Vous êtes sur la page 1sur 206

Progress in Materials Science Vol. 28, pp. 229 to 434, 1984 0079-6425/84 $0.00 + .

50
Printed in Great Britain. All rights reserved Copyright © 1985 Pergamon Press Ltd

PHASE STABILITY U N D E R IRRADIATION

K. C. Russell
Department of Materials Science and Engineering, Department of Nuclear Engineering,
Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, U.S.A.

(Received 21 February 1984)

"Systems requiring a continuous inflow o f energy to maintain their state exhibit


many analogies to equilibrium but their behavior is more strongly influenced by
noise."
Landauer (1978)

ABSTRACT
Irradiation of metals and alloys with neutrons, electrons, heavy ions, or 7-rays may introduce
up to 10s J/mol of energy in the form of atomic displacements. This energy, which is in the form
of vacancies, self-interstitials, and cores of displacement cascades is then available to produce
a range of phase changes and microstructural alterations which are not observed under thermal
conditions. There exist numerous mechanisms to convert part of this displacement energy into
microstructural change, including irradiation-induced solute segregation, Frenkel pair re-
combination at the particle:matrix interface, irradiation disordering or amorphization, and
recoil resolution of atoms from precipitates. In addition, the cores of displacement cascades may
act as precipitate nucleation sites and Frenkel pair recombination may trigger spinodal-like
instabilities.
The theory of these mechanisms is developed in some detail, and is followed by a systematic
review of experimental observations of irradiation-altered phase stability. The observations
include enhanced nucleation on displacement cascades, precipitation induced by solute
segregation to defect sinks, nucleation of wrong phases, disordering and amorphization,
Frenkel pair recombination driven precipitation, and inverse Ostwald ripening.

CONTENTS

1. INTRODUCTION 231
2. GENERAL 232
2.1. Interaction of Radiation with Solids 232
2.2. Displacements 233
2.3. Neutron Irradiation 234
2.4. Heavy Ion Irradiation 236
2.5. Electron Irradiation 237
2.6. The Post-Irradiation State 238
2.7. Enhanced Diffusion 241
2.8. Thermodynamics under Irradiation 243
3. THERMODYNAMICMODELS FOR PHASE STABILITY 244
3.1. Free Energy Calculations 244
3.2. Inclusion of Irradiation-Induced Effects 245
4. POINT DEFECT:PARTICLE ASSOCIATION 246
4.1. Incoherent Particles 247
4.1.1. Particle growth 248
4.1.2. Particle nucleation 251
4.1.3. Nucleation in a single phase region 252
4.2. Coherent Particles 254
JPMS 28/3-4~ A 229
230 PROGRESS IN MATERIALS SCIENCE

5. SOLUTE SEGREGATION 255


5.1. Size Factor 257
5.2. Solute :Defect Binding 257
5.3. Inverse Kirkendall Effect 260
6. SPINODAL INSTABILITIES 262
7. RECOIL RESOLUTION 265
7.1. Infinite Inter-Particle Spacing 266
7.2. Cell Models 267
8. ORDERING AND DISORDERING 270
8.1 Nelson, Hudson and Mazey Model 270
8.2. Liou-Wilkes Model 271
8.3. Banerjee-Urban Model 273
9. PARTICLE COARSENING 274
9.1 Disorder-Dissolution 275
9.2. Recoil Resolution 276
9.3. Point Defect Recombination 277
10. ALUMINUMBASED ALLOYS 278
10.1. AI-Ag and AI-Ag-Zn Alloys 278
10.2. AI-Cu Alloys 279
10.3. AI-Ge Alloys 284
10.4. AI-Li Alloys 286
10.5. AI-Mg and AI-Mg-Si Alloys 293
10.6. AI-Si Alloys 294
10.7. AI-Zn Alloys 295
11. COPPER BASED ALLOYS 299
11.1. Solute Segregation 300
11.2. Cu-Au Alloys 301
11.3. Cu-Be Alloys 302
11.4. Cu-Co Alloys 304
I 1.5. Cu-Fe Alloys 306
11.6. Cu-Ni Alloys 307
12. IRON BASED ALLOYS, FERRITIC 309
12.1. Fe-C Alloys 309
12.2. Fe-Cr Alloys 310
12.3. Fe-Mn Alloys 312
12.4. Fe-N Alloys 312
12.5. Fe-Ni-V Alloys 312
13. IRON BASED ALLOYS, AUSTENITIC 319
13. I. Solute Segregation 319
13.2. Types 316 and 304 Stainless Steel 320
13.2.1. Phases in austenitic stainless steel 320
13.2.2. Temperatures of appearance of phases 324
13.2.3. Microchemical evolution 324
13.2.4. Neutron spectrum and He effects 326
13.3. Precipitation of NbC 339
13.4. Ferrite Formation 342
13.4.1. Proton irradiation 342
13.4.2. Neutron and heavy ion irradiations 343
13.5. Miscellaneous Observations 350
14. NICKEL BASED ALLOYS
351
14.1. Solute Segregation 352
14.2. Ni-Al Alloys
354
14.3. Ni-Be Alloys
358
14.4. Ni-C Alloys
365
14.5. Ni-Cu Alloys
365
14.6. Ni-Ge Alloys
365
14.7. Ni-Mo Alloys
366
14.8. Ni-Nb Alloys
366
14.9. Ni-Si Alloys
371
14.10. Ni-Th Alloys
377
14.11. Ni-Ti Alloys
377
14.12. Nimonic PE-16
377
14.13. Inconel Alloys
378
PHASE STABILITY UNDER IRRADIATION 231

15. RF2RACTORYALLOYS 385


15.1. Mo-Ti Alloys 385
15.2. Mo-Ti-Zr Alloys 385
15.3. Mo-Zr Alloys 385
15.4. Nb-Sn Alloys 386
15.5. Nb-Ti Alloys 386
15.6. Ti-AI, Ti-AI-Mo, Ti-AI-V, Ti-V Alloys 387
15.7. V-B Alloys 388
15.8. V-C Alloys 388
15.9. V-Cr Alloys 393
15.10. V-Fe, V-Mo, V-Nb Alloys 393
15.11. V-Ni Alloys 394
15.12. V-Si Alloys 394
15.13. V-Ti-O Alloys 395
15.14. W-Re Alloys 395
15.15. Zr-AI Alloys 396
15.16. Zr-Nb Alloys 397
16. MISCELLANEOUSALLOYS 398
16.1. Mg-Cd Alloys 398
16.2. Pb-Ni Alloys 411
16.3. Pd-Si Alloys 411
16.4. Pt-C Alloys 411
16.5. Sn Alloys 412
16.6. U-AI Alloys 412
16.7. U-Fe Alloys 413
16.8. U-Mo Alloys 413
16.9. U-Nb Alloys 414
17. SUMMARY 414
ACKNOWLEDGMENTS 419
REFERENCES 419
NOTES1N PROOF 430

1. INTRODUCTION
The study of irradiation damage in metals largely parallels the development of the nuclear
reactor, which has provided most of the technological applications and economic incentive.
Electron and ion beams and ~,-rays have also been employed to produce various features of
neutron damage as well as being interesting in their own right.
Although most of the work has been accomplished in the last twenty-five years, there exist
some very early studies of radiation damage which pre-date not only the availability of
radiation sources but even the discovery of radioactivity itself. Historians of science may
recall that in 1815 Wollaston and Berzelius noticed that the mineral gadolinite displayed some
of the attributes of both crystals and glasses. This was the first report on a group of materials
later classified as metamict (From Latin, "mixed otherwise"). It is now recognized that
localized disorder was introduced into the original crystalline structure by g-particle
bombardment from the naturally occurring uranium and thorium which were consistently
present in concentrations greater than 0.19/o.
In 1943, a decreased lifetime for structural materials experiencing fast neutron bombard-
ment in the just-invented nuclear reactor was predicted by Wigner (1946). Some of the
external symptoms of irradiation damage shown by an otherwise sound metal can be a
discoloration of flaking of the skin, a uniform swelling of the bulk, both increases and
decreases in various aspects of physical strength, and even growths visible on x-ray
photographs. As might be expected, these phenomena became known as "Wigner's disease".
The field of irradiation-altered phase stability was only moderately active until the 1970's
when active development of core materials for the liquid metal fast breeder reactor (LMFBR)
began. Materials in the L M F B R core are damaged at some 103 times greater rates than in
232 PROGRESS IN MATERIALS SCIENCE

earlier reactors. These higher rates give rise to a host of microstructural changes which usually
do not occur in the less severe environment of earlier reactors. These changes include:
Precipitation of phases not thermally stable
Dissolution of thermally stable phases
Disordering of ordered precipitates
Rendering crystalline phases amorphous
Inverse Ostwald ripening
Irreversible segregation of solute to or away from dislocations and interfaces.
In that any change in a carefully tailored metallic microstructure usually causes undesirable
changes in the material properties, effects such as those just listed pose a serious problem in
reactor material development. As a result, many experimental studies have been performed
of microstructural stability under irradiation.
Irradiation-altered phase stability has at the same time become an area of intense scientific
study.
Materials under irradiation are a class of dissipative systems which may reach a steady
state, but not equilibrium (Landauer, 1978). The behavior of this class of systems is also very
sensitive to small changes in environmental and material parameters. Small changes in such
parameters often cause large changes in the behavior of the system.
This monograph begins with a brief review of the interaction of various types of radiation
with metals and the production of a defect structure. Then, the various theoretical models
for microstructural phase stability under irradiation are discussed in turn. Finally, the various
experimental studies of phase stability in metals and alloys under neutron, electron, heavy
ion, and 7-ray irradiation are reviewed on a system by system basis. Wherever possible the
observed effects are attributed to a particular mechanism (or mechanisms). The experimental
observations are summarized in tabular form in each section. Those interested in irradiation-
altered phase stability in ceramic materials are referred to the review of Clinard and Hobbs
(1984).
Irradiation effects which have been the topics of recent comprehensive reviews are discussed
only briefly. These effects include enhanced diffusion (see Adda et al., 1975 and Sizmann,
1978), irradiation disordering (see Schulson, 1979) and the effects of irradiation on super-
conducting properties (see Brown et al., 1978).
The number of displaced atoms per atom (dpa) has become the preferred unit of irradiation
damage, and is used wherever possible. Where authors cite numbers of dpa, their value is
given in the text and in the tables. However, irradiation exposures are given in a number of
other units, including neutrons or ions per unit area, sometimes of unspecified energy, as
energy flux of radiation, and sometimes as degree of burnup of nuclear fuels. Whenever
possible an estimate of number of dpa is given. In the case of neutron exposures, the rough
rule of thumb that 2 x 10 21 fast neutrons/cm2 gives approximately 1 dpa, regardless of alloy,
is used in the absence of a better value. This estimate is good to within about a factor of 2.

2. GENERAL

2.1. Interaction of Radiation with Solids


There are many kinds of interactions between energetic particles and solids. Of interest here
are atomic displacements and transmutation reactions caused by neutrons, heavy ions (Z > 1)
PHASE STABILITY UNDER IRRADIATION 233

protons and electrons. Neutrons are typically produced by fission or fusion reactions, and
charged particles by particle accelerators or fusion reactions. Particle energies are typically
in the MeV range, in comparison to the ca. 25 eV needed for an energetic atom to displace
a lattice atom of comparable mass.
Numerous books have been written on the physics of atomic displacements (Dienes and
Vineyard, 1957; Thompson, 1969). Unfortunately, no volume has appeared which incorpo-
rated recent advances in the field. The following treatment generally follows the simplified,
pedagogical approach of Olander (1976).
An energetic particle will transmit the maximum possible energy to a lattice atom in a
head-on collision. A non-relativistic particle of energy E and mass Mt may transfer to a
stationary particle of mass M a maximum energy of
Tm~x= AE, (2.1)
where
A = 4M1M
(Ml + M) z . (2.2)

The energy transfer will be less for glancing collisions.


Electrons in the MV range must be considered as relativistic particles. The maximum
energy transferred is given by Dienes and Vineyard (1957) as:
2(E + 2mC2)E
Tm~x = MC 2 , (2.3)

where m = electron rest mass


M = mass of atom
C = velocity of light.
The closer M1 and M are to equality, the more efficient is the energy transfer. For M t = M,
Tmax= E, as often occurs in collisions between billiard balls. At the other extreme, electrons
may transfer at most a few percent of their energy to a metallic lattice atom.
The probability of a collision is given in terms of a cross section. Consider a slice of material
of unit area and dx thickness, containing N particles per unit volume. The probability of
collision of an incident particle of energy E with a target particle in dx is:
N ~r(E) dx,
where a ( E ) is the total collision cross section. The mean free path, or distance between
collisions by an energetic particle of energy E is related to the collision cross section by:
1
t'(E) - Ua
(E---~ " (2.4)

2.2. Displacements
A minimum amount of energy must be transferred to displace a lattice atom, creating a
self interstitial and a vacancy. This displacement energy, Ed, depends on the metal or alloy,
the crystal structure, and direction of displacement (e.g., (100) versus (110) versus (111)).
In addition, E d has been found to decrease significantly with increasing temperature (Urban
and Yoshida, 1981). Ed is a difficulty quantity to measure or calculate, and as such is relatively
poorly known. A value of about 25 eV, independent of material and averaged over the various
234 PROGRESS IN M A T E R I A L S SCIENCE

1
0 Ed 2E d Ec
PKA Energy (E)

FIG. 1. Kinchin-Pease model for number o f displaced atoms as a function o f primary knock-on
atom energy. (After Kinchin and Pease, 1955.)

possible directions for displacement is sometimes taken as a reasonable estimate for Ed. This
25 eV is far more than the typically 5 eV energy for thermal formation of self interstitials.
The reason for the difference is the highly irreversible nature of the displacement event, which
must require more energy than the equilibrium thermal creation process.
An atom which has been displaced by a bombarding particle is known as a "primary
knock-on atom", or PKA, which may have an energy from the eV to the keV range. PKAs
with energies below E d cannot displace further atoms and will lose their energy due to various
inelastic interactions which cause only very localized heating of the lattice. The more energetic
PKAs may create one or a number of secondary knock-ons. Each of these secondary
knock-ons, given enough energy, may in turn displace other atoms. The result is a shower
of displaced atoms. The problem is to calculate the total number of atoms displaced in a
particular material by a PKA of a particular energy.
The Kinchin-Pease (1955) model, based on elastic, two-body collisions between atoms
gives a simple, physically reasonable estimate of the total number of displaced atoms due to
a PKA of energy E. Figure 1 plots the number of displaced atoms, v(E) versus E. The plot
of v (E) versus E is seen to have four distinct regions:
(I) E < Ed; v ( E ) = 0. For these low energies, the PKA cannot transfer the minimum
displacement energy, so v ( E ) = O.
(II) Ed < E < 2Ed; V(E) = 1. In this range the PKA can displace a lattice atom, but will
itself fall into the vacancy just created. As such, only one lattice atom is displaced,
which in fact just replaces the PKA.
(III) 2Ed < E < E~, v(E) = E/2E d. In this range, the PKA displaces one atom per 2Ed of
energy.
(IV) Ec < E, v ( E ) = Ec/2Ed. Above a certain cutoff energy, Ec, losses in energy are by
electronic excitation which gives no atomic displacements. Only below Ec, where
Ec -~ MI (KeV) (M l is the mass number), do energy losses become primarily elastic.
It must be emphasized that the Kinchin-Pease model is highly simplified. In particular, the
sharp breaks in v (E) at Ed, 2Ed, and E c do not exist, and Ed should be replaced by some mean
threshold energy averaged over several directions.

2.3. Neutron Irradiation


The total number of atomic displacements produced by a neutron of energy E, is given
by:

Ra= N aa(E.) dp(E.) dE., (2.5)


Ea/A
PHASE STABILITY UNDER IRRADIATION 235

3000 -

2000
los

I~ I000

010"3
,.
10"z 10-= I
J/7 10
Neutr0n enerqy MeV

Fio. 2. Displacement cross section for stainless steel. (After Olander, 1976.)

where the displacement cross section, ad, is

aa(E,) = AE, Ed a (E,, E) v (E) dE. (2.6)


I ^E"
,lEa
In the above:
4)(E,) = flux of neutrons of energy E, (number/m z" s),
a(E,, E) = cross section for transfer of energy E to a PKA by a neutron of energy E,,
v(E), as given in the Kinchin-Pease model, is shown in Fig. 1.
The displacement cross section for stainless steel is shown in Fig. 2. The displacement cross
section is seen to increase with E,, but by no means proportionally to it. The 14 MeV fusion
neutrons have a a d about four times that of the ~ 1 MeV fission neutrons, rather than the
fourteen times which would be the case if E, and a~ were proportional.
Figure 3 shows the neutron spectra for three reactors of interest in irradiation testing. The
Experimental Breeder Reactor II (EBR-II), which has been heavily used in materials testing
work, has a neutron spectrum which peaks between 105 and 106 eV. The High Flux Isotope
Reactor (HFIR) has a high energy spectrum similar to EBR-II but also has a high low energy
and thermal neutron spectrum, which often leads to large numbers of transmutation events.

,0.-\ '~ HFIR-IO0 MWj . , ~ - ,

I013
'~'F:: IOIZ EBR-
,.., 6Z.5 M W t / / '
i= 10 9 7/
//Fusi~
iolO // I MW m"2

iog
• I • • I ,
//
, 1,111 , , i , hi, ,I , ,
10" l I 10 10z 103 104 10~ 106 107
Neutron energy (eV)
FIG. 3. Neutron spectra in three reactors: Experimental Breeder Reactor II (EBR-II), High Flux
Isotope Reactor (HFIR), and at the first wall o f a controlled fusion reactor with a 1 MW/m z power
density. (After/Kulcinski, 1979.)
236 PROGRESS IN M A T E R I A L S S C I E N C E

The fusion reactor first wall sees a neutron spectrum with a high energy peak near 14 MeV,
which again leads to transmutation events.
Neutrons may also enter into transmutation reactions, wherein a neutron is captured by
the atomic nucleus. Such absorption may give various effects including:
(a) Simply raising the atomic mass, which gives an atom unchanged chemically, but which
may respond differently to subsequent neutrons. For example, Ni 5s captures a neutron to
become Ni 59. Neutron capture by Ni 59does not, however, give Ni 6°, but instead produces Fe 56
and He 4.
(b) Transmuting the absorbing atom to a new chemical species one mass unit heavier than
the original atom.
(c) Fission of the nucleus:neutron complex to produce two atomic nuclei. One nucleus is
often a proton or a-particle (hydrogen or helium nucleus, respectively). Transmutation
reactions are an important consideration in phase stability because they change the chemical
composition of the irradiated material.
The various transmutation reactions may be described by an equation similar to Eq. (2.5),
but with ad replaced by a transmutation cross section, at, which is often strongly dependent
on E,. Some reactions occur almost entirely over a limited range of E,. More often, trr
increases with E,, and for many reactions is much greater for 14 MeV fusion neutrons than
for 1 MeV fission neutrons. This variation with E, means that transmutation-induced
composition changes are more likely to be important in fusion than in fission environments.

2.4. Heavy Ion Irradiation


Heavy ion irradiation of metals is of interest from several aspects, including:
(i) composition modification (Pfister, 1975)
(ii) as a scientific tool
(iii) as an aid in understanding the behavior of materials under neutron irradiation
(Steigler, 1975).
Heavy ions for irradiation studies are typically produced in Van de Graaff accelerators or
in cyclotrons, and as such have energies in the range of a few to a few tens of MeV. Such
ions, once inside the metal, behave very much like PKAs. The initial energies are usually far
above Ec so that relatively little displacement damage is done until the ions have slowed to
this level. Most of the displacement damage by heavy ions is in a few hundred nm-thick layer,
typically located somewhat less than 1 #m beneath the sample surface where the beam enters.
Heavy ions used in surface modification may have energies in the tens of keV, and much lower
penetrations.
A heavy ion of initial energy E~ loses energy by electronic stopping according to:

E,(x) = [ ( E , o ) '/~ - ½kx] ~ (2.7)

where E~(x)=energy at point x, and for like ions (as in self-ion irradiation)
k = 0.3 N Z 2/3eV/nm, with N = atoms/unit volume. The rate of production of displaced atoms
at depth x is given by Olander (1976) as:

Ra(x ) = N I a(E~, E) v(E) dE displacement/m3. s. (2.8)


PHASE STABILITY UNDER IRRADIATION 237

80
............ o~ . . . . . . . . . . . . . . . . . .

6o

io

2o

0.5 1,0 15 20 2,5


E (MeV)

FIG. 4. The damage cross section for electrons bombarding copper, taking Ed= 25 eV. (After
Thompson, 1969.)

In Eq. (2.8):
I = current of bombarding ions,
a(E~, E) = cross section for an ion of energy Ei to transfer E to a lattice atom,
v(E) = number of atoms displaced by PKAs of energy E, and may be obtained from a
displacement model, such as that of Kinchin and Pease.

2.5. Electron Irradiation


Electron irradiation is used for much the same reasons as given for heavy ion irradiation,
except for composition modification. In the context of phase stability studies, essentially all
electron irradiations are done in the high energy electron microscope (HVEM). Micro-
structural evolution may then be studied during irradiation.
As noted earlier, in irradiation experiments electron energies are typically in the 1 MeV
range with velocities near the speed of light, so that they must be considered as relativistic
particles. The maximum energy transferred, T~ax, is given by Eq. 2.3.
For 1 MeV electrons /'max is only a few E d. As such, a PKA in electron irradiation will
produce no more than a few displaced atoms.
Figure 4 gives the displacement cross section for electrons bombarding copper, based on
Ed = 25 eV. For high energy electrons ad approaches an asymptotic value given by Thompson
(1969) as:
8na~Z2E:
a d --
MC2E~ '
(2.9)
where a0 = Bohr radius of hydrogen,
E = 13.6 eV, the Rydberg energy,
Z = atomic number of the displaced atom.
The asymptotic value of a d is seen to scale a s ZE/MEdfor various displaced atoms. At lower
electron energies a d has a more complex dependence on Z, M, E, and E d.
Figure 5 shows the penetration and displacement behavior of the various types of particles
used in irradiations; neutrons are seen to have great penetrating power, but displace very few
atoms. Heavy ions are seen to penetrate less than 1/~m into the metal, but to give a great
many displacements per ion. The heavier ions are seen to give the greatest number of
displacements but to have the lowest penetration. Electrons have moderate penetrating power
and by far the lowest efficiency in creating displacements.
238 PROGRESS IN MATERIALS SCIENCE

i0-i4 •

~5"Mev TO+
I0-i51 ~~1.31
5 Ni *- MeV

I0-i61
20-MeV C+

i0-17

,a.> 10-18

<,~ 10-19
lo-zO
~-MeV neutrons
~..~
-~

10-21 ~'"- I- MeV neutrons

iO-ZZ 1.5-MeVe-- ~

10-z3 I I l I I
Z 4 6 8 IO 12
Dislonce inlo solid, y.m

FIG. 5. Penetrating power of various particles impinging on nickel. (After Olander, 1976.)

2.6. The Post-Irradiation State


The preceding section describes the irradiation process up to the instant of displacement
event. We will now consider the residue of irradiation-induced defects and their annealing
and rearrangement into other defect aggregates.
The probability of PKA displacing other atoms increases as it slows down, so that the
displaced atoms become more closely spaced. Energetic higher order knock-ons may also
travel through the lattice displacing atoms at closely spaced intervals. The result is the
root-like structure shown schematically in Fig. 6, known as a displacement cascade. Figure
7 shows the results of a field ion microscope study of displacement cascades in 20 keV W ÷

-- PKA
..... ,Secondary "-..._._I. .., I. ,~.~

• f" ",,..?~ :-~-"+..._


~+.-~-~>"-
-..
,, --,._._/
,' ,' ".",' ..,"--/- ~ ~--.V:
'~, ,~ . . . . . . . " ,-,," --X " ',
o~ "~'-" , ,+ ~ " .... ~l "; i---' ;,
-, >. ?--
I

[100] - - ~
FIG. 6. Proposed structure of displacement cascade caused by a 5 keV PKA in iron. Out-of-plane
damage has been projected onto the (001) plane shown in the figure. (After Beeler, 1966.)
PHASE STABILITY UNDER IRRADIATION 239

QO 0 0 ~. 5A =:

,~~#~ooO
o o

[1251
o o
-\lolo] o zo k,v w*[
~7[,oO)(oo,] ~ ~° IT'll I
FIG. 7. Visualization of depleted zone in W caused by a 20 keV W + ion incident at the angle shown.
Circles connected by lines indicate vacancies on nearest neighbor sites. (After Pramanik and Seidman,
1983).

ion irradiated W. The core of the cascade is seen to contain a large number of vacancies, many
on nearest neighbor sites. A shell approximately 10 nm from the center of the cascade is
enriched with self-interstitials. The size and structure of the displacement cascade will be seen
to have important roles in the destruction or disordering of small precipitates.
As noted earlier, an electron in the MeV range has a Tm,x of a few Ed at the most, and
can at most displace a few atoms. As such, electron irradiation does not give displacement
cascades. The absence of displacement cascades constitutes an important difference between
electron and heavy particle irradiation.
Most experimental studies of phase stability are performed between 0.25 Tm and 0.6 Tm
(Tin = absolute melting point of the material). In this temperature range both the vacancy and
self interstitial are mobile, and may migrate through the crystal lattice. Vacancy:interstitial
pairs within a few atomic spacings of one another are attracted due to elastic interaction.
Because of this binding and the defect mobility, part of the displacements, both isolated and
in cascades, will annihilate almost immediately after their creation. The remaining defects are
free to migrate through the lattice until they:
(i) annihilate through vacancy:interstitial recombination,
(ii) annihilate at a fixed point defect sink, such as a dislocation, grain or interphase
boundary, or free surface,
(iii) aggregate into a more complex defect, such as a void, vacancy or interstitial dislocation
loop, or precipitate particle.
A number of important processes depend on the defect concentrations. There are several
derivations of the steady state concentrations of vacancies and interstitials in irradiated
metals. The following simple treatment is from Brailsford and Bullough (1972):

= K + Kth -- DvCvk~ - otC, Cv, (2.10)


8t
ec,
c3--t = K -- D , C , k 2 - ~C~C~, (2.11)
240 PROGRESS IN MATERIALS SCIENCE

where K = atomic displacement rate,


Kth = rate of thermal vacancy creation,
Cv, C~ = concentrations of vacancies apd interstitials,
Dr, D~ = diffusion coefficients for vacancies and interstitials,
0¢ = recombination coefficient -~ 47r(D~ + D~)
a 2

a = lattice spacing,
k~ = fixed sink strength for vacancies, and
k~ = fixed sink strength for interstitials.
At steady state we have Cv = (~i= 0, and Eqs. (2.10) and (2.11) may be solved to give, for
the atomic fractions of defects:
D,k~
C~ = ~ [ - 1 - p + x/(1 + #)2 _ q]. (2.12)

D,k~
C~ = ~ [ - 1 +/~ + x/(1 + p)2 + q], (2.13)

where
4olK Kthq
,7 = D,Dvk:,k:. ' - 4K "

In the case of dislocations and voids being the only point defect sinks:
k~ = 415cpc + Zvpd, (2.14)
k 2 = 4~Fcp c + Zipa, (2.15)
where Pc = void number density (number/m 3)
I%= mean void radius,
Pd = dislocation line density (m/m3),
Zv, Zi = dislocation sink strength constants for vacancies and interstitials.
Other terms may be added for defect loss at other sinks such as grain boundaries, precipitate
particles, and external surfaces.
Both Zv and Z~ are of the order unity, but Z~ is the larger, due to the larger elastic
interaction between self interstitials and dislocation lines. The absolute and even relative
values of Z~ and Zv are matters of active debate (Bullough and Willis, 1975; Heald and
Speight, 1975). However, Z~/Z~ is thought to lie between about 1.02 and 1.4. As a result, more
interstitials than vacancies arrive at dislocations, while the reverse is true at such so-called
neutral sinks as voids, grain boundaries, and free surfaces.
The defect concentrations take on a particularly simple form in certain special situations.
At low temperatures (T-~ 0.25 T,,), Kth is negligible and defect loss is primarily by direct
recombination. One may use Eqs (2.10) and (2.11) with Kth and losses to fixed sinks equated
to zero, and take C~D~~- C,,D~ to obtain:

K~Di, (2.16)
C , = ~ ~D~

Ci= K~D~. (2.17)


N/ aD,
PHASE STABILITY UNDER IRRADIATION 241

In the limit of D~---~0, steady state will in fact never be established and the preceding analysis
does not supply.
At moderate temperatures Kth is still negligible and defect loss is mostly to fixed sinks, so
that:
K
C~, - D~k~' (2.18)

K
Ci = Dike (2.19)

At still higher temperatures (T ~0.6 Tm) the irradiation-induced defects anneal out very
rapidly and, in addition, Kth>>K. As such, the vacancy concentration is the thermal
equilibrium value and the self interstitial concentration is so much smaller that it may be set
equal to zero.
The preceding derivation for reasons of space and simplicity ignored such important
considerations as multiple sinks, size dependent defect capture efficiencies, spatial arrange-
ment of sinks, free surfaces, impurity trapping of point defects, and solute segregation. Those
interested in a more complete treatment are referred to the critical review of Mansur (1978).

2.7. Enhanced Diffusion


Atom movement in substitutional alloys usually occurs by the interchange of position
between atoms and adjacent vacant sites. The diffusion coefficient in cubic structures is given
by: D = Fd2/6, where d = atomic jump distance, and F = atomic jump frequency for
uncorrelated jumps. For diffusion by the vacancy mechanism, F equals the product of the
vacancy concentration, Cv, the vacancy jump frequency, Fv, and a correlation factor,fo, which
is nearly unity for pure metals and accounts for the fact that the directions of successive jumps
of an atom may be correlated because they may involve the same vacancy. Thus, F =f~CvF~.
The diffusion coefficient for vacancies, Dv = F~d:/6, may be used to express the atomic
diffusion coefficient as
D =f~C~Dv. (2.20)
Diffusion under irradiation may also take place by the motion of self-interstitials which
are not present during equilibrium conditions. Divacancies or other defect aggregates may
also make some contribution to atomic motion. In general:

D =f~DoC~ +fDiCi +f2vDzvC2v + . . . . . (2.21)

Irradiation enhancement of vacancy, interstitial, or divacancy concentrations will then give


a corresponding increase in D (Sizmann, 1978; Adda and co-workers, 1975). Binding energies
for divacancies and higher order aggregates are seldom known, so it is common practice to
consider only the contributions of single vacancies and self interstitials to the diffusion
coefficient.
The irradiation-enhanced diffusion coefficient takes on one of three forms depending on
the value of the temperature with respect to three temperatures ranges. At high temperatures
defect mobilities are so high that the irradiation induced defects are lost quickly and the
thermal equilibrium vacancy concentration is large compared to the irradiation induced
concentration. In this regime D is unaltered by irradiation and has an activation energy of
E~ + E~, the sum of the energies of vacancy formation and migration respectively. In the
242 P R O G R E S S IN MATERIALS SCIENCE

intermediate temperature range most of the vacancies and interstitials will be lost to fixed
sinks, which leads to
K K
DiCi=k~, and DvC~-~.

The atomic diffusion coefficient is then:


ki + v 2K
D~- K ~kikv ~2' (2.22)

The diffusion coefficient is nearly independent of temperature (varying only as the dislocation
density), and has been found to so behave in a number of cases (Adda et aL, 1975; Hudson,
1978).
At still lower temperatures, the concentration of irradiation produced vacancies becomes
high enough that direct recombination with interstitials begins to dominate over diffusion to
fixed sinks. The rate equations reduce to:

D,C, ~- DvCv ".. /D,D,K. (2.23)

In the case where Di>>D~ we may take ~ = 4~Dja 2 to obtain:


,~ ~/~--~v a 2
D _ X/ ' (2.24)

These three regimes are shown schematically in Fig. 8 for three damage rates. At high
temperatures the rate of thermal vacancy production is so high that the irradiation-induced
defects contribute negligibly to the diffusion coefficient. At intermediate temperatures the
irradiation-induced defects annihilate at dislocations so that the number of jumps, hence the
amount of diffusion, and the diffusion coefficient are nearly independent of temperature.
At the lowest temperatures the irradiation-induced vacancies have only limited mobility and
are present at high concentration. Defect annihilation is by direct vacancy:interstitial

Q=0 /.HIGH K

¢3
g
.J

LOW K /

- -I .--...
T
FIG. 8. Schematic plot of different regimes for diffusion coefficient in irradiated metal at different
displacement rates.
PHASE STABILITY UNDER IRRADIATION 243

recombination and the diffusion coefficient is given by Eq. (2.24) and the activation energy
is E~/2.
The shifts in the activation plots with K come from higher displacement rates giving greater
defect concentrations. The low temperature (direct recombination) and intermediate tem-
perature (fixed sink) portions of the activation plots are thus elevated and all intersections
are shifted to higher temperatures.
Those interested in a more detailed analysis of irradiation-enhanced diffusion are referred
to the recent comprehensive treatment by Sizmann (1978).

2.8. Thermodynamics under Irradiation


One might expect that the irradiation flux, q~, incident on a metal could be treated as an
intensive thermodynamic variable (as are temperature, pressure, chemical potential, and
magnetic fields). Then, a modified free energy could be derived to dictate a true equilibrium
state under irradiation. It has been asserted that simply adding the contributions of the
irradiation-induced point defects to the free energies of various actual and potential phases
will account for the effects of irradiation. Then, whatever combination of phases gives the
minimum free energy will be the final result.
Both expectations are in fact ad hoc, as irradiated metals violate a number of conditions
for equilibrium. First, minimization of free energy applies only to a system which is isolated,
except for interactions with thermal and mass reservoirs (Gibbs, 1960). An irradiated metal
is not an isolated system, but is subjected to a flux of energetic particles across its boundary.
It would be hard to ascribe a chemical potential to these particles, and their temperatures
(with energies typically in the MeV range) are orders of magnitude higher than that of the
irradiated metal.
An isolated system is obtained only by encompassing the nuclear reactor or other
irradiation source with the material being irradiated. Thermodynamics then govern the final
state of the system--but not the path or time to reach equilibrium or intermediate steady
states, should such exist.
Metals under irradiation (even at steady state) also contain concentration gradients in point
defects, and often in solutes, due to solute segregation. These gradients violate the
thermodynamic requirement that at equilibrium the chemical potential of each species be
constant throughout the system. The fundamental statistical mechanical principle of detailed
balancing, which states that at equilibrium every microscopic process and its inverse occur
at the same rate, is also violated. In irradiated metals, point defects are created by
displacement events and diffuse to sinks, with no reverse flux. This is a gross violation of
detailed balancing, which according to Fowler (1966) is equivalent to "preservation of
equilibrium".
It should be clear from the foregoing that irradiated metals fall in the regime of irreversible
thermodynamics, and may reach a steady state but not equilibrium (Landauer, 1978; Haken,
1975). Such systems may be approached by one of two methods. The first is the continuum
approach (Prigogine, 1967; Denbigh, 1951) which appeals to a synthetic optimization
principle such as minimum entropy production to dictate the steady-state configuration of
the system. Such schemes often give wrong answers (Cahn and Mullins, 1962). The second
approach is physically based, and usually atomistic, whereby one considers the kinetics of
the various processes involving irradiation-induced defects and the rate laws governing them,
and attempts thereby to determine the final state of the system, and possibly the rate and
path of approach to this state as well. This is the approach which has been particularly fruitful
244 PROGRESS IN MATERIALS SCIENCE

in radiation damage work to date and is the one which is followed in most of the theoretical
analyses of phase stability under irradiation presented herein.

3. THERMODYNAMICMODELS FOR PHASE STABILITY


A zeroth order approach to prediction of phase stability under irradiation would be to
assume that the equilibrium phases will finally appear but with the kinetics of nucleation,
growth and coarsening speeded by irradiation-enhanced diffusion, as described in Section 2.
The next approximation would be to modify the free energies of the various potential phases
for the presence of irradiation-induced defects, and then assume that the final microstructure
would be composed of the phase or combination of phases giving the lowest free energy.
As discussed earlier, such thermodynamic arguments are of necessity ad hoc in a system
under irradiation. However, several investigators have taken this general approach, with
various levels of sophistication and approximation. In support of this approach, Gittus and
Watkins (1977) have shown that void swelling (see Nelson, 1975; Corbett and Ianniello,
1972), which is often associated with microstructural instability, increases in Fe-Ni-Cr
alloys as the FCC matrix phase becomes increasingly less stable against the FCC--,BCC
transformation.

3.1. Free Energy Calculations


A starting point for thermodynamic arguments is a knowledge of how much the stability
of the relevant phases would have to be changed to cause appreciable deviations from the
non-irradiated condition. Clearly, very strong irradiation effects would be needed to reverse
the stability of two phases differing in free energy by tens of kJ/mol, whereas much more
subtle effects would be important if the phases were within a few tens or hundreds of J/mol
of one another in free energy.
Gittus and Miodownik (1979) calculated the free energies needed to give precipitation of
Ni3Si from an undersaturated, binary, Ni-Si solid solution. The calculated free energies
ranged from tens of J/mol at low temperatures and high silicon concentrations to kJ/mol
under the opposite conditions. Accordingly, irradiation must somehow supply free energies
at this level to induce precipitation in these undersaturated solid solutions.
Kaufman et al. (1977) performed a series of more extensive and elaborate free energy
calculations based on solid solution models which have been used extensively to calculate
phase diagrams. These calculations were performed for the W-Re and Fe-Ni-Cr systems,
both of which are of interest in experimental studies of phase stability.
In the case of W-Re, there is evidence that thermally single phase W-rich alloys form the
(WRea) phase under irradiation (R. K. Williams et al., 1983). In addition to the alloys being
thermally single phase, the Re-rich X phase forms rather than the nearest phase, the
equiatomic tr phase. An over 13 kJ/mol free energy decrease for the Z phase was found
necessary to give its precipitation and make it more stable than the a phase. This is a large
free energy change, which would take powerful irradiation effects to produce.
The Fe-Ni-Cr system is of great technological interest in that it is the basis for alloys used
for breeder reactor cladding and ducts and for many of the alloys proposed most seriously
for first walls of controlled thermonuclear reactors. The a phase forms readily in type 316
stainless steel thermally reacted in the temperature range below 700°C, but often does not
appear under irradiation (Lee et al,, 1981). Kaufman et al. therefore calculated the free
energy shifts which would cause the tr phase to become unstable in the alloy. The required
PHASE STABILITY UNDER IRRADIATION 245

free energy modifications are not as simple to present as in the binary W-Re case, in that
the calculations are based on a complex ternary solid solution model. However, a change in
the appropriate solution model energetic parameters of 2-6 kJ/mol was found necessary to
suppress tr phase formation.

3.2. Inclusion of Irradiation-Induced Defects


Several authors have proposed that some irradiation-induced changes in phase stability
may be accounted for by changes in the relative free energies of the phases due to the excess
point defect concentrations: Hauser and Schenk (1967) first made this proposal on a
qualitative basis, and Wilkes et al. (1976), Bocquet and Martin (1979) and Yamauchi et al.
(1979) later presented more quantitative treatments. Even though self interstitials have
formation energies several times higher than do vacancies, the vacancies are usually present
in orders of magnitude greater concentrations (see Section 2), and thus make the greater
energetic contribution.
Wilkes et al. (1976) estimated the contribution of the excess vacancies as CvE~ where
E~=formation energy. Taking Cv= 10 -4 (an upper limit in irradiated metals) and
E~ = 100 kJ/mol, the free energy change would be 10 J/mol, a fairly small change. The authors
noted that they had ignored entropic effects, so that the actual contribution would be even
less. Furthermore, Cv = 10 - 4 is a very high vacancy concentration, so that in most cases the
excess point defect contribution would be less than 1 J/mol, a usually trivial free energy
difference.
Bocquet and Martin (1979) and subsequently Yamauchi et al. (1979) presented substan-
tially more sophisticated treatments of the effects of excess point defects on thermodynamics
and phase stability in both clustering and ordering systems. Bocquet and Martin treated their
system as a ternary alloy (A, B, C), and developed the statistical mechanics by the Kikuchi
cluster variation model. The Kikuchi treatment is a more sophisticated varsion of the familiar
pairwise interaction model. Rather than considering only pairs of atoms, Kikuchi takes a
regular tetrahedron as the basic unit, with atoms at the corners.
Bocquet and Martin did their analysis in terms of the energetic parameter, W~j:

.W~j= E~j-½(Eii-t- Ejj) (3.1)

where Eij is the nearest neighbor pair energy between species i and j, where i, j are A, B, or
C. When WAB< 0, like atoms attract and the system tends to cluster, while for WAB> 0 like
atoms repel and the system tends to order. Since the Wij are seldom known, they were used
as parameters, and assigned potentially interesting but physically reasonable values. Bocquet
and Martin found 22 different potential ternary (A, B, C) phase diagrams. These were
classified on the basis of the possible combinations of positive and negative WAC and WBc,
and of which component constituted the defect. A search was then made for cases where:
(i) the defect concentration required to produce precipitation is on the order of those
encountered in irradiated metals; and
(ii) the range of solute concentrations over which such precipitation occurs is great enough
to be observable.
When the binary (A, B) shows ordering, for precipitation to be induced, one must also have
WBc < 0. This inequality implies that defects in pure B also order, something which the
authors note has never been observed. In addition, the defect-induced precipitate contains
J . P M . S 28/3-4~ B
246 P R O G R E S S IN MATERIALS SCIENCE

the order of 20~ defects. Such defect-compensated deviations from stoichiometry have been
observed, but not in irradiation-induced precipitates.
In a clustering system (WAB> 0) a necessary condition for precipitation was that the
temperature and composition be near the top of the miscibility gap in the AB binary. The
authors note that in the previously observed case of irradiation-induced precipitation of a
disordered phase (A1-Zn) the temperature is far from the critical temperature.
Yamauchi et al. (1979) took an approach somewhat similar to Bocquet and Martin, but
(1) used the simpler, Bragg-Williams approximation for the free energy, and (2) assumed a
constrained equilibrium in which the defect concentrations of the various phases were pre-set,
rather than being established by creation and annealing processes.
The authors considered two cases:
(i) a clustering system in which two different disordered phases, ~ and ~', coexist, and
(ii) an ordering system in which a disordered phase, ~, coexists with an ordered phase, r,
of the A3B structure.
Substantial shifts in the phase boundary were found for an ordering system, as seen in Fig.
9 where the energetic parameters are those for the Ni-Si system. It is the disproportionation
of vacancies between the two phases which gives shifts in the phase boundaries.
Phase boundaries calculated for a clustering system containing excess vacancies also
(Fig. 10) show substantial shifts. The calculated shifts are consistent with the ,--65 K shifts
observed by Cauvin and Martin (1979) in A1-Zn alloys. Again, disproportionation of
vacancies between the two phases causes the substantial shift.

4. POINT DEFECT: PARTICLE ASSOCIATION


Several analyses have been made of the effects of excess vacancies and interstitials
co-precipitating with solute atoms. In the case of incoherent particles, the point defects are
annihilated at the precipitate:matrix interface. In the case of coherent particles, the defects

2.0 I I
1 PHASE~UNDARY UNDER IRRADIMION
. . . . . PHASE ~UN~RY WITHOUTIRRADIATION

_~. ,.5 - "a~

,.o -

o I
0 IO 20
.,o -2)

FIG. 9. Effect of irradiation induced vacancies on the phase fields in an ordering system. (After
Yamauchi et al., 1979.)
PHASE STABILITY UNDER I R R A D I A T I O N 247

0 20 40 60 80 I00
a (x tO'2l
CV

FIG. 10. Effect of irradiation induced vacancies on phase fields in a clustering system. (After
Yamauchi et at., 1979.)

are trapped at the interface. Significant alterations in phase stability are predicted for both
cases.
These analyses are basically kinetic in nature in that the effect of irradiation-induced defects
on the stability of particles is approached from the rates at which these and other atomic
species add onto and otherwise interact with the precipitates. Such analytical treatments are
in sharp contrast to those of the preceding section, where only the thermodynamic properties
of defects were considered.

4.1. Incoherent Particles


The Maydet and Russell (1977) analysis assumes precipitates to be spherical and to be
incoherent with the matrix to provide interfaces which are good sinks for vacancies and
interstitials. The matrix and precipitate are binary substitutional solutions dilute in solute and
in solvent, respectively. The precipitates are characterized by two variables: the number of
solute atoms (x) and the number of excess vacancies (n). This latter is the difference between
the number of matrix atoms displaced by the precipitate and x. A precipitate with a greater
atomic volume than the matrix would thus have n > 0 to relieve the strain energy, and an
undersized precipitate would have n < 0. The behavior of a precipitate particle may then be
described by its movement in a phase space of coordinates n and x.
The processes giving rise to motion are shown in Fig. 11. A particle moves in the x direction
by transfer of solute atoms across the incoherent matrix: precipitate interface. The arrival of
vacancies or self-interstitials at the particle or loss of vacancies gives movement in the n
direction. The rate of mass transfer due to irradiation resolution is initially ignored. The
particle moves with a velocity equal to the frequency of addition times the jump distance--in
this case unity.
Yc = fix(n, x ) - ctx(n, x ) (4.1)
h = fir(n, x ) - ~tv(n, x ) - fii(n, x )
where fix(n, x), fir(n, x ) and fi;(n, x) are the arrival rates of solute, vacancies, and interstitials,
respectively. ~tx(n, x ) and 0tv(n,x) are the rates of loss of solute and vacancies respectively.
The fi's are determined by the concentrations and mobilities of the respective point defects.
248 P R O G R E S S IN M A T E R I A L S SCIENCE

,~¥

o , -_ _B,

x (ATOMS)

FIG. 11. Phase space for irradiation-altered phase stability showing processes which give changes
in the particle.

The rates of emission of vacancies and solute atoms are determined by assuming that these
emission rates are unaffected by the presence of the non-equilibrium irradiation-induced
interstitials. Then:

( 1 dAG°']'~ (4.2)
: ~ = f l x ( 1 - e x P \ k - T ~x ]/1'

h=fl~ l-~-exPkk-T ~nn ] ] " (4.3)

The free energy change on forming a precipitate particle from a solid solution super-
saturated with solute and vacancies was calculated from the capillarity model as:
AG o = - x k T In Sx - n k T In Sv + A k T x 2/3 - B k T x ( 6 - n / x ) 2 (4.4)
where Sx = ratio of the actual and saturation concentrations of solute,
sv = ratio of actual and saturation concentrations of vacancies,
A = (367~f~2)l/a(y/kT),
B= t a E / 9 k T ( 1 - v),
f~ = atomic volume of the precipitate,
6= (~ - f~)/tam + ~,
= fraction of atoms in precipitate which come from interstitial sites in the matrix,
Y= particle: matrix surface energy,
E = Young's modulus,
v = Poisson's ratio.
Matrix and precipitate are assumed to have the same elastic constants.

4.1.1. Particle g r o w t h
Nodal lines are obtained by setting ~ and h individually equal to zero as is done
schematically in Fig. 12 for a system in which 6 > 0. The arrows show the direction of motion
of particles at various points in the plane. Any particle on an ~ nodal line may have a velocity
only in the n direction, and vice versa. If the nodal lines intersect (as at n*, x* in Fig. 12),
PHASE STABILITY UNDER IRRADIATION 249

~ O i 0

0 Xt~ x'-'~

FIG. 12. Nodal lines for a particle subject to the processes shown in Fig. 11. The particle is
immobilized at the critical point; particle trajectories are shown as arrows.

the point of intersection is known as a critical point. The particle has zero net velocity at a
critical point.

For h = 0: n = x{6 + (1/2B) In [S~(1 - fl,/flv)]}. (4.5)


For .~ = 0: n = x6{1 + 2A/3Bf2x 1/3 - (1/B6 2) In Sx} ~/2. (4.6)

The equations for the nodal lines are seen to be independent of the kinetic parameters--
other than ~/fl,, which being equal to Z,/Z~ is actually energetic in nature. Further, Sv and
fl~/flv appear only in the equation for h = 0 and Sx appears only in that for ~ = 0.
Simultaneous solution of Eqs (4.5) and (4.6) show that the critical point is located at a x*
given by
x* = - 327r7T~2/3(A4 )3 (4.7)
or a radius of
r* = -2yfl/At~, (4.8)
where A~b is an irradiation-modified potential given by

Adp = - k T In Sx[S~(1 - fli/flv)]~ - [kT In S,(1 - ~/~)]2/4B. (4.9)


If A~b > 0, the nodal lines do not intersect (for x > 0) and all particles will eventually decay.
In the absence of excess defects Aq~ = - k T In Sx and the familiar Gibbs-Thomson equation
is recovered from Eq. (4.9).
In general, particles with x > x* will grow, and those with x < x* will decay; in each case
the path lies somewhere between the nodal lines. Particles not between the lines will approach
them, curving in to cross in the appropriate direction. A particle which is located at (or very
near) the node will be immobilized at that point. (Immobilization is in the absence of
statistical fluctuations, which are not included in the model.) Such a particle resembles the
critical nucleus in nucleation theory in that it is in a zero growth rate situation. These various
trajectories are illustrated by particle flow lines in Fig. 12.
Figure 13 is a schematic illustration of the effect of irradiation on the stability of two phases
in a matrix of ~c The 0 phase is stable under thermal conditions, but having 6 < 0, is
destabilized by irradiation. Irradiation has the reverse effect on the ~Ophase, which would not
250 P R O G R E S S IN MATERIALS SCIENCE

I ~ ++ ,go, ++ +

/ , ,++,o>o
~',.. <0

++ / . . ~ +o<o

- - Thermot

. . . . . Under Irrod,ohon

FIG. 13. Schematic diagram showing the effect of irradiation induced defects on the stability of
precipitates which are oversized (¢) and undersized (0) with respect to the matrix.

appear thermally in ~ of composition C~. However, 6 > 0 for the ~k phase, which is stabilized
by irradiation to the extent that it should precipitate in favor of the thermally stable 0 phase.
Figure 13 is, however, only schematic, as the relationships illustrated have not been
experimentally verified.
Figures 14 and 15 show the results of calculations made for the stability of a Cr203
dispersoid in type 316 stainless steel. It was assumed that oxygen forms an interstitial solution
in austenitic steel, so that ~ > 0. Figure 14 shows the variation of x* with temperature, for
several values of solute supersaturation. Irradiation-induced vacancies are seen to stabilize
particles even for under-saturated (Sx = 0.5) or just saturated (Sx = 1) solutions (Fig. 14). A
stable particle size for Sx = 0.5 or S~ = 1 is seen to exist only up to about T/Tm = 0.55. Above
this temperature range S~---,1 and the stabilizing effect of the irradiation-induced vacancies
is lost. Figure 15 shows growth trajectories of particles for several solute supersaturations.
Particles are seen to grow even for S~ ~< 1, although not as rapidly as when Sx > 1.

8 =1.0574
K= 10-6 dpo/s
II
I0 316 SS+Cr203ppt Sx~//
9
8
7 ¸

ta 6
o
" 5
4

I l I i I I I I
.?..5.30.35.40.45.50.55.60
T/T M
FIG. 14. The effect of irradiation induced defects on the minimum size of stable Cr203 particles in
austenitic stainless steel.
PHASE STABILITY UNDER IRRADIATION 251

316SS +Cr203 ppt. K; 10"8dpo/s


T= 0.35 TM Irradiation Time =105 hr
8 • 1.037
I0
=10
9
5G 8
Z

u 7

~6
~5
45 / x=2

- .

L [ I l I I I I
2 3 4 5 6 7 8 9
X (xlO 4) (MOLECULES)

FIG. 15. Growth of Cr203 particles in irradiated stainless steel which is thermally undersaturated
(S x < 1) or oversaturated (S x > 1) in solute.

Irradiation may have a corresponding destabilizing effect on particles when 6 < 0. The
destabilizing effect is greatest at low temperatures where Sv is highest.

4.1.2. Particle nucleation


The previous discussion concerned the growth or decay of existing particles. Irradiation-
induced defects may also have a strong effect on the nucleation of incoherent particles, where
the critical nucleus is unlikely to contain more than a few tens of atoms.
The key quantity in the nucleation rate calculation is p(n, x) = the number of particles of
n excess vacancies and x solute atoms, per unit volume of material. The time rate of change
of p(n, x) is obtained from a balance between particles entering and leaving a given (n, x)
size class.
The calculation focuses on the number of nuclei growing past a certain value of x, per unit
volume and time. At steady state, this nucleation flux is independent of x (except for the
effects of K*, the rate of particle destruction by displacement cascades) so the count of nuclei
may be made at any convenient value of x. The nucleation flux of particles is given by Mruzik
and Russell (1978) as:

Js(x) = ~ [fix(n,x)p(n, x) - ctx(n, x)p(n, x + 1)]. (4.10)


n= --oo

The summation in Eq. (4.10) converges rapidly for large positive or negative n, due to strain
energy effects.
This theoretical analysis paralleled an experimental study by Bertram et al. (1978) of the
effect of irradiation on the nucleation of Ge precipitates from a dilute A1-Ge solid solution.
Accordingly, the material and environmental parameters chosen for evaluation of the theory
were those which corresponded to the experimental work.
252 PROGRESS IN MATERIALS SCIENCE

IOZ°~ , , ,

!
lOi°~
-~ I~ ~ S~ = I 0 0

-I0 ~ °v = u
I0 ~ S x = I00
Sv = 102

,620 -102 I I
0 .25 .50 .75 1.00
13i / Bv
FIG. 16. Effect of irradiation on the nucleation rate of Ge particles from AI-Ge solid solutions.
The presence of self interstitials reduces the huge enhancements due to vacancies only slightly.

The system of simultaneous differential equations for p(n,x) was solved by numerical
computation, although serious problems were encountered with convergence and stability.
Initial calculations showed that supersaturated vacancies alone have a profound effect on the
precipitate nucleation rate, with up to 60 orders of magnitude increases seen at rather modest
vacancy supersaturations. Furthermore, no reasonable values of solute supersaturation and
interfacial energy would give an observable rate of nucleation in the absence of a vacancy
supersaturation.
Figure 16 shows nucleation rates calculated for precipitation of elemental Ge from A1-Ge
solid solutions as a function of fl~lflv for several solute supersaturations. The presence of excess
interstitials (fl~/flv > 0) is seen to reduce the nucleation rate, but only slightly as compared to
the some 60 orders of magnitude enhancement by the excess vacancies alone.
Nucleus destruction by displacement cascades was found to have an important effect only
for destruction rates so high as to be physically unreasonable.

4.1.3. Nucleation in a single phase region


The detailed analysis of Mruzik and Russell (1978) found that for the conditions studied,
the presence of excess interstitials and destruction of nuclei by displacement cascades had only
a modest retarding effect on the nucleation rate. Then, to a first approximation the free energy
of nucleation may be calculated from Eq. (4.4) as the value of AG ° giving a maximum with
respect to x and a minimum with respect to n. The significance of AG* is that nucleation is
typically relatively easy and rapid for AG*/kT ~ 60 and is too slow to observe for
AG*/kT ~ 60. It should be noted that use of A~b (Eq. (4.9)) as a driving force for nucleation
gives a grossly erroneous result. The function A~b is not a true irradiation-modified free
energy.
Calculations were performed by Best (1984) for alloys of iron with major amounts of nickel
and small amounts of carbon, a system selected in anticipation of experimental phase stability
studies.
PHASE STABILITY UNDER IRRADIATION 253

Single Phase //j


I00 --- Two Phase//

/I' ~ %
,/,f--- 0T
I0 '7 wt%C
I-.-

/;;, .¢ --0.05
~-0.1

K= I0-6 dpa/s
Graphite . ~-,,'~'¢~~
9'~ : 0,5 J/m 2

I I I I I I
I00 200 300 400 500 600 700 800
T (*C)
FIG. 17. The effect of irradiation on the calculated activation energy for nucleation of F%C and
graphite from Fe--35%Ni~C alloys. The atomic displacement rate is K = 10-6/s.

Figures 17 and 18 show calculations performed for conditions of neutron irradiation


(K = 10-6/s) and of heavy ion irradiation (K = 1 0 - 3 / s ) both for the stable graphite phase and
for the metastable cementite phase. The graphite phase is predicted to nucleate much more
readily under all circumstances. The ease of nucleation is due almost entirely to the stronger
effect of vacancies, and is only slightly due to the greater thermodynamic stability of graphite.
Since carbon is in an interstitial position in the matrix, and the atomic volume of the matrix
and graphite are about the same, 6 ~ 1 and excess vacancies are very effective in promoting

Sinqle Phase //~


---Two Phase / // I

wt%c
oo--7-
. O.O5
I0 c^ r .,~E~_(-" / 1/
f¢3~' ~.~ / //
,.---" ///_
//,, W,~oC
/-Z¢-o-~-
///---~ o o5
!- ~../,~- Qi

Graphile ~ ~.~.~.~~ ~
:_.===--~.~.~"' K = I0-~ dpa/s
T~ = 0.5 d/m

I I I I I I
I00 200 300 400 500 600 700 BOO
T (%)
FIG. 18. As in Fig. 17 except at a displacement rate of K = 10-3/s.
254 PROGRESS IN MATERIALS SCIENCE

nucleation. The Fe3C cementite phase is mostly iron, so that 6 ~ 0. I and excess vacancies have
a much less powerful catalyzing effect.
Figure 18 shows that heavy ion irradiations may give nucleation of either graphite or
cementite at carbon concentrations well below the solubility limit. Under neutron irradiation
conditions, mostly graphite is expected to form at carbon concentrations significantly below
the solubility limit.
It must be emphasized that these calculations and predictions are based on the absence of
the other mechanisms of irradiation-altered phase stability. Such effects as solute segregation,
recoil resolution, and disorder dissolution could easily have a large influence--positive or
negative--on nucleation processes and must ultimately be included in calculations such as
the foregoing.
4.2. Coherent Particles
Cauvin and Martin (1981a,b) used a formalism similar to that of Russell (1971) and
Maydet and Russell (1977) to develop a theory for the effect of excess vacancies and
interstitials on the nucleation of coherent particles. In the coherent case, point defects are
trapped at the particle: matrix interface, as opposed to being absorbed. There then exists a
distribution of coherent clusters containing, for example, x solute atoms and having trapped
various numbers (n) of vacancies or interstitials. A negative n indicates that interstitials are
trapped. Interstitials and vacancies were not allowed to be trapped simultaneously.
The assumption of equilibrium defect trapping at the particle:matrix interface reduces the
problem from one of two dimensions (n,x) to one dimension (x). In the one-dimensional case
it is always possible (Russell, 1971) to define a pseudo-equilibrium distribution p'(x), which
gives zero flux of clusters along the x coordinate.
The pseudo-equilibrium distribution is related to p°(x), the equilibrium distribution in the
absence of irradiation-induced defects, by:
p'(x) p°(x)
B(x), (4.11)
p'(x + 1) p°(x + 1)
where, using the notation of Urban and Martin (1982),
fl,,(x)[p~(x) + p,(x)] + fl~(x)[p,(x) + pn(x)]
B(x) - (4.12)
[fl,,(x) + fli(x )] pt(x )
As before, fly(x) and fl~(x) are the rates of vacancy and interstitial capture by a cluster of x
atoms and
p,(x) = number of clusters which have trapped excess interstitials,
pv(x) = number of clusters which have trapped excess vacancies,
p,(x) = number of clusters which have trapped neither defect,
p,(x ) = p~(x ) + pn(x ).
p,(x) +

Physically, B(x) is the fraction of solute impingements on precipitates which does not result
in defect annihilation since the particle is either neutral or has trapped the same type of defect
which enabled the solute atom to diffuse. In addition, the thermal and irradiation-altered
solubilities are connected by:
fir r = Ce n ( ~ )
A pseudo-free energy, G'(x), is then defined in terms of p'(x):
p'(x) (l
p'(x + 1) = exp \k-T ~x /" (4.13)
PHASE STABILITY UNDER IRRADIATION 255

t
eva

X~

FIG. 19. Del~ndence of the stability criterion, B, on the coherent particle size, x. (After Cauvin
and Martin, 1981.)

Just as an ordinary solid solution is metastable if G ( x ) shows a maximum with x, the


metastability criterion for the irradiated solution is that G'(x) show a maximum, or
alternatively, that 8 G ' ( x ) / S x becomes negative for x > x*. Equations (4.11) and (4.13) may
be rewritten as:

8G'(x) _ SAG °
- - + kT In B ( x ) (4.14)
8x 8x
where p ° ( x ) = exp ( - G ° ( x ) / k T ) . The function B ( x ) shows one of two characteristic forms,
depending on the sign of:

t7 = (1 -- Pi/Pv) (4.15)
(1 - B,//L)
where Pi, Pv = probability of a particular trapping site being occupied by an interstitial or
vacancy, respectively. These two cases are shown in Fig. 19. In each case B(0) = l, and B(oo)
approaches a constant value. Similarly:

SAG °
- - = -kTlnSx (4.16)
8x
for large x.
In the case o f ~ < 0 (Fig. 20a) the solution will be metastable if - k T In B(oo) > - k T In Sx.
If the inequality is satisfied there is a critical nucleus at a size x*. In the case of tT > 0,
8 G ' ( x ) / S x = 0 may have zero, one, or two roots (Fig. 20b, c). In the case of zero roots, the
solid solution is stable. The case of one root gives a single maximum and the solution is
metastable. In the case of two roots, the solution is metastable but only to fluctuations up
to x**. Fluctuations beyond x** are unstable and will decay. The solid solution would thus
contain embryos of size x**, but these could never develop into macroscopic particles.
It should be noted that in this model irradiation always decreases solid solubility. In that
many important precipitate phases are coherent with the matrix, the Cauvin and Martin
model is expected to be of fairly broad applicability.

5. SOLUTE SEGREGATION
The phenomenon of solute segregation, first proposed by Anthony (1972), is a vivid
illustration of the irreversible nature of the irradiated state and the linkage of displacement
256 PROGRESS IN MATERIALS SCIENCE

(a)l

aAG__2
o-<0

-kT~nB
X* (b)

if>0

-kTtnB~

X* (c)
SAG*
0">0

X~ X*~
×---~

FIG. 20. Schematic illustration of stability altered by defect recombination at the particle:matrix
interface. The parameter ~ (Eq. 4.15) plays an important role in determiningsolutionstability.(After
Cauvin and Martin, 1981a.)

energy to microstructural and microchemical change. Irradiation-induced solute segregation


may superficially resemble equilibrium adsorption but is in response to point defect fluxes,
rather than thermodynamic considerations.
Under irradiation there are net fluxes of vacancies and interstitials to the various
sinks---dislocations, internal interfaces, and free surfaces. If the composition of the flux of
vacancies or interstitials differs from that of the matrix--the usual case--irradiation-induced
solute segregation will occur.
Solute segregation has a number of important consequences. Precipitates may form in
locally enriched areas even though the precipitated phase is unstable at the average matrix
composition. Conversely, precipitates which are stable at the average matrix composition may
dissolve in depleted regions. Non-equilibrium surface compositions may alter the activation
energy for the nucleation of voids or precipitate phases. In addition, the ability of a void
surface to accept point defects depends on its composition. A local alteration in surface
composition may give significant changes in the rates of void nucleation and growth. Such
changes in microstructure are expected to have major, and often deleterious, effects on the
mechanical properties of the irradiated material.
Irradiation-induced segregation is a complex physical phenomenon for which no complete
physical model exists. We will consider three models here. These are based on:

(1) Size factor


(2) Dilute solution solute:defect binding
(3) Concentrated solution, inverse Kirkendall effect.
PHASE STABILITY UNDER IRRADIATION 257

5.1. Size Factor


Self interstitials in BCC and FCC metals are believed to exist in a "dumbell" configuration
where two atoms share a single lattice site (Haubold and Martinson, 1978).
The dumbbell configuration has a lower strain energy than does a substitutional atom on
an interstitial site, but still has an energy the order of 5 eV. Accordingly, incorporation of
an undersized solute atom into the dumbell would reduce the lattice strain and thereby the
strain energy. The small atom would tend to remain part of the dumbbell, which would
migrate by interchanging solvent atoms. Thus, any solute which is preferentially incorporated
into interstitials will tend to be segregated to point defect sinks (Okamoto and Wiedersich,
1974).
The criterion for undersized atoms is that the partial atomic volume be less than the atomic
volume of the solvent. The size factor, f~sf, is defined as:
f~sf = (fi~ - f ~ ) / t ~ (5. l)
where ~B is the partial atomic volume of the solute and f~Ais the atomic volume of the solvent.
Undersized atoms have f~,f< 0, and should be segregated toward point defect sinks. Table
1 extracted from King (1966) gives ~,f for a number of binary alloys.

5.2. S o l u t e : D e f e c t Binding
Johnson and Lam (1976, 1977) and Wiedersich et al. (1977) have developed a detailed
model which is applicable to solutions dilute enough that a point defect will usually interact
with no more than one solute atom at a time. In this limit, solute:defect complex formation
and decay may be considered as isolated pairwise interactions. The various complexes are
continually forming and dissociating, and the various mobile defects diffuse through the
matrix in response to concentration gradients. The Johnson-Lam analysis considers isolated
vacancies (v), interstitials (i), solute atoms (s), two kinds of interstitial-solute complexes (isa
and isb) and vacancy-solute complexes. The atomic displacement rate is K0. Vacancies and
interstitials combine with a rate constant K~ (referred to as ~ in Eqs (2.10) and (2.11). Of the
two types of interstitial-solute complexes, the isa complex (formation and dissociation rates
of K2 and K2) has the solute at one end of the dumbbell and is mobile. The type b complex
has the solute on a nearest-neighbor site on a plane normal to the dumbbell. The type b
complex (formation and dissolution rates K3 and K~) is immobile and cannot lead to solute
segregation. However, type b complexes do tie up solute and interstitials, and must be
considered in the calculation. Vacancy-solute (vs) complexes, where the vacancy and solute
are nearest neighbors, form and dissociate at rates K4 and K~. Interstitials and vacancy-solute
complexes combine at rate/(5, vacancies and isa complexes combine at a rate K6, vacancies
and isb complexes combine at a rate K7, while interstitials reach fixed sinks of concentration
Cis at a rate Ks and vacancies reach fixed sinks of concentration C~s at a rate/(9. The factor
¢ri accounts for solute encounters with interstitials giving a flux in the direction of the
interstitial current and ao accounts for solute encounters with vacancies giving a flux in a
direction opposite the vacancy current.
The resulting differential equations are then:

Ot = V. (1 + oiC~)D~VC ~+ Ko - K1C~Cv- K2CiCs

+ K'zC,s. - K3CiCs + K3C~b - KsC~C,s - KsC, C ~, (5.2)


258 PROGRESS IN M A T E R I A L S SCIENCE

ac
~---~-= V-(1 + avC~)hvVCv + Ko - K,C, Cv - K4CvC s

+ K'4C~s - K6C~C,.a - -
KTC~C,,b - Kg(C~ __
Cv )C~s,
cq
(5.3)
ac~
a--i- = V ' ( a , D , C V C , - a . D ~ C V C ~ ) - KzC, Cs + K'2C,sa

- K 3 G C + K;C,,b -- K4C~C + KICk,


+ KsC~C~, + K6CvCisa + K7CvCisb, (5.4)
ac~
- -
at
= D,,,V2C,~a + K2C, C~ - K~C,,, - K6CvCisa, (5.5)

-- = K3C~C - K;C,sb -- K7C~C,,b, (5.6)


Ot
a~
at
= D~,V2C~. + K 4 C v C - KICv. - KsC, C~.. (5.7)

Table 1. Volume size-factors (f~sf) and lim-


iting concentrations (Cmax) for metallic solid
solutions

Cmax Q~r
Solvent Solute (at.%) (~)

AI -Ag 20 +0.12
42u 2.2 - 37.77
--Ge 2.0 +13.13
-Li 12 -2.10
-Mg 15 +40.82
-Si 1.0 - 15.78
-Zn 25 - 5.74

Cu -Aut 40 +47.59
-Be 13 - 26.45
~2o 15 - 3.78
--Cr 1.6 + 19.72
-Fe 2.6 +4.57
-Nit 32 - 8.45
-Si 12 + 5.08
-Ti 0.25 +25.74
-Zn 15 + 17.10

Fe -A1 25 + 12.79
--Co 20 + 1.54
--Crt 9.0 +4.36
~Cu 0.7 + 17.53
-Mn 10 +4.89
-Mo 2.0 +27.51
-Nb 3.0 + 17.58
-Ni 5.0 +4.65
-P 3.5 -13.16
-Si 10 -7.88
-Ti 1.8 + 14.44
-Vt 30 + 10.51
-W 7.0 + 32.99

Mo -Tit 8.0 +0.56


PHASE STABILITY UNDER IRRADIATION 259

Table 1 (contd.)
Cm.~ fgr
Solvent Solute (at.%) (%)

Ni -A1 10 + 14.70
-Cr 40 + 10.34
~2ut 68 +7.18
-Fe 58 + 10.57
--Ge 10 + 14.76
-Mn 25 + 23.20
-Mo 22 + 22.27
-bib 8.0 + 51.24
-Si 12 -5.81
-Ti 9.0 + 29.43
-V 20 + 13.34
-W 15 + 36.93

Ti -A1 27 - 20.09
-Cr 20 -37.71
-Fe 20 - 53.62
-Mn 13 -51.17
-Mot 40 - 20.96
-V 8.0 - 15.40
-Zrt 30 + 30.08

U -Cr 1.4 + 4.03


-Nbt 30 - 24.49
-Ti 3.0 - 8.00

-AI 50 + 8.84
42rt 40 - 15.91
-Fet 10 - 18.86
-Mot 50 + 9.83
-Nbt 20 + 27.93
-Tat 70 + 36.06
-Wt 50 + 12.64

t Lattice parameters have been measured


across these mutual solid solutions from 0 to
100~ solute.

Analytical solution of Eqs (5.2)-(5.7) is not possible and numerical methods must be used.
Even so, the equations present formidable computational problems with stability and
convergence. The equations are usually solved for a one-dimensional case, which is spherical
coordinates for solute segregation to a void, cylindrical coordinates for segregation to a
dislocation, or Cartesian coordinates for segregation to a planar surface or interface.
Equations (5.2)-(5.7) have been evaluated for a number of cases of experimental interest.
Figure 21 shows calculations made for irradiation-induced segregation of Zn to the surface
of a foil of a dilute Ag-Zn solid solution. The various vibrational frequencies and binding
and migration energies involved in Eqs (5.2)-(5.7) were obtained primarily from measured
values, although some were estimated from physical considerations.
Huge enrichments---concentrations up to a hundred or more times the original foil
composition--are seen to be predicted in the absence of precipitation. If the surface
concentration cannot rise above the solubility limit due to precipitation, reduced but still large
surface enrichments are predicted, especially at higher temperatures.
Figure 22 shows that a higher displacement rate shifts the region of segregation to higher
temperatures. The shift arises from the higher concentrations of point defects created by the
higher displacement rate. These defect concentrations persist to a higher temperature, to give
260 PROGRESS IN MATERIALS SCIENCE

iff i

"~ ial Concentration


10-3 T : I00*C

I I I I
1°-40 ioo 200 soo 400 500
Distance from surface (A)
FIG. 21. Irradiation induced segregationof Zn to the surface of an Ag foil. K = I0 -3/s. (After
Johnson and Lam, 1976.)

the high temperature shift. However, at low temperatures most of the defects annihilate due
to direct recombination, which does not give solute segregation. As such, lower displacement
rates give more segregation than high ones at low temperatures.

5.3. Inverse Kirkendall Effect


In the well-known Kirkendall effect, unequal diffusion coefficients of various components
of a solid solution give rise to a vacancy flux and destruction and creation of lattice planes.
In the inverse Kirkendall effect, a flux of irradiation-induced vacancies gives rise to a flux
of solute atoms and produces segregation of the solid solution. Marwick (1978) and
Wiedersich et al. (1979) have presented analyses of this effect which are applicable to
concentrated solutions, where the Johnson-Lam model does not apply. Analysis of solute
segregation in the concentrated solution is in a sense simpler than the dilute case, in that each
vacancy, interstitial, or lattice atom sees approximately the same surroundings as do all other
of its species, and solute:vacancy interactions may reasonably be ignored.

160 I I I I I I .' "J


,/~\ Rv =50 A=]
140 - / \ --- Rv = 5 0 0 A 7

,20 '-°~ ~k~'\ -3!


"

°c~
xloo

0,.,.,8o,..
- II;I // k\ \ , \/

40- --
\\_
o .,,iS" i i I "~ 'r-~',,--
200 400 600 800
T (K)
FIG. 22. Effect of displacement rate on the fractional enrichment of Zn solute at void surfaces in
Ag. (After Johnson and Lain, 1977.)
PHASE STABILITY UNDER IRRADIATION 261

The Marwick analysis, based on work by Manning (1968), includes the effects of solution
non-ideality. In general, replacing an A atom by a B atom involves a free energy change,
which biases the jump direction.
The fluxes of the two components are given by the familiar equations:

[IjA = _ D A OCA r ~D*


~x + "~A~-x ' (5.8)

f ~ B = -- DB dCB r
~O ~
t3x + ,~B ~-x-x' (5.9)
where D is atomic volume, assumed independent of composition, and concentrations are in
atomic fractions. The intrinsic and tracer diffusivities, DA and D*, are related by:
DA = ~rA D*, (5.10)
d In 7A
where ~ = l + - -
d l n CA
)~A activity coefficient of A,
=

N A(O* - D*) (1 -)Co)


rA: 1+
fo(NAD* + NBD*)
where f0 = correlation factor for the particular crystal structure. The tracer diffusivity is given
by:
D~, = yooAfA d ECv, (5.1 l)
where y = number of equivalent jump directions,
~OA= basic rate of exchange of a vacancy with an A atom,
fg = correlation factor for A atoms,
d = jump distance.
Similar expressions obtain for DB and D*.
Marwick assumes the interstitial flux to be neutral, that:
J~ = CA'Ji (5.12)
where Ji = total interstitial flux, and
J~ = flux of A atoms moving interstitially.
Combining the various contributions to JA gives the flux at a surface of:
•JA = - - (DACa -k DaCg) VCA + CACn V(D* -- D*) (5.13)
or by factoring DA -- FACv, D* - F* Cv, etc.,
D.JA = -(FACa + FnCA) CvVCA + CACa(FA* -- F*) VCv. (5.14)
Thus, the flux of A atoms (assumed the faster diffusing species) up the vacancy gradient
is proportional to the difference in tracer diffusion coefficients of the A and B atoms. The
slower moving species, B, will then diffuse down the vacancy gradient and be segregated to
the point defect sink. In addition to segregating the slower diffusing element to the defect sink,
the concentration gradient gives rise to a Kirkendall flux of vacancies, jk:
fljk = (DA -- DB) VCA. (5.15)
The Kirkendall flux opposes the vacancy flux to the sink and would retard the growth of
voids.
J.P.M.S 28/3-4~C
262 PROGRESS IN MATERIALS SCIENCE

Wiedersich et al. (1979) extended the analysis of Marwick to account for migration of
irradiation-induced interstitials also leading to solute segregation. Their analysis, which
generally parallels that of Marwick, arrives at an expression for the steady-state solute
concentration gradient near a sink. The concentration gradient is equal to a positive quantity
(a function of diffusion coefficients) times:
l"Av FAi"~
F-~B~)VC~,. (5.16)
Fa~
The Fs are simply the effective jump frequencies of "A" and "B" atoms into neighboring
vacancies or interstitials.
During irradiation, the vacancy concentration decreases near a sink, and the element A
becomes enriched at sinks if FA]FBi > FAv/FB~,,and vice versa. This criterion for the direction
of solute segregation is reduced to the simpler Marwick case (Eq. (5.15)) by setting FA~/F~ = 1
and multiplying by FBv.
The theory of solute segregation in concentrated solutions is thus simpler and has fewer
unknown energetic factors than the theory for dilute solutions. It is also far more transparent,
mathematically, and involves only four quantities (the Fs) and those only as two ratios.
Unfortunately, diffusion coefficients, hence values for the F's, are often very poorly known
for concentrated alloys.

6. SPINODAL INSTABILITIES
The treatments of solute segregation discussed so far all assume that the solid solution is
stable under irradiation, that concentration fluctuations do not spontaneously increase as
occurs under thermal conditions during spinodal decomposition. Martin (1975, 1980) has
considered this assumption of stability and found it to be unjustified under a range of
experimental conditions.
Martin considers a binary A-B solid solution which also contains irradiation-induced
vacancies and interstitials. The conservation equations are:

~Ci - K - - ~ C i C v -- V" Ji, (6.1)


Ot
aC
- K - ~tCiC~ - V" Jr, (6.2)
Ot
ac. = _ f. L.
(6.3)
Ot
In that CB + CA = 1, CA need not be considered explicitly.
The familiar phenomenological equations for the fluxes may be written as:

JP "~--2 OpqVCq, (6.4)


q
where p, q are i, v, B. The Dpq may in general be functions of concentrations and concentration
gradients, but when considering only small fluctuations may be taken as constants.
The solid solution is assumed to be perturbed by small concentration fluctuations given
by:

Cp( r, t) = Cp + Ap( i , t) COSk. ; , (6.5)


PHASE STABILITY UNDER IRRADIATION 263

where r = position vector


~p = concentrations in a homogeneous solid solution, where p is in turn i, v, or B
Ap(k, t) = wave amplitude

[ k I = 2n/2, where 2 = wavelength of fluctuation.

The amplitudes evolve with time according to:

10Ai DiiAi + Di~A~ + D,~AB + K'~tAv + K"~Ai, (6.6)


k ~ c3t
1 OA.
- - -- D~,iAi-t- D,,A~ + DvnAB + K'~iA, + g'~.~,Ai, (6.7)
k 20t
10A B
- - = DBiAi + DBvA, + DBBAn, (6.8)
k 20t

where K' = ot/k 2.


The system was analyzed only for conditions where the solid solution is stable in the
absence of mutual recombination. The stability of the solid solution in the presence of
recombination is governed by the sign of the eigenvalues of the matrix, ~. These eigenvalues
are the solutions of the characteristic equation:
Det. (t~ - o97) = 0,
where 7 is the unit matrix, and

D. + K'~v Div + K'Ci Din


(~ ~ - - k 2 D~ + K'~'~ D~ + K'~.~ D~B (6.9)
DBi DB~ DnB

We see that 0 is the matrix of coefficients of Eqs (6.6)-(6.8).


The solution may be unstable if one or more of the eigenvalues has a positive real part.
As Martin notes, a general discussion of eigenvalues of 3 x 3 matrices is tedious; we only
summarize his results here.
The signs of the eigenvalues are determined much more by the signs of the various Dpq than
by their magnitudes. That is, the signs of the cross terms in the Dpq are of special importance.
Furthermore, recombination ( K ' # 0) is required for instability to occur. Martin discusses
instability in a two-dimensional F - G space, where F and G are complicated polynomial
functions of the various terms in the matrix O. The criteria for instability are given by two
inequalities (criteria I and II) in the terms in 0, and both must be satisfied to yield instability.
Figure 23 shows eight different cases for the conditions of instability. In case 1, the same
result is obtained whether the signs of D~B, Dn~, D,a and DBi, alternate + -- + -- or
-- + -- + . The inequality referred to in the figure is a complicated function of the
displacement rate, recombination coefficient, and the various defect diffusion coefficients. In
cases 2-6 the criteria are satisfied in different, but overlapping, regions. In case 7 the two
regions do not overlap, and for case 8 criterion II is satisfied nowhere in the F - G plane, and
the solid solution is therefore stable against concentration fluctuations.
264 PROGRESS IN M A T E R I A L S S C I E N C E

>m ,~ m m
> Domain of Validity Domain of Compatibility of I and IT
"" "" " "" of Condition
Case ~,~c~[~ld~t:~l ~ CriterJ0nl Criteri0n~ Shapeand location [ of Existence

4- -- '!" --
G uGl ~ I ~ . , /"
1.'" Same as the domain I
I No extra
1 - + - + _ F ~ of validity of inequality I condition

2 + +-+- G

++ 4-
3 4- -+

4- 1~ . iB BV
++

4 +++ iV . ,

+ ++ L" F
5 4" "!" +

6
"i + + + '/I OBVOiB~ OVBOBi

+ +
7
÷ 4- Conditions I and 11
i are not compatible
4-'1"+-I-
8 +++i+

FIG. 23. Criteria for an irradiation induced spinodal instability in a solid solution. The signs of
the diffusion coefficients are the main determining factors. (After Martin, 1980.)

It is interesting to note that in both cases 1 and 7 all the relevant products of Dpq'Sare
negative, but that the different sign variations giving the negative products give very different
results. Case 1 gives a large region of instability while case 7 has none. The importance of
alternating signs in the Dpqin promoting instability is illustrated by case 8. Either all positive
or all negative Dpq gives all positive products and no region of instability.
Martin notes that just as irradiation may destabilize a stable solid solution, the reverse
effect is also possible. Irradiation, through the kind of solute segregation effects just discussed,
could very well stabilize a solid solution which would otherwise decompose.
Figure 24 shows calculations of regions of instability for A1-Zn solid solutions, in a
temperature-displacement rate space. Except for the highest solute concentrations, the solid
solutions are thermally stable over most of the temperature range shown. Irradiation clearly
has a strong destabilizing effect, even at relatively low displacement rates. The complexity of
the criteria for destabilization are vividly shown by the lack of simple variation in stability
PHASE S T A B I L I T Y UNDER IRRADIATION 265

tg itI 11 ."';L"
If i : / -s.x_"
I ,/.C.-7
,, .... ...... . . : y
: d S .." ~'s • .w.w

10-1 , ..i
i ,,..'/" /z~,~d"
,.,'/ /X,~¢ .+ ~ , ~

~lO.-O
7Y
. . . . . .

FIG. 24. Isothermal cuts of the calculated spinodal surface in an irradiated AI-Zn solid solution
under irradiation. (After Martin et al., 1981.)

with temperature, composition, or displacement rate. For example, at 450 K, with increasing
displacement rate, alloys of a range of compositions go from stable to unstable to stable
again.
The physical basis of the instability is that a local increase in solute concentration attracts
vacancies. In turn, more interstitials are annihilated in the vacancy rich region, which in turn
increases the local solute concentration yet more. The physics are the same as those for
irradiation enhanced precipitation of coherent particles described in Section 4.2. Only the
mathematical formulations are different.
Very recently, Krishnan and Abromeit (1984) presented a theory for radiation-induced
spinodal type instabilities in concentrated alloys. The physical basis of the instability is a
composition-induced bias in A or B interstitial:vacancy recombination. In this way, the bias
could give, say, fewer A interstitials in A-rich regions than elsewhere in the alloy. Then, A
interstitials would diffuse down the concentration gradient into the A-rich region, and enrich
it further. In this way, a spatially periodic structure would be built up. The model yielded
a temperature and dose independent wavelength, which was stated to be in agreement with
experimental observations.
Although Krishnan and Abromeit used a formalism similar to that of Martin, the physical
bases of the two theories are entirely different. In Martin's model decomposition is induced
by coupled diffusion fluxes, whereas in the model of Krishnan and Abromeit decomposition
arises from a bias in vacancy:interstitial recombination.

7. RECOIL RESOLUTION*
Several analyses have been made of the effect of ejection of atoms from the particles by
impinging energetic atoms on particle stability. These analyses are of two basic types. In one
case, the particles are to be infinitely spaced, so that the solution to the diffusion field around
a spherical particle in an infinite medium may be employed. In the second case, a cell model
is employed and the diffusion fields around the particles depend on the inter-particle spacing.

*See also notes in proof (pp. 430-434).


266 PROGRESS IN MATERIALS SCIENCE

7.1. Infinite Inter-Particle Spacing


The earliest and probably most easily understood analysis of recoil resolution is that of
Nelson et al. (1972) who considered the balance between precipitate shrinkage due to
irradiation resolution and re-precipitation as speeded by irradiation-enhanced diffusion.
The objective of the calculation was to determine the growth (or dissolution) rate of
particles as a function of particle size and number density, temperature, and irradiation
conditions. No attempt was made to include transient effects, to calculate concentration
profiles in the matrix, or to determine the evolution of a system of particles in a Wagner (1961)
and Lifshitz and Slyozov (1961) type of calculation. Spherical particles were assumed, and
no account was taken of purely thermal dissolution. Infinite inter-particle spacing and
uniform dispersion of recoiled atoms through the matrix were implicitly assumed.
Each atom which recoils across the particle:matrix interface with an energy of greater than
about 25 eV will be dissolved in the matrix. The flux of atoms (per unit area of interface)
was estimated as ¢ K, where K = atomic displacement rate, and the constant ~, for other than
electron irradiation, was estimated at 1018/m2. The particle radius, r, would then vary due to
recoil resolution as:
dr
- - = - ~ g f2, (7.1)
dt
where fl = atomic volume in precipitate.
The requirement of conservation of solute and use of the usual equations for the growth
rate of spherical precipitates yields:
dr 3DC
- - = - ~ K f~ -t Dr2pp, (7.2)
dt 4~rCp
where r = particle radius
C = total amount of solute (atomic fraction)
Cp = atomic fraction solute in particle
pp = number of particles/unit volume.
Figure 25 is a schematic plot of d r / d t versus r. Particles are predicted to approach an
equilibrium radius from larger or smaller sizes. At this equilibrium size, or node, the rate of
+

dr
I

dt

0
----o.

FIG. 25. Growth rate o f a precipitate particle which is experiencing recoil resolution. The smaller
particles are seen to be more stable than the larger ones. (After Nelson et al., 1972.)
PHASE STABILITY UNDER IRRADIATION 267

re-precipitation just balances the rate of solute loss from the particle due to disorder
dissolution. The usual size dependence of particle stability is seen to be reversed, with small
particles more stable than large ones. This inverse coarsening (or inverse Ostwald ripening)
is another example of the harnessing of irradiation energy to reverse the path of evolution
which a system would take in the absence of irradiation.
Brailsford 0980) performed an analysis of recoil resolution which again assumed infinite
inter-particle spacing, but assumed a uniform rate of resolution in a shell of thickness R
around the particle. The difference between the initial particle radius, r0, and final radius, rp,
was found to be:
3 - 3R2rp(3rp + 2R) l? (7.3)
r~ - r p - 8rc Dp,[(rp + R) 3 - rp3l 4rcr~'
where pp = particle number density and V = volumetric loss rate due to recoil resolution.
Equation (7.3) reduces to the Nelson et al. treatment (1972) in the limit of R - . o o .

7.2. Cell Models


Cell models take a boundary condition of (OC/Or)L = 0, where 2L is the particle spacing.
Wilkes (1979) employed a cell model in his analysis and assumed the recoiled solute to be
uniformly distributed throughout the cell. The steady state concentration profile around a
particle of radius rp was found to be:
~ Kr2p F2L 3(r - rp) ]
C(r) = Ce -~ 2D~-L5~ r3) L ~rp rE - rzp " (7.4)

where Ce = equilibrium solubility,


r = distance from the particle center.
The concentration gradient at the particle surface is seen to be inversely proportional to the
diffusion coefficient. Accordingly, unlike the previous analyses, the rate of particle growth or
dissolution is independent of the diffusion coefficient.
Frost and Russell (1982a,b) presented an analysis based on a cell model and finite
resolution distance, and also included transient effects. The treatment considers an array of
spherical particles of radius rp and interparticle spacing 2L, placed in a matrix with initial
solute concentration Co. The rate at which solute enters the matrix is the rate at which recoils
originate in the particle and end in the matrix. This can be described as a source generation
term g(r) in the matrix which depends on the radius r from the particle center. The source
term is given by Gelles and Garner (1979) as:

G(r) = 4 ~ [r2 - (r - R)2], (7.5)

where S is the resolution rate, per atom, and R is the recoil distance. A more complex
expression obtains when R is not the same in the particle and matrix. Since no deposits occur
beyond one recoil distance from the surface, G(r) = 0 for r >i rp + R.
With the same boundary conditions as used by Wilkes, the concentration profiles were
obtained by solving:
8C
O~ = D V e C + G. (7.6)
268 PROGRESS IN MATERIALS SCIENCE

The steady state solution within one recoil distance of the particle is:

c~(r) = Co+~ ~3 - - Tp~- - 4R(r 2 - - r~) - - 6(r,~-R2)(~-r,)

The maximum concentration, C~,x, is:

Cmax--C(r>~rp+R)=Co+i~l - . (7.8)

This maximum is seen to depend linearly on the recoil rate, S (or the radiation flux), on the
recoil distance squared, inversely on the diffusion coefficient, and for rp > R, only weakly on
rp.
Transient values for the concentration profiles were determined in a closed form series by
summing contributions from concentric, incremental, spherical shells. Figure 26 plots
concentration against distance from the particle center for various values of ~v/~.
Steady state is approached when t ~ T, where:
L 3
Z ~, rp--D" (7.9)

Thus, steady state is reached when the characteristic diffusion distance x / ~ is comparable
to the interparticle half-spacing, times a geometric factor which accounts for the ratio of
particle radius to spacing.
Neither the resolution rate nor recoil distance enters into T. Taking L = 10-4cm,
r = 1 0 - 6 cm, and D = 10 -8 cm2/s, z -- 100 s, which is small compared to most experimental
times. Thus, even for lower D and larger L, which give larger T, steady state should be reached
very early in essentially all neutron irradiations and in many heavy ion irradiations.
The thermal equilibrium concentration increases with temperature whereas the concen-
tration from recoil resolution decreases; at some temperature the two will be equal. Figure
27 shows this behavior for TiC particles in stainless steel in which an excess of Ti over
stoichiometry leaves the free carbon in such low concentration as to be rate controlling for
coarsening or inverse coarsening.

L:iOrp : R:O.2 rp
64
g 0.003 32 -

e ~"I ° 0.002 16 --

~ 0.001 = --
z

0 2 4 6 B
r/rp

FXG. 26. S o l u t e c o n c e n t r a t i o n profile a r o u n d s p h e r i c a l p a r t i c l e u n d e r g o i n g recoil r e s o l u t i o n a t a r a t e S.


rp = p a r t i c l e r a d i u s , L = p a r t i c l e s p a c i n g , R = recoil d i s t a n c e .
PHASE STABILITY UNDER IRRADIATION 269

T (°C)
600 400 300 200
I I I I I
S: 10-4/$~ ~ S : ~'71sec
10-4

z
o 10-8
104. i0 I0
/\ ! i CARSON ,~
/ ~1 J AUSTENITE
u
z
o io-~
x
1- "4," '~rp = IOOA
.
i0 -~

WITH 0.1 AT% TITANIUM "~t


IC)z£ IN SOLUTION I I ~'~'~

1.0 1.5 2.0 2.5


I000
T
FIG. 27. Calculated thermal and irradiation-altered solubility of TiC particles in austenitic stainless
steel for two different displacement rates. Symbols are the same as in Fig. 26.

In Fig. 27, the equilibrium carbon concentration is plotted together with Cm~xfor recoil
rates typical of fast neutron irradiation and heavy ion bombardment. Concentrations are
shown for r = 100 ,~ and r = oo; the difference between the two drives either thermal or
inverse coarsening. Under recoil resolution the small particles are seen to have a lower
solubility than large ones, which will lead to inverse coarsening.
Figure 28 illustrates some limits on the recoil resolution model. At sufficiently low
temperatures (e.g., less than 300°C for S = 1 0 - 7 / s ) , the matrix concentration increases to the
point that all the solute is dispersed into the matrix. At this point all the particles will be
dissolved, regardless of size.
In many cases the predicted recoil induced concentration exceeds the equilibrium concen-
tration by many orders of magnitude so that new particles will nucleate and grow in the
matrix until the supersaturation is lowered to the point that nucleation ceases. This
re-precipitation will start near the particle as the solute builds toward the steady-state profile
and results in a tight halo of sub-particles.
The time required for an initial distribution of particles to reach a uniform dispersion of
fine sub-particles can be estimated by treating the dispersion of solute as a diffusion problem
in which the jump distance is about one or two recoil distances, and the jump frequency is
the recoil rate. A jump distance of 200 A and a frequency of S = 1 0 - 7 / s give Deer= 10-~9 ClTl2/s
(for reactor irradiation). With an initial spacing of 0.2/am, we approach steady state in 109 S
(about 30 years). For S = 10 -4, typical of heavy ion bombardment, the time is 106 seconds
(about 11 days). Both times are very long as compared to the time scale of experiments so
270 PROGRESS IN M A T E R I A L S SCIENCE

T (*C)
IO
600 400 300 200

=o 05
S 10.4 /$eC
=

FRACTION:
)t!t!t5"I
Vf : 10-3 I--- ~
I
S I0-r/sec
:

o
LI- vf= Io-2 I -..I
I I
I I
1 tl
I0 15 2.0 2.5
I000 / T

Fro. 28. Volume fraction of TiC particles in austenitic stainless steel not dissolved by recoil resolution
as a function o f temperature for two initial volume fractions and two displacement rates.

that only the early stages of the process are likely to be observed, as was done by Jones (1978)
for ThO2 particles in ion-irradiated Ni-ThO2.

8. ORDERING AND DISORDERING


The effect of irradiation on ordered alloy phases can be described as a balance between
two conflicting processes: the disordering produced by the atomic replacements during
cascade and point defect production, and the ordering facilitated by the migration of
vacancies and interstitials. Both these processes have been observed experimentally, as has
been reviewed recently by Schulson (1979).

8.1. Nelson, Hudson and M a z e y Model


The analysis by Nelson et al. discussed in Section 7 was in fact developed for both recoil
resolution and irradiation-induced disordering. In the latter case, a fraction, f, of the
disordered atoms in a shell of thickness, ~, at the surface of an ordered precipitate are
assumed to be lost by diffusion into the matrix. Then the factor ~f~ in Eq. (7.2) is replaced
by ~f to give:
dr 3DC
- f f K -t Dr 2pp (8. l )
dt 4nrCp
where the variables were defined after Eq. (7.2). It was estimated that for coherent, ordered
precipitates. E ~-10-6cm, f - 1 , so that Ef---10-6cm. The corresponding factor in their
treatment of recoil resolution is ~ f~ -~ 10-11 cm, so that disorder dissolution was expected to
be by far the more effective process. A plot of Eq. (8.1)* would resemble Fig. 25, except that
the stable particle size would be much smaller, due to the greater efficiency of disorder
dissolution.

*All quantities other than f f being the same as in Eq. (7.2).


PHASE STABILITY UNDER IRRADIATION 271

8.2. Liou-Wilkes Model


The detailed kinetics of the ordering and disordering processes were modeled recently by
Liou and Wilkes (1979). The rate of change of the order parameter, S, is the sum of the
irradiation disordering and radiation-enhanced thermal ordering rates:
dS
()
dS
d---t= - ~ irr
+ -~
th" (8.2)
Under constant flux and temperature, S will approach a steady state value which will not
exceed the equilibrium value for that temperature. The Bragg-Williams definition of the long
range order parameter for a binary alloy with two sub-lattices c¢ and/3 is:

S = fa= - CA (8.3)
1--CA
where fA= is the probability of an A atom being on an ~ site, and CA is the atomic fraction
of A atoms. S = 1 when the alloy is completely ordered, and S = 0 when it is completely
random.
The disordering rate depends on the rate of atom replacement, which should be linear in
the irradiation flux. If the displaced atoms are replaced randomly by other atoms, the
disordering rate is proportional to the order parameter:

(dS)=~KS, (8.4)
-'~ irr

where e is the disordering efficiency, which depends on the type of radiation. (e is the order
of 10 to 100 for heavy ions and neutrons, 1 to 3 for electrons.) In this approximation the
disordering rate does not depend directly on temperature.
Both interstitials and vacancies can move through a disordered lattice in such a way as to
increase the order, so that the ordering process occurs by the motion of point defects. If the
local degree of order is lower than the equilibrium level, the energy differences between
various possible jumps will induce the defect to follow an ordering path more often than a
disordering path. The ordering rate depends on the local order parameter because the fraction
of available jumps that will increase the local order depends on the local order.
The ordering rate is proportional to the concentration of mobile defects, the rate at which
they jump, and the degree to which ordering jumps exceed disordering jumps. The ordering
rate therefore depends on the irradiation flux through the dependence on defect concen-
tration. It depends on temeprature through the defect concentration and the jump rate and
jump preference terms. The defect concentrations were given as functions of temperature and
flux in the earlier section on enhanced diffusion. The defect jump rate is proportional to the
defect diffusion coefficient times a correlation factor that takes into account the local
ordering.
At intermediate temperatures where recombination is negligible, we found that DiCe~- K/k~
and DvCv~-K/k 2. In this case, if the sink density does not change with temperature, the
number of defect jumps per unit of damage will remain constant with temperature.
As the temperature is lowered, the fraction of ordering jumps should increase because of
an increasingly favorable Arrhenius factor.
The foregoing factors lead to the approximation that in the regime of diffusion to fixed
sinks, the ordering rate should increase as the temperature is lowered. Since the disordering
rate is constant, the steady state order parameter should increase (or remain constant) as the
272 P R O G R E S S IN MATERIALS SCIENCE

temperature is lowered. Only when the temperature is reached at which direct recombination
of defects becomes important will the trend reverse.
Liou and Wilkes discuss models for the ordering and disordering jumps of each type of
defect. For S approaching unity they quote from Nowick and Weisberg (1958):
s,

where x includes the defect concentration, the jump frequency, and a term for the energy
difference between a right pair and a wrong pair, and Se is the equilibrium order parameter.

x= vvCvexp (Z=+Z,-2)+vitrt~iexp~---~-)~-~s) Z, exp ,(8.6)

where vv and vi are frequency factors for the exchange by vacancy and interstitial, E~, and
E~, are activation energies for defect jumps, Z~ is the number of fl sites neighboring an ¢ site
and Z= is the reverse, a is an interstitialcy reordering efficiency containing geometric factors,
and V0 is the energy gained by transforming a wrong pair into a right pair.
The ordering and disordering terms can be combined to calculate S(t), which at steady
state is:
[ ZK2 2 4eKx ) (1 -- Se)
S= 1+ eK-~/e +4x + ~ 2-x (8.7)

The effect of decreased order in ordered precipitate phases will be to raise the free energy
over the equilibrium value. A free energy increase will also occur in solid solutions with short
range order, but to a much lesser extent. The energy of disordering is also large compared
to the energy of the point defects present at a given time in any irradiated phase.
The free energy increase will change the compositions that result from the local approach
to equilibrium at the particle:matrix interface, as seen in Fig. 29. The ordered fl phase is in
equilibrium with ct phase of composition C A. If fl is partially disordered to a state of higher
free energy, fl', local approach to equilibrium will tend toward a matrix composition of C~,.
If the alloy composition is between CA and CA, the precipitate will be dissolved.
In the extreme case, the disordered state, for example fl", may have such a high free energy
that the phase does not form at all, or completely dissolves, and the next phase (in this case
V) may appear. In this case the solute concentration in the matrix would approach C~,.

PARTIALLY
DISORDERED
DISORDERED , /

>-
IZ
bJ
Z
LU
W
W
n,."
IJ..
I II ~ ORDERED
i I I B
I i

CA C~ C~ COMPOSITION

FIG. 29. Schematic plot of the effect of irradiation induced disordering on the free energy and
stability of an ordered phase. (After Liou and Wilkes, 1979.)
PHASE STABILITY UNDER IRRADIATION 273

3000
(o) i I
LIQ. / ~

2000

T,K

IO00

E E+TiRu 1
0.5 I.O
CRu.__~ RU
3000 i
(b)
LIQ.

2000

T,K

iOOOi /~ / ~+TiRu ~' + TiRu

O 0.5 I.O
Ti -('Ru~ Ru
FIG. 30. TiRu phase diagram, thermally(a) and in the presence of irradiation induceddisordering
(b). (After Liou and Wilkes, 1979.)

The effect of irradiation on the stability of ordered phases may also be illustrated by an
irradiation-altered phase diagram; Fig. 30 shows the Ti-Ru phase diagram under thermal
equilibrium conditions and under irradiation. Irradiation destabilizes the ordered Ti-Ru
phase; at low temperatures it disappears, to be replaced by the e, e', and fl phases. Once again
the energy of irradiation has been harnessed to produce a microstructure of higher than
equilibrium free energy.

8.3. Banerjee-Urban Model


Very recently, Banerjee and Urban (1984) developed a theory of the order-disorder
transformation which follows the same general lines as that of Liou and Wilkes (1979) but
differs in the following significant respects:
(i) Vacancy concentrations on the two sub-lattices are distinguished.
274 P R O G R E S S IN MATERIALS SCIENCE

(ii) Does not require the same vacancy to cause jumps of A and B atoms involved in an
interchange.
(iii) Is fully consistent with the Bragg-Williams solid solution model and is usable all the
way to To.
(iv) Is self-consistent concerning ordering energy, vacancy formation energy, and diffusion
activation energies, all of which are physically interconnected.
Figure 31 shows the results of calculations for the Dla structure, which is exhibited by
Ni4Mo. The figure shows two ordering temperatures. The upper ordering temperature, To,
is defined thermodynamically and is independent of displacement rate. The lower ordering
temperature, T*, is determined by the relative rates of irradiation-induced disordering and
ordering, and is dependent on displacement rate.

9. PARTICLE COARSENING

Under thermal coarsening conditions (Wagner, 1961; Lifshitz and Slyozov, 1961) the
particle size distribution, p(r, t) reaches an asymptotic form such that it is the product of two
functions:
p(r, t) = f ( r /~)g(t ), (9.1)
where r and ~ are actual and mean particle radius, respectively. The evolution of mean particle
size is described by:
~3 __ ~ 3 __ 8DC7 V2mt
9RT ' (9.2)

where ~0 = mean radius at t = 0,


D = diffusion coefficient of solute in the matrix,
C = solubility of particle of infinite radius,
Vm = molar volume,
7= particle: matrix surface energy,
R = gas constant,
T= temperature.

1.0

0.8
o
g
0.6
E

.~o 04
ihq--5 ,0-5
0.2
,r'h

0 500 I000
T(K)
FIG. 31. The steady state degree of long range order, S, in a DI a structure as a function of
temperature and displacement rate. (After Banerjee and Urban, 1984.)
PHASE STABILITY UNDER IRRADIATION 275

Bhattacharyya and Russell (1972) showed that for compound precipitates, Eq. (9.2) is
modified by a numerical factor close to unity, and D and C refer to the component for which
the product D C in the matrix is least.

9.1. Disorder-Dissolution
Bilsby (1975), M. Baron et al. (1977a,b) and M. Baron (1981) produced formulations to
describe the evolution of a system of precipitates with a distribution of sizes. Both treatments
used the concepts of Nelson et al. (1972) of irradiation altering the particle growth or
dissolution rate by a combination of disorder dissolution and enhanced diffusion.
Bilsby presented a detailed derivation which roughly parallels that of Wagner. Irradiation-
enhanced diffusion is accounted for by an increased D, and the effects of disorder dissolution
enter through a factor P, which is the ratio of the irradiation induced dissolution rate to the
thermal coarsening rate. For P <<1, the effect of irradiation dissolution is to introduce a factor
of (1-4P) on the RHS of Eq. (9.2). Bilsby calculated that for ion bombardment disordering
rates, P is in the 105-10s range, far beyond the range of his treatment. In the range of validity
of the Bilsby treatment, P, hence the deviation from ~3 versus t behavior, increases with time.
M. Baron et al. (1977a,b) were apparently unaware of the earlier Bilsby treatment when
developing their theory. The fundamental assumption of their treatment is of dynamic
scaling, i.e., that the particle distribution function was of the form given by Eq. (9.1). They
considered the effects of irradiation disordering and enhanced diffusion and, unlike Bilsby,
allowed the matrix concentration to rise above the thermal value. Particles were allowed to
change size by two processes:
(1) steady drift due to the rates of thermal and irradiation mass loss being greater (or less)
than the rate of re-precipitation from the matrix;
(2) random walk, due to statistical fluctuations in the rates of solute addition and loss.
The resulting equation for p (r, t) is of the Fokker-Planck* type (Lothe, 1966). The solution
for p (r,t) as t ~ oo is an exceedingly complex expression, which involves tabulated functions.
Simpler expressions are, however, obtained for several quantities of interest. The maximum
stable particle size is given by:

= ( 3 a D ~ 1/2 (9.3)
r max \ ~fK ]

where f f is defined in section 8.1, and a = lattice spacing.


The solution of p(r, t) for finite times is an even more difficult problem than for infinite
times. However, it was found that, as in the thermal case, the average particle size was
proportional to the maximum particle size. In the case where the particles are initially rather
larger than the maximum stable size, the maximum size decreases according to:
rmax(Kt) - rm~x(Kto) = - f f K ( t - to). (9.4)
Equation (9.4) applies only for K t considerably less than the value needed to approach the
final, constant size distribution, p(r, oo).
In the other extreme, where the initial particle sizes are much smaller than given by p (r, oo),
then:
r3m~x(Kt ) - r3ax(Kto) = 9 a D ( t - to). (9.5)

*Fick's second law, written for diffusion in the presence of a potential field, is also of this type.
276 PROGRESS IN M A T E R I A L S SCIENCE

The order of K t = 10 dpa produces a constant size distribution, depending somewhat on the
particle sizes involved, but regardless of whether the approach is by growth or by dissolution.
The similarity between Eq. (9.5) and Eq. (9.2) for thermal coarsening is striking. Both have
the r 3 versus t dependence, and the proportionality coefficient in both cases involves the
diffusion coefficient. By contrast, in Eq. (9.4) for particle shrinkage r varies linearly with t,
with the proportionality constant depending only on the irradiation parameters.

9.2. Recoil Resolution

Frost and Russell (1982a,b) considered the coarsening of disordered particles by recoil
resolution, in what is an extension of their treatment and that of Nelson e t al. (1972) described
in Section 7. No attempt was made to calculate the particle size distribution and its evolution
with time, as was done by Wagner (1961), Lifshitz and Slyozov (1961), Bilsby (1975), M.
Baron e t al. (1977a,b), and M. Baron (1981).
In Section 7 the matrix solute concentration was shown to depend on particle radius for
a system of particles of identical radius; this size dependence is what drives inverse coarsening.
When the particles are not of identical size, however, the calculation is not exact. An
approximate description of the concentration in a system with a range of particle sizes is made
by assuming that the diffusion near each particle can be calculated based on the boundary
condition that the concentration at the edge of its cell, r = L, is an average matrix
concentration, Cm. This average matrix concentration will, of course, depend on the particle
radii throughout the volume, and the shape of the distribution of radii.
To get an approximation for the rate of volume change of the various particles one assumes
that there is no flux to or away from a particle of some radius r m.
The time required for the inverse coarsening process can be found by solving the differential
equation for particle radius:

dt= 1- P(L-r) 48
There are three different regimes for the evolution of r with time:
(A) r <<rm
( B ) r >>rm
(C) r x r,,.
In all cases, it is assumed that r <<L, the particle spacing; the starting particle size is r0.
In case (A), Eq. (9.6) may be integrated to give:

r3(t)
= 1+- t (9.7)
r3(0) ~A
where
48r°3 (9.8)
ZA = S R 3 •
For case (B):

= 1 ---, (9.9)
TB
where
24rm r2
z B -- SR 3 . (9.10)
PHASE STABILITY UNDER IRRADIATION 277

Thus, the particle size varies as r 3 versus t when approaching rm from below, and as r: versus
t when approaching from above.
In case (C):
r r., = (ro rm) exp (-- t/Zc),
- - - - (9.11)
where
48r~
Tc = SR3" (9.12)

In all three cases, for R, r 0, and r,, the order of 100/~, T lies approximately between IO/S
and 100/S. In fast reactor experiments, S ~- 10-7/S, SO that 108 s ~< r ~< 1 0 9 S. In heavy ion
irradiations S -~ 10-3/S, SO that 104 s ~< • ~< 105 s.
For larger r,, and r0 the incubation times would be longer. Even for particles the order of
100 ~, the incubation times are seen to be several years (for reactor experiments) and several
hours (for heavy ion irradiations). Both times are at the upper limits for the respective
irradiations. As such, one would expect to see substantial inverse coarsening by recoil
resolution only in the most extended experiments.

9.3 Point Defect Recombination


Urban and Martin (1982, 1983) extended the Cauvin and Martin (1981a,b) model for point
defect recombination altered phase stability (Section 4.2) to describe coarsening of coherent
particles under irradiation. The pseudo-free energy, G'(x), was used as the driving force for
coarsening in an analysis which generally paralleled those of Wagner (1961) and Lifshitz and
Slyozov (1961).
In the absence of irradiation, the expression for the number of clusters which contain x
atoms, pt(x, t), may be separated into functions of x and t. The resulting universal size
distribution may then be solved for p,(x, t). Urban and Martin found, as did Bilsby (1975)
that analytical solution for p,(r, t) was impossible. They found, however, that setting the
irradiation terms constant allowed them to use the arguments of Wagner and Lifshitz and
Slyozov to obtain solutions for short time intervals where the steady-state assumptions were
justified. Figure 32 compares the changes in mean radius with time according to the Urban

15-- Urbon-M ~

! //// -
"~ / ./Wogner-
V _ / ,/ Lifshitz- _

~.-----'r//// I I I
I0° I01 I0z 103
Time (min)
FIG. 32. Mean particle size as a function of time in the absence of irradiation (WLS) and as affected
by recombination of irradiation induced defects at the particle:matrix interface (UM). (After Urban
and Martin, 1982.)
J.PM.S. 28/3-4 D
278 PROGRESS IN MATERIALS SCIENCE

and Martin theory and to the Wagner, Lifshitz and Slyozov (WLS) theory using an
irradiation-enhanced diffusion coefficient. The asymptotic WLS distribution was taken as a
starting point. Material and irradiation parameters are for 10-2 dpa/s HVEM irradiation of
an Al-la/oZn alloy at 400 K.
For small particles, the size distribution function is considerably wider than in the WLS
case. Above 10 nm the size distribution becomes very narrow and radiation slows coarsening
until at about 15 nm both treatments give the same distribution function and mean radius.

10. ALUMINUM-BASED ALLOYS


Aluminum-based alloys are used in a wide variety of structural applications. Aluminium
has a substantial solid solubility for many other metals. The solubility usually decreases with
temperature, which makes these alloys condidates for strengthening by precipitation hard-
ening. Usually heat treatment of Al-based alloys produces not one, but a sequence of phases
(Russell and Aaronson, 1976) which must be manipulated, often through complex time-
temperature-mechanical deformation cycles, so as to give such desired properties as strength,
ductility, and corrosion resistance. In that the size, morphology, distribution, and identity of
the phases in these alloys are all very sensitive to alloy composition, heat treatment, and
deformation, one might expect they would be altered by irradiation.

10.1. AI-Ag and Al-Ag-Zn Alloys


Cauvin and Martin (1983a) studied the response of several binary Al-based solid solutions
to 1 MV HVEM irradiation at displacement rates from 3 × 10-4-10 -2dpa/s. Irradiation
temperatures ranged from 25 to 260°C.
Irradiation of alloys of 0.1, 0.2, and 1.2 a/oAg triggered homogeneous precipitation of a
metastable Ag-rich phase. This phase formed readily either above or below the equilibrium
thermal solvus. This behavior was explained on the basis of the undersized Ag atom diffusing
as part of an interstitial dumbbell, to give point defect recombination-driven precipitation of
the type discussed in Section 4.2. The metastable precipitates observed showed significant
coherency with the matrix, so the Cauvin-Martin model for irradiation-induced precipitation
of coherent particles should be applicable.
K. S. Liu et al. (1972) irradiated Al-la/oAg and A1-5a/oAg-5a/oZn alloys to 3.2 × 1015-
1.5 x 1016 fast n/cm 2 ( ~ 2 x 10-6-10-Sdpa) at T < 20 K. The alloys had been given various
aging treatments to produce GP zones with Guinier radii of 2.5-6 nm. The experiment was
designed to determine the effect of neutron irradiation on the stability of GP zones at
temperatures where neither self-interstitials nor vacancies are mobile.
Irradiation to 1.5 x 1016 n/cm 2 gave approximately a 10~ decrease in Guinier radius for
A1-Ag alloys with initial radii of 4.4 or 5.5 nm. The A1-Ag-Zn alloy contained both Ag and
Zn zones; irradiation of a specimen with zones in the 2.5-5 nm range to 3.2 × 1015fast n/cm 2
gave decreases in the Guinier radius of 0.5-6 nm, with the lesser changes occurring with the
smaller zones.
The results on both the AI-Ag and A1-Ag-Zn alloys were attributed to displacement
cascades partly or completely dissolving the GP zones. The smaller zones could be completely
dissolved. Since a small zone size implies a small inter-zone spacing, energetic atoms removed
from these zones could wind up at another zone and irradiation might have little effect on
the average zone radius. Large zones would be only partially dissolved and the removed
atoms would not have enough range to reach another zone. Accordingly the average zone
size would decrease significantly.
PHASE S T A B I L I T Y U N D E R I R R A D I A T I O N 279

The maximum damage level, however, was very low to have caused a significant amount
of recoil resolution from the zones (Section 7).

10.2. AI-Cu Alloys


Alloys which contain from ca. 1-4 w/o copper may be age-hardened to high strengths, and
are widely used in structural applications. Different heat treatments may bring out any of four
different phases. These are:

GP-I: coherent Cu platelets 1-2 atoms thick lying on {100} matrix planes.
0" (also called GP-II): coherent Cu-rich platelets on {100} matrix planes.
0': platelets of approximate composition CuAI 2 lying on and coherent across { 100} matrix
planes, incoherent around their circumference.
0: equilibrium phase of CuA12, largely incoherent with the matrix.
Figure 33 shows the Al-rich end of the AI-Cu phase diagram with solvi for all four
phases. Various heat treatments may produce complex precipitation sequences. The order
of appearance of the phases is required by thermodynamic considerations to be
GP-I--,0"--,0'-,0, although one or more phases may not appear for a particular alloy and
heat treatment.
Many studies have been made of irradiated AI-Cu alloys. These include elevated
temperature irradiations by electrons and neutrons and post-irradiation annealing studies
after irradiation by electrons, neutrons and deuterons.
Sklad and Mitchell (1974, 1975) irradiated solution treated and quenched alloys in a 650 kV
high voltage electron microscope (HVEM). In the 1974 study samples of an A1-3.98 w/oCu
alloy were irradiated to 2dpa at 120°C at 4 x 10-4dpa/s and to 18dpa at 100°C at
5 × 10-3 dpa/s. Time in the microscope hot stage without the beam on raised the total heating
time to ~ 180 min for both sets of samples. A control sample was also heated for 180 min
at 120°C in the absence of irradiation. The starting state for all samples was GP-I in an Al-rich
matrix. The control sample showed no further precipitation while both irradiated samples
showed 0' formation with more precipitation in samples exposed to the higher flux and
fluence. The 0" phase, intermediate in stability between GP-I and 0', was not formed. The

600

o
~.
400

200
,
S ~ e" $olvus

t J I I I
I 2 3 4 5
Wt. % Cu

FIG. 33. Aluminum-rich end of A1-Cu phase diagram showing solvi for GP zones, the transition
phases 0" and 0', and the equilibritzm 0 phase. (After Russell and Aaronson, 1975.)
280 PROGRESS IN M A T E R I A L S SCIENCE

enhanced formation of 0', was attributed to irradiation enhanced diffusion and to formation
of small defect clusters which served as nucleation sites.
Sklad and Mitchell (1975) irradiated both solution treated and quenched, and quenched
and aged A1-3.50w/o Cu in the 650kV HVEM at 3 × 10-4dpa/s and 5 x 10-3dpa/s at
temperatures from 20 to 200°C. In all cases the samples were heated for 30 min for
equilibration before irradiation. Irradiation of the quenched samples to 9-18 dpa between
20°C and 150°C gave precipitation of 0' and disappearance of the GP-zones. High flux
irradiation enhanced the formation of 0' more than low fluxes. At the lower temperature 0'
does not form under thermal aging. Figure 34 illustrates the effect of high flux irradiation
at 100°C on a solutionized specimen. Figure 34a shows that no precipitates existed prior to
irradiation though some streaking in the diffraction pattern suggests the presence of GP
zones. Figure 34b, taken after 9 dpa, shows a high density of ~ 95 A diameter precipitates
aligned along {110} matrix planes. The diffraction pattern is characteristic of 0' precipitates.
Figure 34c, taken after 18 dpa shows that the particles are larger and the maxima along (100)
directions in the diffraction pattern are better developed.
A sample pre-aged at 165°C was irradiated at 150°C at a high flux to up to 45 dpa and
a sample pre-aged at 190°C was irradiated at high flux at that temperature. In both cases 0"
disappeared and was replaced by 0'.
The results of both studies show that electron irradiation causes dissolution of GP zones
and 0" and accelerates formation of 0'. Acclerated formation of the 0' was attributed to
irradiation-enhanced diffusion and nucleation sites provided by defect clusters. Growth rate
measurements of 0' indeed showed an enhanced diffusion coefficient below ~ 200°C.
Figure 35 shows the coarsening rate of 0' precipitates with and without irradiation.
Electron irradiation at 5 x 10 -3 dpa/s is seen to give a dramatic increase in the coarsening
rate at 100, 120 and 150°C.
The enhanced dissolution of GP zones and 0" and prevention of 0" nucleation were
attributed to irradiation "mixing", to irradiation-altered thermodynamic stability, or to
alteration of the nucleation process.
Irradiation could destabilize 0" by disordering it, as was discussed in Section 8 and was
observed by Potter (1981) in a Ni-Si alloy (Section 14). Irradiation is certainly capable of
causing significant changes in nucleation mechanisms (Section 4) but, as was shown in Section
3, is without much effect on solution thermodynamics.
Weaver et al. (1977) studied the effects of both HVEM and or-particle irradiation on
precipitation in an A1-3.8 w/oCu alloy which had been solution treated and quenched to form
GP zones. Some samples were irradiated at 190°C and precipitation observed in situ; others
were aged thermally at 190°C, then subjected to room temperature electron irradiation to
0.45-9 dpa, and with room temperature s-particle irradiation to 10 appm He. The HVEM
had a displacement rate of 1.5 x 10-2/s, which is much higher than in the experiments of
Sklad and Mitchell and would be expected to affect precipitation at higher temperatures.
All irradiation treatments promoted the rapid formation and increased number density of
0' particles at the expense of the pre-existing GP zones and without formation of the
intermediate 0" phase. All irradiated specimens showed an increased coarsening rate of the
0' phase, with the greatest increase in samples which were electron irradiated at 190°C. These
samples contained only a modest concentration of 0' particles, and would maintain high point
defect concentrations and a higher diffusion coefficient than those with a higher particle
number density. Enhanced coarsening of the specimen electron irradiated at room tem-
perature and thermally aged was interpreted as growth of 0' particles in the irradiated region
at the expense of the surrounding GP zones.
PHASE STABILITY UNDER IRRADIATION 281

FIG. 34. Electron micrographs and selected area diffraction patterns showing the effect of high flux
irradiation at 100°C on a solutinized specimen of A1-3.5~Cu. Doses are: (a) 0 dpa; (b) 9 dpa; and
(c) 18 dpa. (Electron micrographs courtesy P. S. Sklad.)
282 PROGRESS IN MATERIALS SCIENCE

i6106 x
- Irrodiated at /o
150%/

o,,~ 12

,..-, -
{0 o
'E
~G
E 8

~. 6

4 0t 120'C

-~o~.----'ff~,r~Jio,e, o,,oo.e
x-~ ,Ag~,,,:~ o.__~t
,sp.c
1.0 2.0 3.0 4.0
Irradiation time, t l. (Hrs)

FIG. 35. Coarsening of 0' without irradiation (bottom line) and during electron irradiation at
5 × 10 -3 dpa/s. Irradiation is seen to give a markedly enhanced coarsening rate. (After Sklad and
Mitchell, 1975.)
PHASE STABILITY UNDER IRRADIATION 283

The preferential precipitation of 0' particles was explained as irradiation-induced defects


somehow relieving coherency strains and enabling the GP zones to transform to the 0' phase.
This mechanism is at least qualitatively similar to that of nucleation of 0' on irradiation-
produced defect aggregates advanced by Sklad and Mitchell (1974, 1975).
Tucker and Webb (1959) irradiated solution treated and quenched AI-2a/oCu alloys at
- 80°C with 1.4 MV e- to approximately 10 -5 dpa. A resistance increase of approximately
1~o during subsequent aging was interpreted as being due to solute clustering facilitated by
irradiation-induced point defects. The authors estimated that each defect would make about
105 jumps in its lifetime, so that each solute atom would jump approximately once.
Carpenter and Yoo (1978) irradiated an A1-3.8w/oCu alloy at 328 K to 2.8 x 1022n/cm 2,
E > 0.1 MeV (,-~ 14 dpa). The alloy had been preaged to form 0' precipitates. The particle size
distribution of 0' did not change during irradiation, though there was some loss of coherency
on the broad faces of the 0' particles.
Several investigators studied the effects of neutron irradiation on subsequent annealing
kinetics. Pieragostini et al. (1966) used AI-4w/oCu alloys which had been solution treated,
quenched, and aged 24 hr at room temperature, annealed for 1 min at 200°C (reversion) and
stored in liquid N2. After reversion the alloys were expected to be a uniform solid solution
Samples were irradiated to 1.7 x 1017 n/cm 2 (E > 1 MeV) or ,-~ 10 - 4 dpa at - 195°C, and
irradiated and non-irradiated samples were given isochronal anneals from -200°C to 300°C.
The irradiated samples showed a stronger resistivity peak at ~ 80°C than did the non-
irradiated sample, which was interpreted as irradiation-enhanced clustering of Cu into zones.
Katz et al. (1968) studied an Al-lw/oCu alloy which had been solution treated and
quenched into water at 25°C. Samples were irradiated at 40°C to 1.8 x 10'4-1.2 x 1017 fast
n/cm 2 ( ~ 1 0 - 7 - 1 0 - 4 dpa). Annealing kinetics of irradiated and irradiated alloys were mea-
sured at 110 or 130°C, where the 0' phase is expected to form. Irradiated alloys showed a
more rapid resistivity decrease, indicating accelerated formation of the 0' phase. The
enhancement in 0' formation did not depend on neutron fluence over the range studied. The
absence of a fluence effect suggests that some extraneous cause, such as unintentional cold
work, may have enhanced 0' formation.
Al-4w/oCu samples were irradiated by Katz et al. (1968) to 1019 fast n/cm 2 ( ~ 5 x 10 -3 dpa)
at a temperature stated as being well below 100°C. Subsequent annealing of irradiated and
unirradiated samples at 200°C showed a marked enhancement of 0' formation.
Vogl and Weiss (1965) compared the annealing kinetics of irradiated and unirradiated
A1-5.17w/oCu. Irradiation to 10TMn/cm 2 (E > 1 MeV) or ,-~5 x 10 - 4 dpa gave an accelerated
rate of resistivity decrease on annealing in the 160-220°C range, which was interpreted as
enhanced 0' formation. Although the rate of 0' formation was increased by irradiation, the
activation energy was unchanged. The lack of change in activation energy shows that while
irradiation altered the rate of the process, it did not alter the process itself.
Katz and Herman (1967) studied the effects of 10.5 MeV deuterons on precipitation in a
reverted Al-l.7a/oCu alloy. Samples were irradiated near 77 K to 2.5 × 1015 d/cm / and
5 × l015 d/cm 2 (,-~2.5 - 5 × l0 -5 dpa). Resistivity changes at 25°C were essentially the same
for irradiated as for unirradiated alloys which indicates that irradiation did not change the
kinetics of GP zone formation. The authors explained this negative result on the basis of
displacement cascades not providing nucleation sites for GP zones, which are fully coherent
with the matrix. This analysis is consistent with the calculations of Marth et al. (1976) who
show that whereas semi-coherent phases often form on heterogeneities, coherent phases
usually form by homogeneous nucleation or spinodal decomposition.
Miura et al. (1975) used 1.5 MeV y-rays in studies of A1-3w/oCu and the commercial 2017
284 PROGRESS IN MATERIALS SCIENCE

alloy, which contains (in w/o) 4.42 Cu, 0.49 Mg, 0.52 Mn, 0.58 Si and 0.50 Fe. Prior to
irradiation the alloys were solution treated, quenched to room temperature, and aged to a
fully hardened state at 190°C. Irradiation for 25 h at 25°C was estimated to give only
~ 3 x 10-12 dpa (,-~3 x 10-17 dpa/s), but that value would produce three orders of magnitude
more vacancies than the thermal equilibrium concentration.
A transmission electron microscope (TEM) study showed that room temperature irra-
diation of the 2017 alloy for 25 hr gave an order of magnitude decrease in the precipitate
number density, while similar treatment of an Al-3 w/oCu alloy gave no observable
microstructural change. These results were interpreted on the basis of a strong binding
between vacancies and the Mg atoms in the 2017 alloy which gave the vacancies a longer
lifetime. Certainly the few vacancies produced by this very mild irradiation must be used very
effectively to produce any microstructural change.

10.3. AI-Ge Alloys


The A1-Ge system (Fig. 36) is particularly well-suited for a study of vacancy-assisted
precipitate nucleation in that the system is of the simple eutectic type, with no intermediate
phases and limited solid solubility of Ge in A1. The solid solubility of AI in Ge is extremely
small. The atomic volume of elemental Ge is 36~o larger than that of the A1 solid solution,
so that for about each three Ge atoms precipitated, one vacancy must be consumed so as
to prevent strain energy effects. A vacancy supersaturation would thus make a large
contribution to AG ° for nucleation [see Eq. (4.4)]. Betram et al. (1978) studied the effect of
A1÷ ion irradiation on the nucleation of Ge precipitates from an A1-Ge solid solution. Foils
of an A1-2w/oGe solid solution were irradiated with A1÷ ions at displacement rates from
3.7 x 10-3 s -l to 3.7 × 10 -1 s-'. Irradiations were conducted at temperatures of 25 and
125°C, to doses ranging from 10 to 120 dpa.
Figure 37 is typical of the results of the investigation. The upper right region of the foil
was exposed to the AI ÷ beam, and formed a high density of Ge precipitates. The lower left

WEIGHT PER CENT GERMANIUM


10203040 50 60 7075 80 85 90 95
IO00 II III I I I l II I I II I I i [I ~r ] I I r

900 //~
800 . ~
~"700 660°
~6oo~. /~
5oo ~_
I-- 4 0 0 ' 0 ~2"8(7"2) ~ 424°

30O F(A'e)

200 r
I ool 1 I I 1 J I I i l
0 I0 20:50 40 50 60 70 80 90 I00
A,e ATOMIC PER CENT GERMANIUM Ge

FI~3. 36. Aluminum-germanium phase diagram. (After Hansen, 1958.)


PHASE STABILITY UNDER IRRADIATION 285

region of the foil was shielded from the beam by a copper grid, and shows only a few
precipitates, which formed on the foil surface. The supersaturated vacancies have clearly had
a strong catalytic effect on the nucleation of Ge precipitates.
Such effects were seen in a number of cases, particularly for high displacement rates and
thicker foils, both of which conditions favor higher vacancy supersaturations and more rapid
nucleation. Precipitate number densities in irradiated foils were typically in the 101s-1016/cm3
range.
Rusbridge (ne6 Bertram) (1981) studied the effects of 1 MV HVEM irradiation on
precipitation in A1-Ge alloys. Foils of 0.22, 0.47, 0.98, and 1.90w/oGe were irradiated at
temperatures from ~25°C to 220°C at a displacement rate of 6 × 10-3dpa/s to doses of
0.7-20 dpa. The samples were examined periodically in the HVEM during irradiation and in
a 100 kV TEM afterwards.
Copious Ge precipitation occurred in the 1.90w/oGe foil during irradiation at temperatures
from ~25°C to 175°C. Precipitate number densities were over 5 × 10tr/cm 3 at 25°C and
decreased to 1.2 × 1016/cm3 at 175°C. Particle number densities were lower in the more dilute
alloys. Precipitation was observed in the 0.22w/oGe alloy at temperatures as high as 140°C,
where the particle number density was 0.6 × 1016/cm3.
Rusbridge compared her results with the chemical vacancy model described in Section 4.1,
and found agreement between theory and experiment if a nucleus:matrix surface energy of
approximately 230 mJ/m 2 were assumed. This value is well below the ~700 mJ/m 2 value
estimated by Belier (1973) for incoherent interfaces involving much larger Ge particles.
Rushbridge suggested that the nucleus had some coherency with the matrix, so the lower
surface energy was not unreasonable. Aaronson and Russell (1983) surveyed the most recent
results of high resolution TEM studies in nucleation in solids and concluded that the critical
nucleus in nearly all cases showed a high degree of, if not complete, coherency with the matrix.
As such, a 230 mJ/m 2 surface energy is reasonable for the critical nucleus:matrix interface.
Rusbridge (1981) studied solute segregation in an AI-2w/oGe alloy under 150 keV A1÷ ion
irradiation. Irradiation was to 18, 54, and 90 dpa at 160°C at a displacement rate which was
unspecified but probably in the 10-3-10 -l dpa/s range. After 18 dpa, the solute was enriched
by ~ 25~o at the surface and depleted by the same amount about 70 nm below the surface.
At 90 dpa the surface enrichment was ~ 100~ and the depletion at 70nm depth was
approximately 50~o. Germanium is thus segregated toward the surface, a defect sink. The
segregation is very weak compared to that observed in other systems, particularly Fe and Ni
based alloys. In addition Ge is oversized in AI, so the direction of segregation in this dilute
solid solution is opposite to that predicted by the size misfit theory (Section 5.1).
Since Ge is enriched at point defect sinks, solute segregation may play a role in irradiation
enhanced precipitation in AI-Ge alloys. The degree of segregation is however too modest to
be a major factor in the huge increases in precipitation rate caused by irradiation.
Bertram et al. (1978) held samples for long periods at the various irradiation temperatures
but without irradiation, to give adequate time for diffusion. If irradiation were simply
increasing the diffusion coefficient, the extended thermal treatment would give a similarly high
precipitate density. Only a low precipitate density was observed, so, irradiation-enhanced
diffusion was not the mechanism for enhanced nucleation.
The results of Bertram et al. (1978) and Rusbridge (1981) are consistent with the
experiments of Belier (1973) on quenched A1-Ge alloys. The quenched in vacancies were
found to give increases in the precipitate number density comparable to those obtained under
irradiation.
Cauvin and Martin (1983a) studied the response of Al-based alloys of 0.1 and 0.3a/oGe
286 PROGRESS IN M A T E R I A L S SCIENCE

to 1 MV HVEM irradiation at temperatures of 25-260°C and displacement rates of


3 × 10-4-10 -2 dpa/s. Precipitation occurred below, but not above, the thermal solvus. The
absence of precipitation above the solvus was attributed to the oversized Ge atoms not being
incorporated into dumbbell interstitials, so that the Cauvin-Martin model for irradiation-
induced precipitation in undersaturated solutions was not operative. However, since one
vacancy must co-precipitate for every ~ 3 Ge atoms to avoid a prohibitive strain energy, the
chemical vacancy model described in Section 4.1 could have been operative. However, since
at 125°C the thermal solvus lies at 0.05 a/oGe, the 0.1 and 0.3 a/oGe alloys would have to
be irradiated well above this temperature to test for irradiation-induced precipitation. At such
temperatures the vacancy supersaturations will be relatively low and would have only a minor
effect in promoting precipitation. As such, the absence of precipitation in undersaturated
A1-Ge alloys is not surprising.
Rusbridge (1981, 1983) induced the crystalline-amorphous transformation in Ge precip-
itates in an Al-matrix with either 1 MV e- or 100 KeV A1+ ions. Electron irradiation at
- 130°C at 1.9 x 10 -2 dpa/s (in the Ge) gave a gradual decrease in the Ge diffraction spot
intensity until at ~ 1023e-/cm2 (,-~0.2 dpa) the spots disappeared entirely. Figure 38 shows the
progressive amorphization of Ge precipitates during HVEM irradiation at -130°C. Dis-
placement cascades are not formed by e- irradiation, so they could not be responsible for
the amorphization. Rusbridge interpreted the amorphization as due to a critical vacancy
concentration of ~0.2.
Irradiation at room temperature to 1015Al+/cm 2 (,,~ 1.8 dpa) also rendered the precipitates
amorphous. In this case loss of crystallinity was attributed to the disruptive effects of
displacement cascades in the Ge precipitates.

10.4. AI-Li Alloys


Shiraishi (1970) studied alloys of A1-2.2w/oLi which had been annealed at 350°C and
slowly cooled and were thus in the overaged condition. (The solvus temperature of this alloy
is between 400 and 500°C.) The samples were irradiated at 45°C to 1.2 x 1019 n/cm 2, or about
5 x 10-3 dpa.
Irradiation increased the yield strength of the alloys from ,,~50 MPa to ~ 100-150 MPa.
Post-irradiation annealing at temperatures of 155, 170 and 191°C first gave a small decrease
in room temperature yield strength, followed by an increase to ca. 150 MPa. The yield
strength of non-irradiated samples was not changed by the various annealing treatments.
Shiraishi interpreted the initial decrease in yield strength as due to annealing of irradiation-
induced defects. These defects would have to be individual vacancies and intersittials, since
even neutron irradiation does not give displacement cascades in A1 or Al-based alloys
(Olander, 1975).
The later increase in yield strength was attributed to dislocation pinning by a fine
distribution of A1Li precipitate particles. The solute for the particles was hypothesized to have
come from large precipitate particles which had been destroyed by irradiation.
The calculations given in section 7 showed that sizeable precipitates will be appreciably
dissolved by irradiation only at very high dpa. There was, however, a substantial amount of
Li in solid solution which did not have time to diffuse to the widely-spaced precipitate
particles during the slow cool from 350°C. Irradiation enhanced diffusion would then allow
nucleation of a fine dispersion of particles, either during irradiation or in the early stages of
post-irradiation annealing. These closely-spaced particles could grow during post-irradiation
annealing to give the observed increase in yield strength.
PHASE STABILITY U N D E R I R R A D I A T I O N 287

FIG. 37. A1-2w/oGe irradiated at 80°C to 5dpa with 100keV AI ÷ ions at 3.7 x 10-3dpa/s. The
lower left part of the micrograph shows a part of the sample which was shielded from the A1÷
beam by a copper grid. (Electron micrograph courtesy K. Rusbridge.)
288 PROGRESS IN M A T E R I A L S SCIENCE

Om 2m 5m
FIG. 38. (a) Dark field micrograph recorded at RT during post-irradiation examination of an AI-Ge
alloy. The whole sample had been uniformly seeded by ion irradiation at 140°C and the central
region had been subsequently irradiated with 1 MV electrons at -130°C to a dose of ~ 0 . 2 d p a
at 1.9 x 10-2dpa/s. The absence of a dark field image in this region indicated that electron
irradiation had produced a c-a transition within the precipitates.
Co) Sequence of SADP's recorded in situ in the HVEM during 1 MeV electron irradiation of area
shown in (a) at -130°C showing gradual decrease in second phase Ge reflections.
(Electron micrograph and diffraction patterns courtesy K. Rusbridge.)
PHASE S T A B I L I T Y U N D E R I R R A D I A T I O N 289

FIG. 39. Dark field micrograph of precipitate reflections showing Mg2Si precipitation at grain
boundary and in grains, Alloy was A1-2.2w/oMg prior to neutron irradiation. (Photomicrograph
courtesy R. T. King.)
290 PROGRESS IN M A T E R I A L S SCIENCE

FIG. 40. Commercially pure aluminum irradiated at 328 K to 1.4 x 1023n/cm 2 (E > 1 MeV) and
2.3 x 1023n/cm 2 (E < 0.025 eV) to give about 100 dpa and 5.3a/oSi. Many fine Si precipitates have
formed, some associated with irradiation induced voids. (Electron micrograph courtesy K. Farrell.)
PHASE S T A B I L I T Y U N D E R IRRADIATION 291

FIG. 41. Effect of irradiation on Si precipitate nucleation in a Al-l.4w/oSi alloy air cooled from
540°C and irradiated in the electron microscope, g = [I11], (211) foil orientation. (a) 200°C, 190 min;
(b) 250°C, 210min; (c) 250°C. Unexposed area. (Electron micrograph courtesy M. Nemoto.)
292 PROGRESS IN MATERIALS SCIENCE

FIG. 42. A1-0.5w/oSi after 1 MV electron irradiation to 20 dpa at 165°C. Bright field (left) and dark
field (right). (Electron micrograph courtesy G. L. Kulcinski.)
PHASE STABILITY UNDER IRRADIATION 293

10.5. A I - M g and A I - M g - S i Alloys

Cauvin and Martin (1983a) irradiated Al-based alloys of 1.2, 2.1, 3.5, and 4.0 a/oMg in
the 1 MV HVEM at temperatures between 25 and 260°C at displacement rates between
3 x l0 -4 and l 0 -2 dpa/s. Irradiation did not induce precipitation in thermally single phase
alloys, and in fact had no effect on precipitation kinetics in supersaturated alloys. The lack
of irradiation-induced precipitation was attributed to the oversized Mg atom not being
incorporated in self-interstitial dumbbells, so that the point defect recombination-driven
precipitation would not occur.
Under thermal conditions the intermetallic AlaMg2 fl-phase precipitates from super-
saturated Al-based A1-Mg solid solutions. The fl-phase has a larger atomic volume than does
the Al-based matrix (3 ~ 0.15) so that according to Section 4.1 its formation should be
promoted by a vacancy supersaturation. However, Mg is highly oversized in A1 (D.~f= 41%)
and should be rejected strongly from the particle: matrix interface. The two effects thus would
tend to neutralize one another.
Irradiation with thermal neutrons triggers a novel phase instability in A1 and Al-based
alloys. Aluminum has a transmutation cross section of 0.23 × 1 0 - 2 4 c m 2 for neutrons of
0.025 eV energy to produce Si (Farrell et al., 1977). As such, a reactor with a relatively high
thermal flux may readily produce 1% or more Si by transmutation of Al atoms.
King and Jostsons (1975) made a neutron irradiation study of the commercial 5052 Al alloy
(Al plus 2.2w/oMg and small amounts of other metals). The annealed, solid-solution alloy
was irradiated at 55°C to 3.6 × 10 22 n/cm 2 (E > 0.1 MeV) and 5.8 × l 0 22 n/cm 2 (E < 0.414 eV)
in the High Flux Isotope Reactor at Oak Ridge National Laboratory. The irradiation
produced approximately 20 dpa and lw/oSi. Silicon is relatively insoluble in Al-based alloys,
and in this case precipitated out as a fine dispersion of fl' (MgESi) particles as seen in Figure
39. The primary role of irradiation was to produce Si, as a similar microstructure could have
been developed thermally in alloys of Al-lw/oSi-2.2w/oMg.
Ostrovsky et al. (197 l) studied the effects of neutron irradiation on precipitation hardening
in an A1-0.48w/oMg-0.81w/oSi alloy. The alloy was solution treated and quenched, then
subjected to 5.4 × 10 2° n/cm 2 (E > 1 MeV) at temperatures of 50-60°C and 110-130°C to give
~ 3 dpa. Aging of an unirradiated sample at 230°C gave hardness peaks after 1 and 28 hr.
The first hardness peak occurred in the irradiated alloys almost immediately and the second
after 16-20 hr. Microstructural studies showed that the irradiation had accelerated the
precipitation process.
The authors did not propose a mechanism for the accelerated precipitation. Irradiation-
enhanced diffusion is probably not the cause since the aging temperature is so high (0.54 Tin)
that irradiation-induced vacancies would escape almost immediately. It is more likely that
irradiation assists the nucleation process, possibly by providing catalytic sites for formation
of nuclei.
Vaidya (1979) irradiated an aged AI-0.9w/oMg-0.6w/oSi alloy with 250 keY A1÷ at room
temperature. The alloy contained semicoherent fl' phase precipitates which initially had an
array of misfit dislocations at the particle:matrix interface. By 44dpa the interfacial
dislocations had disappeared and the fl' particles had begun to dissolve.
The loss of coherency appears to be of the same kind observed by L. M. Brown et al. (1968)
in Cu-Co alloys (Section 11). Irradiation induced dissolution could be the result of a recoil
resolution process. (Section 7) but is more probably due to irradiation-enhanced diffusion.
Magnesium is oversized in A1 while Si is undersized, so irradiation induced solute segregation
could also be operative.
294 PROGRESS IN M A T E R I A L S SCIENCE

10.6. A I - S i Alloys
As in the A1-Ge system, quenched-in vacancies greatly enhance the nucleation of the
diamond cubic Si phase in Al-rich alloys (Ozawa and Kimura, 1970; Russell, 1980).
Co-precipitation of vacancies with Si has been proposed as a mechanism (Russell, 1980),
whereas Rosenbaum and Turnbull (1959) and Ozawa and Kimura (1970) preferred to
attribute the increased precipitate number density to heterogeneous nucleation of Si particles
on prismatic loops which had been formed by vacancy condensation. In addition, since Si
is under-sized in A1 and known to segregate to point defect sinks, a mechanism which involves
solute segregation to sinks followed by precipitate nucleation is also possible.
Farrell et al. (1977) irradiated l l00-grade (commercially pure) aluminum at 328 K to
1.4 x 1023n/cm2 (E > 1 MeV) and 2.3 × 1023n/cm 2 (E < 0.025 eV) to produce about 100 dpa
and 5.3a/oSi. the solubility of Si in A1 at the irradiation temperature is < 10-2a/o, and
diffusion is rapid enough to allow precipitation and solute segregation. A TEM study revealed
numerous fine Si precipitates, many of which were associated with irradiation-induced voids.
A STEM study showed the surfaces of the voids to be coated with a 4-11 nm thick layer of
Si. Figure 40 shows Si precipitation in the matrix and at void surfaces. Silicon is undersized
in the AI matrix, so segregation to voids and other point defect sinks is consistent with the
prediction of the atomic misfit model for solute segregation.
Nemoto et al. (1971) studied the effects of 500 kV electron irradiation in the HVEM on
both quenched and air cooled (to 50°C) samples of an Al-l.4w/oSi alloy. The former
treatment would produce a uniform solid solution, and the latter a few large precipitates with
a substantial amount of Si still in solid solution. Irradiation of the quenched alloy for 60 min
at room temperature or the air cooled alloy for 190 min at 200°C or for 210 min at 250°C
gave a high number density of Si precipitates.
The effects of the 200°C and 250°C irradiations are shown in Fig. 41, along with a view
of a control sample for comparison. Irradiation is seen to give a huge increase in the
precipitate number density. In contrast, the only thermal treatment which would give a high
precipitate density in this alloy is a quench to below room temperature, followed by an anneal
at ~ 100°C. The precipitates are clearly irradiation-induced. The e- beam current was not
stated, but one would expect the irradiation to have given the order of tens of dpa at a
displacement rate of ~ 1 0 - 3 / s .
Tjhia et al. (1980) studied Si precipitation in A1-0. lw/o and A1-0.5w/o alloys during 1 MV
electron irradiation in the HVEM. Displacement levels of 0.4, 3, and 20 dpa were achieved
at 20, 110, and 165°C, respectively.
Figure 42 shows the results of irradiation of a 0.5w/o Si alloy to 20 dpa at 165°C. Both
bright field and dark field micrographs show a high density of very fine clusters.
Solute clusters of ,,, 50 ~, diameter formed at number densities of up to 3 x 1015/cm3, as
did both interstitial and vacancy-type dislocation loops. Vacancy loops usually form much
less readily than interstitial loops during irradiation (Russell and Powell, 1973). Tjhia et al.
however, concluded that a solute atmosphere of undersized Si atoms was stabilizing the
vacancy loops by reducing their bias for interstitial capture.
An AI-0.5w/oSi sample which had been irradiated to 3 dpa at 20°C was given post-
irradiation anneals at 50 and 100°C. Whereas no microstructural changes were observed after
the 50°C anneal, increasing the temperature to 100°C caused the dislocation loops to anneal
out in 10rain. The small clusters dissolved within 40 rain and at the same time large
equilibrium diamond cubic silicon precipitates formed.
The small clusters were associated with dislocation loops, and diffraction contrast
measurements indicated some degree of coherency with the matrix. A calculation of the
PHASE STABILITY UNDER IRRADIATION 295

activation barrier for nucleation was made, and it was concluded that the homogeneous
nucleation rate of incoherent Si particles was much too low to give the observed precipitate
number densities. Then, on the basis of the coherency of the clusters, their association with
dislocations, and the low calculated nucleation rates, it was concluded that homogeneous
nucleation by co-precipitation of vacancies and solute was not operative. Instead, the authors
felt that clusters were forming on dislocation loops, a mechanism proposed by Rosenbaum
and Turnbull (1959) and by Ozawa and Kimura (1970). The association of the clusters with
dislocation loops is indeed persuasive evidence for heterogeneous nucleation of the Si
particles.
However, disappearance of the high density of small clusters on annealing does not occur
in quenched AI-Si alloys and has not been reported in other irradiation studies. It is thus
possible that the formation of the coherent clusters on dislocation loops was peculiar to this
investigation. Disappearance of the fine precipitates is to be expected since highly coherent
particles have a higher solubility than do largely incoherent particles, and will dissolve in
favor of the latter.
Tjhia et al. did not state what value was taken for the nucleus:matrix interfacial energy.
Aaronson and Russell (1983) have argued that the critical nucleus will in virtually all cases
show some degree of coherency with the matrix and that the surface energy appropriate to
nucleation will be substantially less than that for macroscopic interfaces.
Cauvin and Martin (1983a) irradiated A1 alloys of 0.05 and 0.5 w/o Si in the 1 MV HVEM
at temperatures from 25 to 260°C and displacement rates from 3 × 10 -4 to 10-2 dpa/s. Results
were not reported in detail, but it was noted that irradiation-induced precipitation above the
thermal solvus probably occurred.
As in the case of AI-Ag and A1-Zn alloys, such precipitation was attributed to the
undersized Si atoms being incorporated in interstitial dumbbells and giving point defect
recombination-driven precipitation (Section 4.2). However, since, to avoid strain energy,
vacancies must coprecipitate with Si atoms, the chemical vacancy effect discussed in Section
4.1 is operative. It is possible that in this case the two mechanisms operated simultaneously
to give precipitation above the thermal solvus.
The conflict of the results of Nomoto et al. (1971) and Cauvin and Martin (1983a) with
those of Tjhia et al. (1980) indicates the need for more studies on the response of AI-Si alloys
to electron irradiation.

10.7. A l - Z n Alloys
Suitable quenching and aging treatments will produce Zn-rich GP zones in Al-rich AI-Zn
alloys. The progress of zone formation is commonly followed by either small-angle x-ray
scattering or electrical resistivity measurement. Quenching the alloy from ca. 500 K is
necessary to produce the vacancies needed for mass transport in zone formation under
thermal ageing conditions. Several investigations have instead employed irradiation to
provide the point defects needed for zone formation.
Herman (1964) irradiated a reverted A1-5.3a/oZn alloy at 77 K to up to 13.3 × 10 L5
10.5 MeV deuterons/cm:. He estimated from resistivity measurements that approximately
102°d/cm2 produced 1 dpa, so that his maximum damage level was about 10-4dpa. On
post-irradiation annealing at 20°C, irradiation was found to accelarate zone formation and
to produce slightly larger zones (11 A vs. 9 A at the highest displacement level).
Cerasara et al. (1964) and Horak et al. (1968) studied the effects of neutron irradiation on
zone formation. The former irradiated both reverted and quenched Al-10w/oZn alloy to
296 PROGRESS IN MATERIALS SCIENCE

2.9 × 10~Tn/cm 2 ( E > I M e V ) at 78K, to produce approximately 10-4dpa, or 10-4at.


fraction vacancies and interstitials. On post-irradiation annealing between - 4 0 ° C and 0°C
the irradiated alloys had an enhanced rate of zone formation. The effect was attributed to
irradiation-induced vacancies; interstitials were believed to have escaped during the 78 K
irradiation.
Horak et al. irradiated both reverted and quenched alloys of 1.7-5.27a/oZn at 4.5 K to
doses of 2.3 × 1017-3.4 × 1017 n/cm 2 (E > 0.5 MeV), or about 1-2 × 10-4dpa. There was no
evidence of zone formation during irradiation. However, the alloys showed an increased rate
of zone formation on post-irradiation annealing at about 250 K which was attributed to
irradiation-induced vacancies.
Schiile (1964, 1969) disagreed with the attribution of enhanced zone formation to
vacancies, arguing that self-interstitials were at least in part responsible. A particularly cogent
argument is that interstitials can give mass transport, and had they annealed out during
irradiation, they would have produced zones. Horak (1968) disputed Schiile's contention,
claiming that the vacancy concentration was adequate to give zone formation and that there
is no need to invoke mass transport by interstitials.
The proper value of the interstitial migration energy in aluminum has been a matter of
controversy for years, with some authors supporting a migration energy close to that of the
vacancy, and others a much lower value. It appears that the argument over the mechanism
of zone formation in irradiated A1-Zn alloys is in large part an extension of this larger
controversy.
Cauvin and Martin (1979, 1981b, 1982) made extensive studies of irradiation induced
precipitation in A1-Zn alloys. In the earlier study a 1.9a/oZn alloy was subjected to 1 MV
e- irradiation at temperatures from 95 to 235°C. Electron fluxes from 5 × 1018 to
1.5 × 102°e-/cm2/s gave displacement rates from 3 x 10 -4 to 9 × 10-3/s. The 1.9a/oZn alloy
has a thermal solvus of 110°C.
The results of the study are shown in Fig. 43. Irradiation produced either coherent, Zn-rich
GP zones, the semi-coherent, hexagonal fl phase, or neither, depending on temperature and
displacement rate. The G P zones were found to be less stable than the equilibrium fl phase,
and disappeared on appearance of the latter.
Figure 43 is the first clear evidence of a displacement rate-dependent solvus. Samples were
irradiated to as high or higher doses at the lower rates as at the higher rates, so the dependence
is on rate rather than on dose. Irradiation has clearly increased the solvus temperature by
50-110°C over the thermal value.

"~I0
2 - /
ID
O~

I I I I
100 150 200 250
T, *C
FIG. 43. Irradiation conditions for the occurrenceof zinc precipitation in AI 1.9 at 9/ooZn. (tO) Guinier
Preston zones; (Or) /3 phase; (l-q) no particles observed; ( .) dose rate threshold for radiation
induced precipitation; (..... ) incoherent solvus temperature. (After Cauvin and Martin, 1979.)
PHASE STABILITY UNDER IRRADIATION 297

Table 2 Phase stability in aluminum-based alloys


Alloy Treatment Result Reference
AI4).l, 0.2, 1.2a/oAg 1 MV HVEM, 3 x 10 -4 to Precipitation of me- Cauvin and Martin
10-2 dpa/s, 25-260°C tastable Ag-rich (1983a)
phase above and be-
low thermal solvus
Al-la/oAg < 2 0 K , 1.5 x 1016 fast Decrease in size of Liu et al. (1972)
n/cm 2 ( ~ 10-5 dpa) GP zones
A1-5a/oAg-5a/oZn < 2 0 K , 3.2 x l015 fast Decrease in size of Liu et al. (1972)
n/cm 2 ( ~ 2 x 10-6dpa) GP zones
AI-3.98w/oCu 120°C, 2 dpa, 650 kV e-, Enhanced 0' for- Sklad and Mitchell
4 x 10-4dpa/s mation, bypass 0" (1974)
Al-3.98w/oCu 120°C, 2 dpa, 650 kV e-, Enhanced 0' for- Sklad and Mitchell
5 x 10 -3 dpa/s mation, bypass 0" (1974)
AI-3.50w/oCu 20-150°C, 9-18dpa, Enhanced 0' for- Sklad and Mitchell
650kV e-, 4 x 10-4dpa/s marion, bypass 0" if (1975)
or 5 x 10-3 dpa/s not already formed
AI-3.8w/oCu 20°C, 0.45-9dpa, l MV e- Subsequent ageing Weaver et al. (1977)
gives enhanced O"
nucleation and
growth
AI-3.8w/oCu 190°C, 135dpa, 1 MV e- Enhanced nucleation Weaver et al. (1977)
and growth of 0'
AI-2a/oCu < - 60°C, ~ 10- 5dpa, 1.4 Increase in resistivity Tucker, Jr. and
MVe- during aging Webb (1959)
A1-3.8w/oCu 328 K, 2.8 x 10:: n/cm 2 Partial coherency Carpenter and Woo
(E > 0.1 MeV),~ 14 dpa loss on broad faces (1978)
of 0' particles
AI-4w/oCu -195°C, 1.7 x l017 Enhanced clustering Pieragostini et al.
n/cm 2 (E > l MeV), on post-irradiation (1966)
~ l0 -4 dpa annealing
Al-lw/oCu 40°C, 1.8 x l014 to Enhanced 0' for- Katz et aL (1968)
1.2 x l017 fast n/cm 2, mation on post-
~ 10-7-10 -4 dpa irradiation annealing
AI-4w/oCu < 100°C, 1019 fast Enhanced 0' for- Katz et al. (1968)
n/cm 2, ~ 5 x 10 -3dpa mation on post-
irradiation annealing
A1-5.17w/oCu 30°C, 1018n/cm 2 (E > Enhanced 0' for- Vogl and Weiss
1 MeV), ~ 5 x 10-4dpa mation on post- (1965)
irradiation annealing
Al-l.7a/oCu 77 K, 2.5 x l0 I~ and No change in rate of Katz and Herman
5 x 10~ deuterons/cm 2 GP zone formation (1967
(10.5MeV), ~2.5 x 10 -5 on post-irradiation
-5 x 10-5 dpa annealing
AI-4.2Cu, 0.5 Mg, 25°C, ~ 3 x 10-t2dpa, 3 x Lower precipitate Miura et al. (1975)
0.5 Mn, 0.6 Si, 0.5 Fe (in w/o) 10-~Tdpa/s, 1.5 MeV number density
-rays
AI-2w/oGe 25°C, 125°C, 150 keV AI ÷, 10ts-1016Ge precipi- Bertram et al. (1978)
10-120dpa, 3.7 x 10-3 tates/cm 3
dpa/s, 3.7 × 10-I dpa/s
A1-0.22, 0.42, 25-220°C, 1 MV HVEM, > 1016Ge particles/ Rusbridge (1981)
0.98, 1.90w/oGe 6 x 10-3dpa/s, cm 3 in 1.9w/o alloy
0.7-20dpa up to 175°C, lower
densities in other al-
loys
A1-2w/oGe 160°C, 150keV A1+ Ge segregation to ir- Rusbridge (1981)
18-90 dpa radiated surfaces
AI~).I, 0.3a/oGe 1 MV HVEM 3 x 10-4 Precipitation in Cauvin and Martin
-10-2dpa/s, 25-260°C super-saturated al- (1983a)
loys only
Al-l.9w/oGe -130°C, 1 MV e-, 0.2 dpa, Ge particles became Bertram (1981,
1.9 x 10-2dpa/s (in Ge) amorphous 1983)
298 PROGRESS IN M A T E R I A L S SCIENCE

T a b l e 2. Continued.
Alloy Treatment Result Reference
AI-1.9w/oGe 30°C, 100 keV AI +, 1 dpa Ge particles became Bertram (1981,
1.4 × 10-2dpa/s amorphous 1983)
Al-2.2w/oLi 45°C, 1.2 × 1019n/cm 2, Post-irradiation an- Shiraishi (1970)
~ 5 x 10 3dpa neal at 155-191°C
increases yield
strength
AI-I.2, 2.1, 3.5, 4.0 w/oMg 1 MV HVEM, 3 × 10-4- Irradiation had no Cauvin and Martin
10 -2 dpa/s, 25-260°C effect on precip- (1983a)
itation
Al-2.2w/oMg 55°C, 3.6 x 1022n/cm 2, Precipitate fl' Mg2Si King and Jostsons
(initially) (E > 0.1 MeV), ~ 20 dpa, during irradiation (1975)
5.8 × 1022n/cm 2
(E < 0.414 MeV)
A1-0.48w/oMg- 50-60°C or 110-130°C, Accelerated precip- Ostrovsky et al.
0.81w/oSi 5.4 × 102on/cm 2 itation on post- (1971)
(E > 1 MeV), ~0.3 dpa irradiation annealing
AI-0.9w/oMg- ~25°C, 250 keV A1÷, fl' particles lose in- Vaidya (1979)
0.6w/oSi 44 dpa terfacial dis-
locations, start to
dissolve
AI (initially) 328 K, 1.4 x 1023n/cm 2 Fine Si precipitates Farrell et al. (1977)
(E > 1 MeV), 100 dpa, form under irra-
2.3 x 1023n/cm 2 diation; Si segre-
(E < 0.025 eV) gation to voids
Al-l.4w/oSi 60 min, 25°C; 190 min, Fine dispersion of Si Nemoto et al. (1971)
200°C; 210min, 250°C, precipitates
0 . 5 M V e - (~tens of dpa)
A1-0.1w/o and 20°, 110°, 165°C, High density of fine Tjhia et al. (1980)
0.5w/oSi 1 MV e- in HVEM, clusters form on dis-
0.4, 3, and 20 dpa location loops, dis-
appear during 100°C
anneal
A1--0.05, 0.5a/oSi 1 MV HVEM, 3 × 1 0 - 4 Precipitation above Cauvin and Martin
-10-2dpa/s, 25- the thermal solvus (1983a)
260°C
A1-5.3a/oZn 77K, up to 13.3 × 10Is Enhanced GP zone Herman (1964)
d/cm 2 (E = 10.5MeV) formation on post-
10- 4 dpa irradiation annealing
Al-10w/oZn 78 K, 2.9 x 1017n/cm: Enhanced GP zone Cerasara et al.
(E > 1MeV), ~10 4dpa formation on post- (1964)
irradiation annealing
A1 1.7-5.27a/oZn 4.5 K, 2.3-3.4 x 1017 Enhanced GP zone Horak et al. (1968)
n/cm: (E > 0.5 MeV) formation on post-
1-2 x 10 -4 dpa irradiation annealing
Al-l.9a/oZn 95-235°C, 1 MV e-, Dose rate dependent Cauvin and Martin
3 x 10 4-10 2 solvus up to 100°C (1979)
dpa/s or more above ther-
mal
A1 0.2--4.4a/oZn 100-250°C, 1 MV e-, Dose rate dependent Cauvin and Martin
6 × 10-4-1.5 x 10 -2 solvus, violation of (1981b)
dpa/s lever rule

Nucleation was homogeneous, not associated with point defect sinks or other hetero-
g e n e i t i e s , w h i c h m a k e s s o l u t e s e g r e g a t i o n a n u n l i k e l y m e c h a n i s m . Z i n c is, h o w e v e r , u n d e r -
sized i n A1. T h e a t o m i c v o l u m e o f t h e fl p h a s e is s m a l l e r t h a n t h a t o f t h e m a t r i x p h a s e , so
t h a t 6 < 0 a n d t h e e x c e s s v a c a n c y m e c h a n i s m ( s e c t i o n 4.1) w o u l d n o t give p r e c i p i t a t i o n .
PHASE STABILITY UNDER IRRADIATION 299

I
/,1.5 x 10"zdpa/s

LSxlO-3
dpn/s
200
/
/J'
Thermol •
Solvus/"
/
150 .."
/
/
/ Precipitotion:
I' / e Observed at low flux
I00 / / _ _ Q Nolobservedof low flux
/ DObservedat high flux
I DNot observed of high
I flux
J Ill
50 0
L I 2 3 4 5
Zn (o1%)

FIG. 44. Thermal and irradiation altered solve for AI-Zn alloys under 1 MV electron irradiation.
(After Cauvin and Martin 1981.)

The 1981 study was a major extension of the earlier work, and it also proposed a physical
basis for the results. Alloys of 0.2-4.4a/oZn were irradiated in the 1 MV HVEM at
displacement rates from 6 × 10 -4 to 1.5 x 10-2/s at temperatures from 100-250°C. the results
of the investigation are shown in Fig. 44, where no attempt is made to differentiate between
GP zones and fl-phase precipitates. The solvus under irradiation is again elevated well above
the thermal value with the change greatest at low solute concentrations and high displacement
rates.
Cauvin and Martin noted that Fig. 44 does not represent a true irradiation-altered phase
diagram in that the lever rule is not always obeyed. In the 4.4a/oZn alloy, the precipitate
volume fraction is only about 10~o of that predicted by the lever rule. The 0.8a/o alloy, on
the other hand, obeyed the lever rule at both high and low temperatures.
The authors attribute the irradiation-altered solvus to the model discussed in Section 4.2,
where recombination of vacancies and interstitials at the particle: matrix interface stabilizes
the precipitate phase.
Figure 45 compares theory and experiment. Reasonable agreement was found for the
2 x 10 -2 displacement rate in Fig. 45, and also for K = 2 x 10 -3 dps/s. Agreement between
calculated and measured solve is reasonable, especially in view of uncertainties in the material
parameters needed to evaluate the theory.
Theory was also found to predict violations of the lever rule. The calculated solvus was
relatively insensitive to changes in energies of vacancy formation and migration.

11. COPPER-BASED ALLOYS

The high thermal conductivity of some Cu-based alloys makes them candidates for fusion
reactor first wall materials. In addition, copper-based alloys show a variety of interesting
precipitation reactions which may be altered by irradiation. A number of such studies has
been made.
300 P R O G R E S S IN MATERIALS SCIENCE
T(K)
55C
Experimentol,, t

50C "~ ~-- ~-Calculoted

450

| ......... Thermol
40C I y." SoIvus
....'
350 - .,"'

•"1 I I I I
I 2 3 4 5
Zn (at %)
FZG. 45. Comparison of theoretical and experimental solvi for AI-Zn alloys irradiated at
2 × 10-2dpa/s. (After Cauvin and Martin, 1982.)

11.1. Solute Segregation


Takahashi et al. (1981) irradiated solution treated and quenched alloys of Cu plus 2a/oAg,
Fe, or Ni in the 650 kV HVEM. The displacement wate was 4-9 × 10 -4 dpa/s depending on
foil orientation with respect to the beam.
The Cu-Ag and Cu-Fe alloys were supersaturated at the irradiation temperature, whereas
Cu and Ni form a continuous range of solid solutions. Precipitation occurred during
irradiation of the supersaturated alloys, which would be expected to interfere with measure-
ment of solute segregation effects.
Irradiation of the Cu-Ag alloy to 10 dpa at 523 K gave solute enrichment peaks on either
side of grain boundaries and a local depletion right at the boundary. Takahashi et al.
interpreted this result as solute segregation from the boundary, which was expected for the
oversized Ag atoms.
Irradiation of the Cu-Fe alloy to 10 dpa at 573 K gave strong Fe segregation to void
surfaces.
After 10 dpa at temperatures of 523-673 K, Ni in the 2a/oNi alloy segregated to grain
boundaries and voids, in agreement with the size misfit theory of solute segregation. Figure
46 shows Ni segregation in a Cu-2a/oNi alloy after irradiation to 10 dpa at 573 K.
Knoll et al. (1981) irradiated Cu-Be, Cu-Co and Cu-Fe alloys with 14 MeV Cu 3÷ ions.
The displacement rate at the sample surface was between 1.7 × l0 -4 and 3 × 10-4dpa/s.
An alloy of 3.4 a/o Be was irradiated to up to 2.6 dpa at temperatures between 350 and
475°C. The highly undersized Be solute segregated to the irradiated surface as expected, as
seen in Fig. 47. The amount of segregation increased strongly with temperature and only

" Cu-2 at % Ni
b 32

~ 3o
o

~ 28
- .-," 0!It~ 1
I I I I I
-2 -I 0 I 2
Dislance from G.B. (/~m)
FIG. 46. Concentration of Ni as a function of distance from grain boundaries in Cu-2a/oNi after
650kV HVEM irradiation. (After Takahashi et al., 1981.)
PHASE STABILITY U N D E R I R R A D I A T I O N 301

o3~
.c~ * A.,~'~-.,o,,- • 430"C, 0.13dpo
~. o- u-..,.zl,.~ o 430"C, 0.06 dpa
~ ~ • Control,450°C
02 ,.
\ .__ _\ _
\ .o,
c,: "" ~•'-a.~o ~."
I I I
o( ,o zo 30 40
Deplh, nrn

FIG. 47. Beryllium concentration vs. depth profiles in Cu-3.4at.~Be irradiated at 430°C to 0.13 and
0.06 surface dpa with 14 MeV Cu 3÷ ions. Concentrations are first-order approximations generated
from AES data. (After Knoll et al., 1981.)

moderately with displacement level. This alloy also showed a substantial but lesser degree of
Be segregation to the surface, even in the absence of irradiation.
Irradiation of a Cu-la/oCo alloy to up to 1.2dpa at 400 or 475°C gave no measurable
segregation. Cobalt is undersized in Cu by ,,~4~o, so that some segregation was to be expected
on the basis of the size misfit criterion.
The Cu-la/oFe alloy was irradiated to a surface damage level of 0.83 dpa at 450, 400 and
350°C. No observable segregation occurred at 350 or 400°C, but Fe segregated from the
sample surface at 450°C. This behavior is opposite to that observed by Takahashi et al. (1981)
in an electron irradiated Cu-2a/oFe alloy.
Ohnuki et al. (1982) irradiated Cu-2a/oSi at 650-1000 kV in the HVEM at displacement
rates between 4 and 9 x 10 -4 dpa/s. Other alloys in the study were irradiated to 3-10 dpa,
so one would infer a similar damage level for the Cu-Si alloy. The oversized Si atoms were
segregated away from grain boundaries, as expected.
Knoll et al. (1981) note that although King (1966) gives Fe as slightly oversized in Cu, other
investigators have found the opposite. Further studies are clearly needed to determine the
segregation behavior of Fe in irradiated Cu-Fe alloys.
v. Lensa et al. (1977) and Bartels et al. (1979) studied solute segregation in alloys containing
17-182 appm Be. The alloys were irradiated with 3 MV e- between 280 and 410 K and the
damage state followed through electrical resistivity changes. Irradiations were to only
5-10 x 10-5 dpa but were suitable for the resistivity studies.
Beryllium was found to segregate strongly to internal point defect sinks. It was found that
one Be atom was removed from solid solution for every 2-4 Frenkel pairs created by
irradiation. As such, solute segregation was very effectively storing the energy of irradiation,
in this case as a non-uniform, hence non-equilibrium, solute concentrations. Beryllium atoms
are undersized in the Cu lattice, so the direction of segregation is in agreement with the atomic
misfit model.

11.2. C u - A u Alloys
Guinan et al. (1981) studied the disordering of highly ordered (S = 0.99) C u 3 A u at 4 K
during irradiation with 14.8 MeV neutrons. The degree of order was determined by resistivity
measurement.
JPM.S. 28!3-4 E
302 PROGRESS IN MATERIALS SCIENCE

In the absence of reordering the order parameter varies as:


dS
dt = - A S,

where ~b = neutron flux and A = disordering parameter. This equation is equivalent to Eq.
(8.4) with A~b replacing eK, the product of the disordering efficiency and the displacement
rate. The study found A = 15.7 × 10-16cm2/n as compared to ~ 4 × 10-16cm2/n for lower
energy fission neutrons. It was concluded that A scaled with the amount of displacement
energy transferred from the neutrons to the alloy.

11.3. Cu-Be Alloys


Copper-beryllium may be age-hardened through formation of coherent Be-rich GP zones
and transition precipitates. A number of neutron and electron irradiation experiments has
been performed on these alloys. Figure 48 shows the Cu-rich end of the Cu-Be phase diagram.
Murray and Taylor (1954) exposed Cu-13.5a/o(2.2w/o)Be to up to 3 x 1019 fast n/cm 2
( ~ 1.5 x 10 -2 dpa) at temperatures between 0 ° and 40°C. The fact neutron flux was about
4 x 10H/cmZ.s. Irradiation gave much the same changes in resistivity and density at does
thermal ageing at higher temperatures. From this result it was inferred that irradiation was
speeding zone formation. Results of subsequent annealing at higher temperatures were
interpreted as enhanced reversion of GP zones in the irradiated samples.
Murray and Taylor (1954) interpreted their results as precipitation induced by fast neutron
irradiation. They suggested precipitation was enhanced by an increased nucleation rate due
to the ~ 10 -3 atom fraction Frenkel pairs induced by irradiation. Richards (1955) noted that
Cu-Be alloys are notoriously erratic in their response to heat treatment and suggested that
the results of Murray and Taylor (1954) could be better interpreted in terms of an
irradiation-increased dislocation density, which effectively cold-works the material and speeds
the aging process.

WI. percentBeryllium
0.51 2 4 6 8 I0 14
N~

130~

120C

100C "
/
k ~ ~.,.. .

80C
<co,/
700 or a
\/
,8
_,
60(]

4O0
g÷a
30~/

2000 I0 20 30 40 50 60
Cu Atomic percent Beryllium Be

Flo. 48. Portion of Cu-Be phase diagram. (After Hanson, 1958.)


PHASE STABILITY UNDER IRRADIATION 303

Yoshida el aL (1971), Yoshida (1969) and Yoshida and Sagane (1972) made an intensive
study of the effects of neutron irradiation on aging in Cu-Be alloys.
In the studies of Yoshida et al. (1971) quenched alloys of 12.6a/oBe, some with minor
additions of Mg, Co, Zn, and Fe were irradiated to 1-4 x 1017 n / c m 2 of unstated energy at
~80°C. If these were mostly fast neutrons the maximum damage level was about
2 x l0 -4 dpa. Resistivity changes during irradiation and during post-irradiation annealing
indicated, as did the work of Richards (1958), that compared to thermal aging GP zones
formed faster during irradiation and 7' replaced the GP zones more rapidly on post-
irradiation annealing. In addition, in TEM studies of alloys irradiated to 3 x 1017n/cm 2 the
as-irradiated microstructure showed cross-hatching parallel to {110} planes, and the electron
diffraction patterns showed streaks through matrix spots running in the (100) directions.
both observations are characteristic of plate-like GP zones on {100} planes in Cu-Be alloys.
Prior to irradiation the alloys showed only very weak indications of GP zones.
The transition 7' phase formed more rapidly during post-irradiation annealing than during
aging of non-irradiated specimens, as occurred in the work of Murray and Taylor (1954).
Yoshida et al. (1971) attributed the accelerated GP zone formation to irradiation-enhanced
diffusion, and possibly to preferential displacement of the light Be atoms. They believed that
GP zones formed during irradiation were less stable than those formed during purely thermal
treatments, and that the 7' nuclei were formed during irradiation. The combination of
unstable GP zones and pre-existing 7' nuclei would then enhance 7' formation during
post-irradiation annealing.
Bystrov et al. (1968) studied the effect of 2.5 MV e- irradiation on alloys of Cu-2.5~Be.
The composition is presumably in w/o, since Cu-2.Sw/oBe is a common alloy. The alloy was
quenched into water from 830°C and irradiated at 20-25°C to 0.55-3.8 x 1016 e-/cm 2. The
maximum exposure would produce about 10 -5 dpa and no displacement cascades.
During post-irradiation annealing at 20 and 80°C accelerated resistivity increases in
irradiated samples were interpreted as due to accelerated decomposition of the solid solution.
At 150°C, however, the unirradiated samples showed the most rapid increase in resistivity.
Irradiation-accelerated decomposition was interpreted as interstitial-enhanced GP zone
formation during irradiation and vacancy-accelerated decomposition at 20 and 80°C.
Irradiation-retarded decomposition at 150°C was attributed to thermally-produced vacancies
migrating as complexes while irradiation-induced vacancies do not. This interpretation is
inconsistent with current knowledge of the structure of vacancies in metals.
Koch et al. (1981) bombarded pre-thinned (-~ 1000 ,~) Cu-l.35a/oBe samples with 300 keV
Cu + ions. Irradiation was at temperatures of 600-670 K, at rates of 5 × l0 -4 to 5 × l0 -3 dpa/s
to damage levels up to 8 dpa.
A high density (1015-1016/cm3) of equilibrium 7 (CsCl structure) precipitates formed in all
cases. The particles were roughly spherical and were distributed at random in the matrix.
Koch et aL attributed the precipitation to Be interstitial segregation to cascade-induced
vacancy loops, followed by nucleation in the Be-enriched region, but were unable to observe
such loops in the electron microscope. It is also possible that some other mechanism, such
as defect:solute co-precipitation or point defect recombination may have induced the
observed precipitation.
Tjhia el al. (1980) attributed vacancy loop formation in irradiated A1-Si alloys to a
diminution of the dislocation bias for interstitials due to an atmosphere of undersized Si
atoms. Beryllium is 26.45~o undersized in Cu, so a similar effect might be expected in Cu-Be
alloys.
Kinoshita et al. (1981) irradiated both under- and over-saturated Cu-Be alloys in a 650 KV
304 PROGRESS IN MATERIALS SCIENCE

HVEM at fluxes of 1-5 x 1018e-/cm 2.s and of 1-5 x 1019 e-/cm2.s for up to about 30 min.
Irradiation of the supersaturated 6 a/o Be alloy at 300, 350 and 430°C at the lower flux gave
the usual =--,GP zones ---,7 "--'7'---'7 precipitation sequence. The kinetics of precipitation were
much accelerated by irradiation.
The undersaturated 0.18w/oBe alloy was irradiated at both the low and high fluxes at
temperatures of 340--410°C. Precipitation of 7' and/or 7 occurred homogeneously in the
matrix, rather than at point defect sinks where the undersized Be atoms would tend to
segregate.
Kinoshita et al. (1981) proposed that homogeneous nucleation in the undersaturated
solution occurred by preferential migration of the Be part of the interstitial dumbbell and
subsequent clustering on {100} planes to form GP zones. This mechanism is something of
a qualitative restatement of the theory of Cauvin and Martin (1981a) for coherent homo-
geneous precipitation in undersaturated irradiated solid solutions discussed in Section 4.2.
Wahi and Wollenberger (1983) irradiated Cu-l.35a/oBe alloys in the 1 MV HVEM at
temperatures between 600 and 700 K at a displacement rate of about 10 -4 dpa/s. GP zones
formed at 700 K, where the alloy is undersaturated even with respect to the stable 7 phase.
At 600 and 640 K evidence was found ofT" or 7' particles, though unambiguous identification
was not possible. In all cases the displacement level was only about 0.1 dpa.

11.4. Cu-Co Alloys


Figure 49 shows the Cu-Co phase diagram. Precipitation in Cu-rich Cu-Co alloys is
usually in the form of a homogeneous dispersion of fine, coherent Co-rich particles. The
particles have a small misfit with the matrix and tend to lose coherency upon growing to a
certain size. The loss of coherency may be accelerated by cold work or irradiation.
Piercy (1962) exposed Cu-2w/oCo alloys to 3.4 x 1019 fission n/cm 2 at 50°C, or about
0.02 dpa. The alloy had been given pre-aging treatments at 395°C or 600°C to produce a
number density of 1017-1018/cm3 of coherent Co-rich particles of mean radii of 1-5 nm.
The experiment was designed to determine under what conditions displacement cascades
would dissolve the fine, Co-rich particles. Progress of the dissolution was detected through

WI, percent Copper


I0 ZO 30 40 50 60 70 80 90
1 5 0 0 ~
1400 ' \
1 3 0 0 ~ [ ~ ~ ~ ~
1 2 0 0 ~
I100 i

~ 1000 , (c=)

~ 800

600
500~
0 I0 20 30 40 50 60 70 80 90 I00
Co Atomic percent Copper Cu
FIG. 49. Copper-cobalt phase diagram. (After Hansen, 1958.)
PHASE STABILITY UNDER IRRADIATION 305

measurement of the saturation magnetism, which is sensitive to cobalt particle size and
number density.
The magnetic measurements were interpreted as showing that displacement cascades had
dissolved 10-20~ of the particles of less than i.2 nm radius, and that the cascades had a
radius of 3.4 nm. Alloys which had been aged to produce a relatively low number density
('-~ 5 X 1016/cm 3) of relatively large (r ~ 3-4 nm) particles showed an increase in particle
number density and a decrease in average size after irradiation. In one case particle number
density increased by 109~ while mean particle radius decreased from 4.7 nm to 3.7 nm. These
changes were attributed to nucleation of new particles during irradiation. Irradiation of a
quenched and un-aged alloy did not give measurable particle nucleation.
The idea of displacement cascades dissolving, or at least disrupting, particles of ~ 1 nm
radius is consistent with the current views on the structure of displacement cascades. Piercy
attributed the nucleation of new particles to diffusion which took place during cooling of the
displacment cascade. It is now known that atoms in a displacement cascade reach thermal
equilibrium after only a few lattice vibrations, and that very little diffusion would occur in
this extremely short period of time. It is, however, possible that irradiation-enhanced diffusion
facilitated the nucleation of new particles at the 50°C irradiation temperature. Approximately
half the Co was still in solid solution at the 600°C aging temperature and would be available
at lower temperatures for nucleation of new precipitates, as would the solute from particles
dissolved by displacement cascades.
L. M. Brown et al. (1968) and Woolhouse and Ipohorski (1971) studied the irradiation-
induced coherency loss of Co-rich particles in a Cu-rich matrix.
L. M. Brown et aL (1968) analyzed the energetics for relief of coherency strains in
undersized Co-rich precipitates by formation of a circumferential prismatic dislocation loop.
Figure 50 shows the results of their calculations; r0 is radius of the precipitate particle, and
rt is radius of the dislocation loop. At rc,t = 3.2 nm a loop around the precipitate will expand,
and at r* = 6 nm creation of a dislocation loop around the coherent particle gives a decrease
in energy.
An alloy of 2.65w/oCo was aged to produce particles with radii in the 3-17 nm range, then
irradiated to a dose of 2 x 1017fast n/cm 2 (~10-4 dpa), at ~65°C. The irradiation was

• • •

40

20

C~

-2C
i I I I I
20 40 60 80 I00

rt ,

FIG. 50. Energetics of coherency loss. With the particle radius, r 0 = 25 ,~, i.e. r 0 < r~t, the coherent
state is the only possible state. W h e n r 0 = 50,~,, i.e. r e d t < r 0 < r*, the incoherent state is possible
but metastable with respect to the coherent state. W h e n r0 = 75/~, i.e. r 0 > r*, the incoherent state
is stable with respect to the coherent state. (After Brown et al., 1968.)
306 PROGRESS IN M A T E R I A L S SCIENCE

expected to provide nuclei for coherency loss: an interstitial dislocation loop would be one
such nucleus.
The irradiated samples were annealed a few minutes at 400°C, then examined at room
temperature. All particles larger than 15 nm radius had lost coherency, whereas none of those
of less than 12 nm radius had lost coherency. According to calculations a matrix dislocation
could have caused coherency loss at 2.5 nm and the particle could have nucleated a misfit
dislocation at 6 nm. The measured radii are thus 2-4 times larger than the calculated values,
which is reasonable agreement in view of such simplifications in the model as considering only
one circumferential dislocation when three are needed for coherency loss.
Woolhouse and Ipohorski (1971) studied neutron and electron irradiation-induced coher-
ency loss in a Cu-lw/oCo alloy. The alloy was given aging treatments to produce particles
of radii of from 2.5-17.5 nm to up to 6.5-21 nm. Irradiations were of 1017-1.6 × 10~8 fast
n/cm 2 (~ 5 x 10 -5 to 10 -3 dpa) at 65-100°C and of 9.5 × 1022 450 kV e-/cm 2 in the HVEM.
The electron irradiation caused particles as small as 2.5 nm radius to lose coherency. This
observation is in good agreement with the calculated size needed to lose coherency with the
aid of a matrix dislocation. Under neutron irradiation and subsequent annealing at
400-600°C the critical size for coherency loss decreased from 19 nm at 2.3 × 1017 fast n/cm 2
to 8 nm at 4.4 x 1019 fast n/cm 2.
Coherency loss under neutron irradiation was attributed to climb and glide of an
irradiation-induced dislocation to the particle:matrix interface. In the case of electron
irradiation, coherency loss was attributed to migration of self interstitials to the interface
followed by prismatic dislocation loop nucleation. The much higher displacement rate and
damage level in the HVEM would favor such loop nucleation.

11.5. Cu-Fe A lloys


Boltax (1956) aged Cu-3.25w/oFe alloys for various times at 400, 550 and 700°C and
exposed then to 5 × 1019 n/cm 2 (E > 1 MeV), or about 0.02 dpa. The irradiation temperature
was presumably less than 100°C. The electrical resistivity was found to decrease by ,,, 5~ in
samples pre-aged at 700°C and to increase by up to 35~o in samples pre-aged at lower
temperatures. The greatest increase was in samples pre-aged the shortest times. Irradiation
increased the saturation magnetization by up to 150~, with the greatest increase coming for
shorter irradiation times.
Boltax (1965) exposed alloys of Cu plus 0.8-5w/oFe to up to 1.7 × 102°n/cm2
(E > 0.25 MeV), or about 0.1 dpa. Again, samples pre-aged at 700°C showed a resistivity
increase after irradiation, while those aged at lower temperatures ultimately showed an
increase, though sometimes after a small initial decrease. Irradiation also gave an increase
in saturation induction whether the sample had been quenched, quenched and aged, or slowly
cooled. The resistivity increases were attributed to irradiation resolution from precipitates and
decreases to irradiation-enhanced precipitation. The increases in saturation induction were
attributed to a small amount of irradiation-induced coherency loss which produced the
ferromagnetic incoherent phase.
Woolhouse and Ipohorski (1971) included Cu-lw/oFe alloys in their study of irradiation-
induced coherency loss. Electron irradiation to ~9.5 x 10:2 450kVe-/cm 2 gave at least
partial coherency loss down to a particle radius of 7 nm. Neutron irradiation to 4.4 × 1019
fast n/cm e ( ~ 0.02 dpa) followed by annealing at 400°C gave loss of coherency in 98~ of the
particles with a size range of 6-18 nm. The critical size for coherency loss by either type of
irradiation is thus about 7 nm.
PHASE S T A B I L I T Y U N D E R IRRADIATION 307

I I l I wi,h ,

1.561~'WiJh~"~
•~ 1.52

1.48
102 103 104 105 106
[rrodiotJon Time, minutes

FIG. 51. Resistivity changes in Cu-55w/oNi during a 144.5°C anneal, with and without irradiation.
Irradiation is seen to give a resistivity decrease. (After Schiile et al. 1975.)

The linear misfit for Fe particles in Cu is only 0.008 as compared with 0.013 for Co
particles, so that refit= 5.8 nm and r* = 13.1 nm (see Fig. 50). The results of both electron
and neutron irradiation are thus in good agreement with theory. The observation of
coherency loss in Cu-Fe alloys provides evidence to support the conclusion of Boltax (1956)
that irradiation-induced coherency loss was responsible for increases in the saturation
magnetism in his Cu-Fe alloys.

11.6. Cu-Ni Alloys*


Poerschke and Wollenberger (1975, 1976) studied the effects of electron irradiation on the
resistivity of alloys containing 42-71a/oNi. In the 1975 study alloys were irradiated to up to
3 x 10~8 3 M V e - / c m 2 ( ~ 3 x 10-4dpa) at 25, 82, and 95K, then given post-irradiation
anneals at up to 350 K. Resistivity decreases found on annealing at 125 K and above were
attributed to decomposition of the solid solution by Ni clustering by means of irradiation-
induced interstitialcy diffusion.
The 1976 study also used a 3 MV Van de Graaff generator as an electron source, but the
irradiations were performed at temperatures between 120 and 570 K. Below 420 K resistivity
decreases were attributed to solid solution decomposition by interstitialcy diffusion. At higher
temperatures vacancies also contributed to diffusion.
Schiile et al. (1975) irradiated alloys with 35-85w/oNi to up to 3 x 10~8 fast n/cm 2
(~ 1.5 x 10-3dpa) at temperatures between 100 and 150°C. The fast neutron flux was
1.7 x 101t/cm2.s. Figure 51 shows that resistivity decreases occurred after fast neutron
fluences as low as 3 x 1015/cm2. As in the electron irradiation studies, the resistivity decreases
were attributed to phase decomposition due to diffusion of irradiation-induced interstitials.
Copper-nickel alloys are known to have a tendency to cluster at low temperatures where
thermal diffusion has become very sluggish. The amount of diffusion caused by irradiation
to even low displacement levels might well be enough to cause a fine scale decomposition of
the solid solution. As such, the interstitialcy diffusion model advanced in all three studies
seems reasonable. The vacancy-interstitial recombination triggered precipitation process
discussed in Section 4.2 would be a probable mechanism except for the results of the
post-irradiation annealing studies of Poerschke and Wollenberger (1975). Resistivity de-
creases were observed at temperatures low for vacancy migration, so that vacancy: interstitial
recombination at the precipitate:matrix interface is improbable.

*See also Ni-Cu alloys.


308 PROGRESS IN MATERIALS SCIENCE

T a b l e 3. Solute segregation in copper-based alloys


Alloy Treatment Result References
Cu-2a/oAg 523K, 650kV e- in HVEM, Ag segregated away Takahashi et al. (1981)
4--9 x 10 -4 dpa/s, 10 dpa from void surfaces
Cu-2a/oFe 523 K, 650 kV e in HVEM, Fe segregated toward Takahashi et al. (1981)
4--9 x 10 -4 dpa/s, 10 dpa void surfaces
Cu-2a/oNi 523-673 K, 650 kV e- in Ni segregated to defect Takahashi et al. (1981)
HVEM, 4-9 x 10 -4 dpa/s, sinks
10 dpa
Cu-3.4a/oBe 350-475°C, 14MeV Cu ÷, Be segregated to irra- Knoll et al. (1981)
1.7-3 x 10-adpa/s, diated surface
up to 2.6 dpa
Cu-la/oCo 400, 475°C, 14 MeV Cu ÷, No observable Co segre- Knoll et al. (1981)
1.7-3 x l0 -4 dpa/s, gation
up to 1.2 dpa
Cu-la/oFe 350, 400, 450°C, 14 MeV Fe segregated from irra- Knoll et al. (1981)
Cu ÷, 1.7-3 x 10 -4 diated surface at 450°C,
dpa/s, 0.83 dpa no observable segre-
gation at 350 or 400°C
Cu + 17-182at.ppm 280--410K, 3 MV e , Segregation of 1 Be v. Lensa et al. (1977);
Be 5-10 x 10 -5 dpa atom to defect sinks per Bartels et al. (1979)
2-4 Frenkel pairs
Cu-2a/oSi 420-670 K, 650-1000 kV Si segregated away from Ohnuki et al. (1982)
HVEM, a few (?) dpa grain boundaries

T a b l e 4. Phase stability in copper-based alloys.


Alloy Treatment Result References
Cu3Au 4 K, up to 1.9 x 1017 Disordering linear with Guinan et al. (1981)
14.8 MeV n/cm2 ftuence, scales with en-
ergy
Cu-2.2w/oBe, 0-40°C, up to 3 x 1019n/cm 2 Accelerated GP zone Murray and Taylor (1954)
Cu-l.8w/oBe (fast), ~ 1.5 x 10-2 dpa formation
0.4 x 1012n/cm:/s (fast)
Cu-12.6a/oBe + 80°C, 1-4 x 1017n/cm2 Accelerated formation Yoshida et al. (1971)
impurities (E > 1 MeV) ~ 2 x 10 -4 of GP zones and 7'
dpa max, ~ 10 -8 dpa/s
Cu-2.5w/oBe 20-25°C, 0.55-3.8 x 10'6 Accelerated decomposi- Bystrov et al. (1968)
2.5 MV e-/cm 2, ~ 10 -5 tion of solid solution on
dpa subsequent anneal
Cu-l.35a/oBe 600-670 K, 300 keV Cu ÷, Copious precipitation of Koch et al. (1981)
5 x 10 -4 to 5 x 10 3 equilibrium ~/phase
dpa/s, 1-8 dpa
Cu-0.91w/oBe 300-430°C, 1-5 × l0 is Accelerated precip- Kinoshita et al. (1981)
650kV e-/cm2-s, ~20rain. in itation sequence
HVEM
Cu-0.18w/oBe 340-410°C, 1-5 x 10~s Homogeneous precip- Kinoshita et al. (1981)
or 1-5 x 1019 650 kV itation of 7' and/or y in
e-/cm2.s, ~20min. in undersaturated solid
HVEM solution
Cu-l.35a/oBe 600-700 K, 1 MV HVEM, GP zones at 700 K, ~" or Wahi and Wollenberger
~ 0.1 dpa, ~ 10 -4 dpa/s y' at lower temperatures (1983)
Cu-2w/oCo 50°C, 3.4 x 1019 Dissolution of < 1.2 nm Piercy (1962)
fission n/cm 2 (~0.02dpa) radius particles, nucle-
ation of high density
of fine particles
Cu-2.65w/oCo ~65°C, 2 x 1017 fast Coherency loss by all Brown et al. (1968)
n/cm2 ( ~ 10-4 dpa) particles larger than
15 nm radius
Cu-lw/oCo 65-100°C, 2 x 1017 fast Coherency loss for Woolhouse and Ipohorski
n/cm(~ 10 4dpa) r > 19nm (1971)
65-100°C, 1.6 x 1018 fast Coherency loss for
n/cm 2 ( ~ 10 -3 dpa) r>8nm
T < 100°C?, 9.5 x 1022 Coherency loss for
450 KV e-/cm 2 r >2.5nm
PHASE STABILITY UNDER IRRADIATION 309

Table 4. Continued.
Alloy Treatment Result References
Cu-3.25w/oFe T < 100°C?, 5 x 1019n/cm2 Increase in saturation Boltax (1956)
(E > 1 MeV) ~0.02dpa magnetization; re-
sistivity decrease in
samples pre-aged at
700°C, increase in
others
Cu-0.8-5w/oFe T < 100°C?, up to 1.7 x Resistivity increase for Boltax (1965)
1020n/cm 2 (E > 0.25 MeV) pre-aging T = 700°C,
~0.1 dpa decrease for lower T.
Saturation mag-
netization increase in
many cases
Cu-lw/oFe T = ?, 9.5 x 10 22 450 kV Coherency loss for Woolhouse and Ipohorski
e - / c m 2 o r 4.4 × 1019 particles with r > 7 nm (1971)
fast n/cm2 (~0.02dpa)
Cu-42-71 a/oNi T<100K, up to 3x l0 ts Resistivity decrease on Poerschke and Wol-
3MV e- (up to ~ 3 x 10-4 annealing at T > 120K lenberger (1974)
dpa)
Cu-58.7a/oNi 120o570 K, up to 6 x 10TM Resistivity decrease Poerschke and Wol-
3 MVe-/cm 2 (up to during irradiation lenberger (1975)
~6 x 10-4 dpa)
Cu-35-85w/oNi 100-150°C, up to Resistivity decrease Schiile et al. (1975)
3 × l0 ts fast n/cm2 during irradiation
~ 1.5 × 10 -3 dpa,
1.7 × I0 ~ fast n/cm2.s,
~ 10-10dpa/s

12. IRON-BASED ALLOYS, FERRITIC*


T w o classes o f ferritic alloys are o f p a r t i c u l a r interest in studies o f i r r a d i a t i o n response.
A l l o y s o f i r o n plus small a m o u n t s o f c a r b o n o r n i t r o g e n have been studied to u n d e r s t a n d
the i n t e r a c t i o n o f interstitial solutes with i r r a d i a t i o n - i n d u c e d defects, in p a r t i c u l a r vacancies,
self interstitials, a n d d i s p l a c e m e n t cascades. O t h e r studies o f s u b s t i t u t i o n a l alloys were
designed to help u n d e r s t a n d p h a s e stability u n d e r i r r a d i a t i o n in ferritic stainless steels, which
are p o t e n t i a l s t r u c t u r a l m a t e r i a l s for c o n t r o l l e d fusion r e a c t o r first walls.

12.1. F e - C A l l o y s
M o s t i r r a d i a t i o n s o f F e - C alloys have been p e r f o r m e d to low fluences in low-flux research
reactors at t e m p e r a t u r e s between - 100°C a n d + 100°C. I r o n interstitials n o t b o u n d to t r a p s
are m o b i l e to well b e l o w - 100°C, a n d the C a t o m b e c o m e s m o b i l e in the - 100 to 100°C
t e m p e r a t u r e range. The t e m p e r a t u r e at which the F e v a c a n c y b e c o m e s m o b i l e is a m a t t e r o f
c o n t r o v e r s y . A l t h o u g h m o s t w o r k e r s a t t r i b u t e a 220 K a n n e a l i n g p e a k to v a c a n c y m i g r a t i o n ,
such r a p i d low t e m p e r a t u r e m i g r a t i o n is inconsistent with high t e m p e r a t u r e data. S o m e
e x p e r i m e n t s have also been d o n e in the H V E M , which p r o d u c e s vacancies a n d self
interstitials, b u t n o t d i s p l a c e m e n t cascades. This absence o f cascades is o f assistance in
d i s t i n g u i s h i n g b e t w e e n v a r i o u s p r o p o s e d m e c h a n i s m s for the effects o f i r r a d i a t i o n on
precipitation.
H u l l a n d M o g f o r d (1961) p e r f o r m e d one o f the early studies o n the effect o f n e u t r o n
i r r a d i a t i o n o n p r e c i p i t a t i o n in F e - C alloys. S a m p l e s o f s o l u t i o n a n n e a l e d F e - 0 . 0 0 4 w / o C were
i r r a d i a t e d to 2.7 x 1015-7.8 × 1016 n / c m 2 ( E > 1 M e V ) o r a p p r o x i m a t e l y 0 . 1 - 4 × 10 -5 d p a at
t e m p e r a t u r e s between 60 a n d 100°C. P o s t - i r r a d i a t i o n a n n e a l i n g at t e m p e r a t u r e s o f 100-250°C
*See also notes in proof (pp. 430~,34).
310 PROGRESS IN MATERIALS SCIENCE

preduced copious carbide precipitation which the authors concluded occurred on displace-
ment cascades.
An extended study of irradiated Fe-C alloys was performed at Brookhaven National
Laboratory in the period 1960-1970 by Arndt and Damask (1964), Johnson and Damask
(1964), Wagenblast and Damask (1962), Wagenblast et al. 0964) and Damask et al. (1970)
in order to determine the mechanism of irradiation-induced precipitation. This group
concluded, as did Hull and Mogford (1961), that the carbides were nucleating at displacement
cascades, and attributed the decrease in resistivity or internal friction during 57°C annealing
to trapping of carbon atoms at vacancies.
In the final paper of the series, Damask et al. (1970) presented a comprehensive scenario
for the effects of neutron irradiation on precipitation in Fe-C alloys. After short-time
irradiations at 30°C, the number of precipitates exceeded the number of displacement
cascades so that another nucleation site had to be operative. This site was concluded to be
a Fe-interstitial bound to two C atoms. A C-Fe interstitial binding energy of 0.5 eV was
calculated by Johnson and Damask (1964) and binding to the second carbon atom would
be expected. Precipitates which form after more extended irradiation were taken to be
nucleated on displacement cascades.
Somewhat surprisingly, irradiation at -78°C followed by a 100°C anneal produced some
20 x fewer precipitates than did irradiation at 30°C. This discrepancy was explained on the
following basis. During the 30°C irradiation, carbon atoms have some mobility, and can
associate to form di-carbons and then the di-carbon-interstitial nucleus. Carbon is immobile,
however, at -78°C, and only C-I complexes may form. Then, on warming the sample to
room temperature, these complexes dissociate and the interstitials diffuse to sinks before
di-carbons have time to form.
Takeyama and Takahashi (1974, 1975a) irradiated a Fe-0.025w/oC alloy at room
temperature to a fluence of 1.7 x l 0 21 e-/cm 2 in a 650 kV HVEM, and then aged the alloy
at temperatures of 130-250°C. No precipitates were visible after irradiation. After aging
20 min at 200 and 234°C, e-carbide precipitation was observed in the non-irradiated region
but, suprisingly, not in the adjacent irradiated area. Figure 52 shows the results of irradiation
followed by a 20 min anneal at 200°C. The authors concluded that the lack of precipitation
in the irradiated area was due to trapping of carbon atoms at vacancies. They acknowledged
that carbon binding to cementite was stronger than to vacancies but asserted that the C-V
complexes would not dissociate at temperatures below 240°C. After aging at 246°C for 1 hr,
cementite formed in the irradiated area, in agreement with their assertion of strong C-V
binding.
Apparently, Takeyama and Takahashi (1974, 1975a) did not consider an explanation based
on carbon-interstitial binding because, due to the low activation energy of interstitial
migration, the complexes would dissociate very rapidly at their aging temperatures. The C-V
complex has a binding energy measured at 0.4 eV (Arndt and Damask, 1964) as compared
to 0.5 eV calculated by Johnson and Damask (1964) for the C-I complex. However, since
the activation energy for dissociation equals the binding energy plus the energy for defect
migration, the C-V complexes will be much more stable than will C-I complexes.

12.2. F e - C r Alloys
Ferritic and martensitic stainless steels have only recently received serious consideration for
use in high-flux irradiation environments. As such, relatively few studies of irradiation
response have been conducted on these alloys, and in most cases were focused on their high
resistance to void swelling rather than on precipitation behavior.
PHASE S T A B I L I T Y U N D E R I R R A D I A T I O N 311

Takahashi et al. (1981) studied solute segregation in Fe-5 and 13a/oCr alloys. Irradiation
was in the 650 kV HVEM at 3 x 1023e-/m 2. s, or a displacement rate of about 5 x 10 -5 dpa/s.
Irradiation of the alloys to 3 dpa at 673 K produced a ~ 200 nm wide Cr-depleted region
each side of the grain boundary, with the larger depletion in the 13~ alloy. Although
Takahashi et al. state that Cr is undersized by 0.63~o in ferrite, King (1966) gives D~f= 4.36~.
If the King value of f~sfis accepted, segregation away from the grain boundary is in agreement
with the size misfit theory of irradiation-induced solute segregation.
Ohnuki et al. (1981) irradiated with 200 keV C + ions up to 118 dpa at 8 x 10 -3 dpa/s at
798 K to induce segregation in Fe-13w/oCr and in Fe-13w/oCr doped with lw/o Si or Ti.
Chromium was enriched at grain boundaries and voids in the Fe-Cr and Fe-Cr-Si alloys.
This apparent segregation was due to formation of Cr-rich carbides. However, Cr was
segregated away from these sinks in the Fe-Cr-Ti alloy. The addition of 1~o Ti has thus
reversed the direction of Cr segregation; Ti was found to segregate toward grain boundaries.
Titanium is 14.4~o oversized in Fe, and on this basis should be depleted at the grain
boundaries.
Ohnuki et al. invoked C-Cr and C-Ti binding to explain this reversal in direction of Cr
segregation. A simpler mechanism would be the heterogeneous nucleation of chromium or
titanium carbides on the grain boundaries. These carbides are so stable that they would
nucleate and grow even if solute were segregated away from the boundary.
Brimhall et al. (1981) studied solute segregation in HT-9, a Fe based alloy with 11.3w/oCr,
0.5w/oMo and 0.38w/oSi, as the main alloying additions. Irradiation with 5 MeV Ni 2+ ions
at 3 × 10 -3 dpa/s gave 2 dpa at the sample surface. Auger electron spectroscopy showed
significant surface segregation of N at 775 K and a lesser degree at 875 K. A lack of P
segregation was attributed to its having combined with Mo in the grain interior. These results
contrast with those from the FCC alloys, PE-16 and type 316 stainless steel, which showed
strong surface segregation of P.
A recent extensive study of precipitation in irradiated ferritic materials was made by Gelles
(1981a) who studied several commercial alloys which cover the compositions 2.25-17w/oCr,
0.04-2w/oMo, 0.04-0.4w/oC. The alloys were irradiated in EBR-II to a maximum fluence of
1.76 × 1023n/cm2 ( E > 0 . 1 MeV) or 84dpa, at temperatures of 400-650°C. The starting
microstructures were various combinations of ferrite and tempered and untempered mar-
tensite.
The dislocation structure in some of these microstructures was so dense that the precipitates
were often hard to identify or even to observe, so that the precipitation behavior could not
be determined quantitatively. Precipitate phases observed varied with composition, tem-
perature, and sometimes from grain to grain. However, M 6 C , M o 2 C , Z, Laves, M23C6, 0t', and
a new unidentified low temperature phase were observed. Figure 53 shows the irradiation
response of alloy H-11. The globular precipitates along the martensite lath boundaries are
MrC, and the numerous rod shaped particles in the laths are Mo2C. (Section 13 gives a
discussion of these phases as they appear in austenitic alloys.)
Little and Stoter (1982) recently published the results of irradiation studies on three
martensitic stainless steels which contained approximately 12w/oCr, 0.5w/oNi, and
0.046-0.19w/oC. One steel also contained 0.96w/oMo and 0.30w/oV, while another contained
0.6w/oMo, 0.14w/oV and 0.26w/oNb. Irradiation was at 380-615°C in the Dounreay fast
reactor to an average displacement level of 30 dpa. Pre-irradiation microstructures were
tempered martensite, except for the lowest carbon steel which also contained some 6-ferrite.
All steels contained M23X6 (X = C, Si) as the primary precipitate phase.
Irradiation induced primarily the M 6 X (.¥ = C, Si), Z, or, and ct' (Cr-rich ferrite) and
312 P R O G R E S S IN MATERIALS SCIENCE

phosphide phases, although some Laves phase also appeared. The appearance of the various
phases depended somewhat on composition and temperature. The results of Little and Stoter
are in general agreement with those of Gelles (1981a).
More studies will be needed to classify the phases in Fe-Cr alloys as irradiation-induced,
irradiation-altered, etc., as has been done in the case of austenitic steels.

12.3. F e - M n Alloys
Takahashi et al. (1981) irradiated a Fe-la/oMn alloy to 3 dpa at 673 K in a 650 kV HVEM.
Manganese, which is oversized in 0t-Fe, was segregated away from grain boundaries as
predicted by the size misfit model.

12.4. F e - N Alloys
Like carbon, nitrogen dissolves sparingly in BCC iron to form an interstitial solid solution.
The migration, binding, and precipitation characteristics of the two solutes are similar unde;
thermal conditions, so that one might expect similar effect of irradiation on precipitation in
Fe-C and Fe-N alloys. These effects in Fe--C alloys were described earlier as being primarily
trapping of solute by irradiation-induced point defects and precipitate nucleation on
displacement cascades.
Stanley (1964) subjected a Fe-0.015w/oN alloy to neutron irradiation at T < - 120°C. He
irradiated to a maximum of ,--4 x 10 -5 dpa, so as to produce displacement cascades for
nucleation sites, while keeping the point defect concentration too low for appreciable solute
trapping to occur.
Irradiation was found to give a large increase in the rate of nitride precipitation during
post-irradiation anneals at 65°C. The effect saturated after 3 x 1015 n/cm 2 (,-~ 10 -6 dpa). By
contrast, irradiation to the same maximum damage level at 0°C with 2 MV e- had no effect
on precipitation kinetics on post-irradiation annealing, supporting the contention that
precipitation enhancement due to neutron irradiation was from displacement cascades.
Takeyama and Takahashi (1975a, 1975b) and Yamane et al. (1979) have studied the effects
of 650 kV electron irradiation in the HVEM on precipitation in Fe-N alloys.
Takeyama and Takahashi irradiated a Fe-0.03w/oN alloy to 1.8 x 1021e-/cm 2 at room
temperature in the 650 kV HVEM. A post-irradiation anneal of 130°C for 30min gave
copious Fe~6N2 nitride precipitation in the non-irradiated but not in the irradiated region.
Only after heating to T > 200°C and re-aging at a lower temperature did nitride precipitation
occur in the irradiated region. It was concluded that the lack of precipitation in the irradiated
region during the 130°C anneal was due to solute capture as V-N complexes, which
decomposed only on heating to T > 200°C.
This explanation is plausible; electron irradiation creates only vacancies and self inter-
stitials, and an I-N complex could hardly be stable to a temperature of 200°C.
Yamane et al. (1979) irradiated a Fe-0.013w/oN alloy with 2.5 x 10212 MV e-/cm 2 at room
temperature in the HVEM. As in the study just described, 160 and 190°C anneals gave nitride
precipitation in the non-irradiated but not in the irradiated region. Like earlier workers, they
attributed this lack of precipitation to V-N trapping.

12.5. F e - N i - V Alloys
Braski (1981) studied the effects of neutron and heavy ion irradiation on ordered alloys
of the (Fe, Ni)3V type. These alloys have a critical ordering temperature of ~670°C.
PHASE S T A B I L I T Y UNDER IRRADIATION 313

FiG. 52. Response of a Fc-0.025w/oC alloy to irradiation and to thermal treatment. (a) 1.7 x 1021
650kV e-/crn 2 at room temperature followed by 2 0 m in anneal at 200°C. Irradiated region is
precipitate free; (b) non-irradiated region near the irradiated region. (Electron micrographs courtesy
T. Takeyama and H. Takahashi.)
0

trJ
rj~
tb~

-t

N
z
t"l'l

FIG. 53. (a-b)


"0

--t
,.¢

-]

FIG. 53. Microstructural response of the martensitic stainless steel H 11 to neutron irradiation at 427°C. (a) control sample; (b) 5 x 1022n/cm 2 (E > 0.1 MeV);
(c) 1.5 x 1023n/cm 2 (E > 0.1 MeV); (d) 1.5 x 1023n/cm 2 (E > 0.1 MeV). The globular particles along the lath boundaries are M6C, and the numerous rod shaped
particles in the laths are Mo2C. (Electron micrographs courtesy D. S. GeUes.)
316 PROGRESS IN M A T E R I A L S SCIENCE

~) ~0
I I I I 1 1 I I
F e - 2 0 g t - 1 2 Ni-I Mo - I S i FS-2Q ~ - 1 2 N i - I I ~ - I S i
430"C;6 dm 51i0"¢; 6 dim

40 4O

Q
m | 3,0 B

i 211

10 ^

"- 9 n
: : :, :
m
:
5 I0 15 20 25 LC . . . . " " - " i 25
I~PTH l l i l O(PTH (~1

I 1 I I
RI-20 Cr- 12Ni-I Ide-ISi Ire-20 ¢r - 12Ni - I Ilo -I Si
100°C~6 aim 6?5*¢ i 6 i ; o

40

zm
-

! °
2C o Cf

Ni
I0 - -

Si

. I~ ==7 . . . .
"0 - ~ ~0 I~ 20 25
I~PTH I ~ | I~PTX (M~I

F[G. 54. Composition vs. depth profiles of the solute elements in Ni ÷ ion irradiated
F e - 2 0 C r - 1 2 N i - l M o - l S i alloy. (Courtesy P. R. Okamato and Journal of Nuclear Materials.)
PHASE S T A B I L I T Y U N D E R IRRADIATION 317

T a b l e 5. Phase stability in ferritic alloys


Alloy
(composition in w/o
unless otherwise noted) Treatment Result References
Fe--0.004C 60-100°C, 2.7 x 10Is- Carbide precipitation Hull and Mogford
7.8 x 1016n/em 2 (E > on post-irradiation an- (1961)
1 MeV) ~0.1-4 x 10-5 dpa nealing at 100-250°C.
Fe-0.01C 200K, up to 2.6 x Accelerated decay of Wagenblast and
10~7n/cm 2 (fast), internal friction peak Damask (1962)
~ 10 -4 dpa on post-irradiation
annealing at ~ 50°C.
Fe-0.011C - 100°C, 4.5 x 1017 Resistivity drop at Wagenblast et al.
n/cm 2 (fast), ~ 2 x 50°C, suppress 160°C (1964)
10- 4 dpa metastable carbide
precipitation.
Fe-0.0075, 0.01, and 30°C, -78°C, 1.2 x 10t4- Copious carbides after Damask et al. (1970)
0.015C 1 × 1017n/cm 2 (fast) 30°C irradiation,
~ 5 x 10-8-5 x 10 -5 fewer after - 7 8 ° C ir-
dpa radiation plus 100°C
anneal.
Fe--0.025C 25°C, 1.7 x 102t, 650 Precipitation in non- Takeyama and
kVe-/cm 2 in HVEM irradiated but not in Takahashi (1974,
irradiated region on 1975a)
anneal at T < 240°C.
Fe-5, 13a/oCr 673 K, 650keV e- in Cr depleted at sinks. Takahashi et al.
HVEM, 3 dpa, 4-9 x 10 -4 dpa/s (1981)
Fe-13Cr 798 K, 200 keV C ÷, 8 x Cr precipitates at Ohnuki et al. (1981)
Fe-13Cr. 1Si 10-3 dpa/s, up to 118 dpa sinks.
Fe-13Cr-lTi 798 K, 200 keV C 4, 8 x Cr segregated from Ohnuki et al. (1981)
10 -3 dpa/s, up to 118 dpa sinks, Ti toward sinks.
HT-9 (11.3Cr, 775, 875 K, 5 MeV Ni 2+, N segregation to free Brimhall et al.
0.52Mo, 0.38Si) 2 dpa, 3 x 10 -3 dpa/s surface. (1981)
Fe + 2.25-17Cr, 400--650°C, EBR-II up to Observed M6C, Mo2C, Gelles (1981)
0.04-2Mo, 1.76 x 1023 n/cm 2 (E > X, Laves, M23C6, Ct',
0.04-0.4C 0.1 MeV) 84 dpa and unidentified phase
in various combina-
tions.
Fe-0.046C, 13.3Cr, 380-615°C, DFR, 30dpa M6X, X, tr, ~', Laves Little and Stoter
0.38Ni; and phosphide phases (1982)
Fe-0.19C, 11.7Cr, form under irra-
0.62Ni, 0.96Mo, 0.30V; diation.
Fe--0.10C, 10.7Cr, 0.65Ni,
0.60Mo, 0.14V, 0.26Nb
Fe-la/oMn 673 K, 650 kV e- in Mn segregated from Takahashi et al.
HVEM, 3dpa, 4--9 × 10-4dpa/s grain boundaries. (1981)
Fe-0.015N T < -120°C, up to 7 x 1016 Enhanced nitride pre- Stanley (1964)
n/cm 2 (fast), up to ~ 4 × 10-5 cipitation during 65°C
dpa anneal
Fe-0.03N R.T., 1.8 x 1021, 650 kV Nitrides in un- Takeyama and Tak-
e-/cm 2 in HVEM irradiated but not in ahashi (1975a,
irradiated areas dur- 1975b)
ing 130°C anneal
Fe-0.013N R.T., 2.5 x 1021, 2 MV 160 or 190°C anneal Yamane et al. (1979)
e-/cm 2 in HVEM gives nitrides in un-
irradiated but not in
irradiated region.
(Fe, Ni)3V 250, 350, 550°C, ORR, Does not disorder be- Braski (1981)
3.8 dpa low Tc.
(Fe, Ni)3V 525, 570, 625, 680°C, Disorders only above Braski (1981)
4MeV Ni ÷, 70dpa L.
318 P R O G R E S S IN MATERIALS SCIENCE

Table 6. Compositions of some commercial stainless


steels*
Steel Composition
304 18-20 Cr, 8-10.50 Ni
0.08 C, 2.0 Mn, 1.0 Si,
0.045 P, 0.030 S
304L 18-20 Cr, 8-12 Ni,
0.03 C, 2.0 Mn, 1.0 Si,
0.045 P, 0.030 S
316 16-18 Cr, 10-14 Ni
0.08 C, 2.0 Mn, 1.0 Si,
0.045 P, 0.030 S, 2.0-3.0 Mo
316L 16-18 Cr, 10-14 Ni,
0.03 C, 2.0 Mn, 1.0 Si,
0.045 P, 0.030 S, 2.0-3.0 Mo
321 17-19 Cr, 9-12 Ni,
0.08 C, 2.0 Mn, 1.0 Si,
0.045 P, 0.030 S (Ti, 5 x C min)
347 or 17-19 Cr, 9-13 Ni,
En58G 0.08 C, 2.0 Mn, 1.0 Si,
0.045 P, 0.030 S (Cb + Ta,
10 x C min)
20/25/Nbt 20.1 Cr, 25.5 Ni,
0.1 C, 0.65 Mn, 0.68Si,
0.01 P, 0.02 S, 0.6 Nb
FV548:~ 16.5 Cr, 11.8 Ni, 0.11 C, 1.14 Mn,
0.35 Si, 0.009 P, 0.006 S, 0.92 Nb,
1.44 Mo, 0.024 N, 0.0025 B, bal Fe.
*From Metals Progress Materials and Processing Data-
book '82, American Society for Metals, Metals Park, Ohio
(1982), except:
?From Fisher and Williams (1972).
:~From Williams (1982).

Irradiation to 3.8 dpa at 250, 350 or 550°C in the Oak Ridge reactor (ORR) left the alloy
in an ordered state. Irradiation with 4MeV Ni ÷ ions to 70dpa at 525, 570, and 625°C
similarly left the alloy in an ordered state. Ion irradiation to 70 dpa just above the 670°C
critical ordering temperature disordered the alloy. The results were not compared with the
theory of irradiation-induced disordering (Section 8); such a comparison would be useful.

Table 7. Solute segregation in austenitic alloys


Alloys
(Compositions in w/o
unless otherwise stated) Treatment Result References
Fe-18Cr-8Ni-lSi 600°C, 3.25 MeV Ni ÷, ~ 15 Segregation of Ni and Si to Okamoto and Wieder-
dpa, ~ 10-3 dpa/s at surface surface, Cr away from sur- sich (1974)
face.
Fe-18Cr-8Ni-I Si 500°C, 1 MV e , 0.3 dpa, Strains around voids, attrib- Okamoto and Wieder-
6 x 10-4dpa/s uted to Cr depletion. sich (1974)
Fe-15Cr-20Ni-0.75Si, 675°C, 2.5 x 1016Ni Ni segregated to free sur- Johnston et al. (1977)
Fe-15Cr-20Ni ions/cm2, ~ 30 dpa at face, Cr segregated away
surface
Type 316 SS 775 & 875 K, 5 MeV Ni 2+, Strong segregation of P and Brimhall et al. (1981)
2 dpa, 3 x 10 3dpa/s S to surface at 775 K, of Si
at surface at 775 K, of Ni at 875 K.
Fe-20Cr-12Ni + Mo, Si 430-675°C, 3 MeV Ni ÷, Si, Ni segregate to surface; Sethi and Okamoto
2dpa, 6 × 10-4 dpa/s Cr, Mo away from surface. (1981)
PHASE STABILITY U N D E R I R R A D I A T I O N 319

13. IRON-BASED ALLOYS, AUSTENITIC


Austenitic stainless steels are of prime interest in both fast reactor core materials and fusion
reactor first walls. Accordingly, many studies have been made of the effects of irradiation on
the microstructural stability of these alloys. Table 6 gives nominal compositions of some of
the commercial and experimental alloys discussed in this section.
Irradiation-induced solute segregation is a key factor in phase stability in these alloys, and
will therefore be discussed first. Then this section discusses the phases observed in types 304
and 316 stainless steels and their behavior under various irradiation regimes. Carbide
precipitation in Nb-bearing alloys, ferrite formation, and some miscellaneous observations
are then discussed. All observations are also summarized in tabular form.

13.1. Solute Segregation


Okamoto and Wiedersich (1974) studied irradiation-induced solute segregation in a
Fe-18w/oCr, 8w/oNi, lw/oSi alloy. Both Ni ÷ ion irradiation and 1 MV e- irradiation were
employed in this early study.
The alloy was irradiated at 600°C with 3.25 MeV Ni + ions to a surface dose of ~ 15 dpa
at a damage rate of ~ 10-3dpa/s. Auger electron spectroscopy showed Ni and Si, both
of which are undersized in stainless steel, were segregated toward the irradiated surface
(Table 7). The sample surface was depleted in Cr, which is oversized. These observations are
thus in agreement with the misfit theory of irradiation-induced solute segregation.
Samples of the same alloy were bombarded at 500°C with 1 MV e- in the HVEM to 0.3 dpa
at a displacement rate of 10 -3 dpa/s. Voids induced by this irradiation showed strong strain
contrast which was attributed to the effect of solute segregation.
Solutions to the diffusion equation were used to estimate the degree of solute segregation
around voids of different sizes, and thereby the resulting strains. Measured and calculated
strains were in reasonable agreement; the greatest strains were predicted to come from Cr
segregation.
Johnston et al. (1977) made Auger electron spectroscopic studies of the surfaces of alloys
irradiated at 675°C with 2.5 x 10164 MeV Ni ions/cm2, or a surface dose of -,,20 dpa. The
irradiated surfaces of both Fe-I 5Cr-20Ni-0.7Si and Fe-15Cr-20Ni were enriched in Ni and
depleted in Cr to depths of ~200/~, in agreement with the results of Okamoto and
Wiedersich (I974). Silicon measurements were not reported.
Brimhall et al. (1981) studied solute segregation in type 316 stainless steel which had been
irradiated with 5 MeV Ni 2÷ ions to a surface dose of 2 dpa at 3 x 10 -3 dpa/s. The surface
concentrations of P and S were increased by several fold at both 775 and 875 K, as compared
to non-irradiated samples. Strong surface segregation of Si was observed at 775 K, and of
Ni at 875 K. The authors note that neither solute-defect binding energies nor atomic misfits
are known for P or S in stainless steel, so comparison of their results on these elements with
theories of solute segregation is not possible.
Sethi and Okamoto (1981) made an extensive study of irradiation-induced segregation in
Fe-20w/oCr-12w/oNi alloys with 1 or 2 w/o of Mo, of Si, or of both. The samples were
irradiated with 3 MeV Ni + ions to 6 dpa at 2 x 10 -3 dpa/s; the surface dose was about 30~o
of this. Irradiation temperatures ranged from 430 to 675°C. Compositions below the surface
were measured with Auger electron spectroscopy after sputtering away the outer material
with 1 keV Ar ions.
Figure 54 shows some of their results; in all cases Si and Ni were segregated toward the
surface and Cr and Mo were segregated toward the sample interior. The former two atoms
320 P R O G R E S S IN MATERIALS SCIENCE

are undersized in the alloy and the latter two are oversized, so that segregation is in agreement
with the size misfit theory. Segregation of Cr, Mo, and Ni all increase with temperature, a
behavior consistent with the greater fraction of defects which migrate to sinks at higher
temperatures and can thereby produce solute segregation. Direct recombination of vacancies
and interstitials predominates at lower temperatures and gives no segregation. Silicon
segregation showed a maximum at 500°C and a minimum at 600°C, followed by a sharp
increase at higher temperatures. This complex behavior was attributed to either a change in
the Si segregation mechanism or a change in the sink structure near the surface, possibly
associated with the formation of Si-rich 7'.

13.2. Types 316 and 304 Stainless Steel


Austenitic stainless steels were the primary early candidates for fuel cladding and duct
materials in the liquid metal fast breeder reactor. Accordingly, the earliest tests of irradiation
response under breeder reactor conditions were made on this alloy. There exists a vast data
base on phase stability of various kinds of austenitic stainless steel under various irradiation
regimes. We restrict the discussion in this section primarily to the response of types 304 and
316 stainless steel and their modifications to neutron irradiation.
Although observations of irradiation-altered phase stability in type 316 stainless steel go
back at least to 1970, it is only with the latest, most sophisticated TEM and x-ray energy
dispersive analysis techniques that a full identification and characterization of phases which
appear during irradiation became possible. Early studies of type 316 and similar stainless
steels irradiated to ca. 1022n/cm 2 (E > 1 MeV) or ~5 dpa in the 450-650°C range often
revealed two precipitate phases: (1) a blocky carbide, of the M23 C 6 type, where M --- (Cr, Fe,
Mo, Ni), ana (2) an acicular phase, probably of the sigma, FeCr type. Some of these early
observations are summarized in Table 8.

13.2.1. Phases in austenitic stainless steel *


Watkin et al. (1977) calculated isothermal sections of the Fe-Ni-Cr ternary phase diagram
by methods developed by Kaufman and Nesor (1975). Stainless steels are positioned in these
diagrams in terms of their equivalent nickel and chromium concentrations, calculated after
accounting for carbide and ?' precipitation.
Figure 55 shows the calculated sectians for 450, 500, and 600°C. In addition to the terminal
solid solutions, the FeCr sigma phase and the low temperature FeNi 3 phases appear under
thermal equilibrium conditions. The common stainless steels are seen to be metastable with
respect to formation of ferrite, sigma phase, or both over the 450-600°C temperature range.
Recent studies by Lee et al. (1981) and Yang et al. (1981) represent the state of the art in
analytical electron microscopy (AEM) studies of the response of type 316 stainless steel to
fast reactor irradiation. This section relies heavily on these two studies. It should be noted
that Williams (1982) has recently provided a systematic data set on SA type 316 stainless steel.
Lee et al. (1981) studied type 316 stainless steel both without and with 0.2-0.25w/oTi which
is added to combine with carbon to form the highly stable TiC phase. Both 20~o cold worked
and solution annealed samples were irradiated in both EBR-II and HFIR, the latter being
a high flux reactor with a large thermal neutron component. Ion irradiations were also
included to up to 100 dpa at a displacement rate of ~ 3 × 10 -5 dpa/s at temperatures from
370 to 700°C. Samples which had been thermally aged to 10,000hr were also studied.
Chemical analysis and x-ray diffraction studies of the various phases were performed on
particles which had been removed from the sample by extraction replication techniques] By
• Weiss and Stickler (1972) performed an exhaustive study of the phases in thermally aged type 316 stainless steel.
PHASE S T A B I L I T Y UNDER IRRADIATION 321
Table 8. Phase stability in austenitic stainless steel
Type of
Stainless Steelt Treatment Result References
SA 316 480°C, 8 x 1021n/cm 2 Rod-like and polyhedral Claudson et aL (1970)
(E > 0.1 MeV) ~ 4 dpa M23C6, needle-like precip-
itates (~r phase?).
SA 316 450--600°C, 2-3 x 1022 M23C6 and needle or lath Bisson and Vouillon (1970)
n/cm 2, ~ 10-15 dpa shaped precipitates.
304 500°C, ~ 4 x 102j M23C6 and cr phase precip- Fish et al. (1970)
n/cm 2 (E > 0.1 MeV), itates.
~ 2 dpa
20~o CW 316 525-700°C, 2.3-3 x 1022 large amounts of a-phase Harries et al. (1971)
n/cm 2, ~ 10-15 dpa at grain triple points.
20~o CW M 316 500°C, 6 x 1022n/cm 2, Intragranular M23C6 precip- Cawthorne et al. (1971)
~ 30 dpa itates.
SA 304, 316 370-600°C, > 1022 M23C6 carbides and rod- Brager et al. (1971)
n/cm 2 (E > 0.1 MeV), shaped particles (a-phase?).
> 5 dpa
20~o CW 316 525°C, 20MeV C 2+, M23C6 precipitates. Mazey et al. (1971)
~ 80 dpa
SA 316, 347 ~550°C, 3-7 x 1022 M23C6 precipitates plus rod Appleby and Wolff (1973)
n/cm 2 (E > 0.1 MeV), shaped particles.
15-35 dpa
SA 316 580-700°C, 1.9 x 1022 M23C6 and ribbon-shaped Bloom and Stiegler (1973)
n/cm 2 (E > 0.1 MeV) precipitates.
~ 10 dpa
SA 316 525-625°C, up to 4 x 1022 M23C6 and a phase. Brager and Straalsund (1973)
n/cm 2 (E > 0.1 MeV), up
to ~ 20 dpa
SA M316, M316L 600°C, 1022n/cm 2, Rod shaped precipitates. Brown (1975)
5 dpa
SA and 20~ CW 370-700°C, EBR-II and Chemical and crys- Lee et aL (1981)
316 and 3 1 6 + T i HFIR, up to 100dpa tallographic character-
ization of all irradiation-
stable phases.
20% CW 316 467-650°C, EBR-II, up Chemical and crys- Yang et al. (1981)
to 1.4 x 1023n/cm 2, taUographic character-
(E > 0.1 MeV), up to ization of all irradiation-
70 dpa stable phases.
20~ CW 316 380-610°C, 1.5-60 q-phase is only carbide type Maziasz (1979b)
dpa, HFIR phase.
SA 316 550-680°C, 30-47dpa, Phases form at lower tem- Maziasz (1979a)
HFIR peratures than thermally or
in EBR-II.
20~o CW 316 550°C (nominal) 650°C Evolution consistent with Brager and Garner (1982a)
(est.), 42 dpa, HFIR EBR-II irradiation at
~ 650°C.
20~o CW 316 375°C, HFIR y' forms after 8.5dpa, dis- Maziasz et al. (1981)
appears after 13 dpa.
20~o CW 316 625°C, EBR-II, 8.4dpa Copious r/and Laves phase Maziasz et aL (1981)
+ 110 appm He formation
+SA = solution annealed,
CW = cold worked.

using this technique, V radiation from the matrix was avoided as was interference with
chemical and crystallographic analysis by the matrix and other precipitate phases. The nine
phases which appear under irradiation are classified according to the following scheme:
Radiation induced: the phase (e.g., V', G-phase) does not form except under irradiation.
Radiation enhanced: irradiation enhances formation of the phase, as compared to thermal
treatment (r/phase).
Radiation modified: irradiation modifies the composition of the phase, usually by solute
segregation to the particle:matrix interface (Laves, M23C6).
J.P M.S. 2813-4- F
322 PROGRESS IN M A T E R I A L S SCIENCE

---- Tie lines / -~ --- Tie line~, /" "~


Phaseboundorie/ ~x Phaseboundaries/ ~,

,,,% ,., \ \

.~?
....:7 " ...k,,.
/. "- h'5
4":l /'. "',, "\A°~-.. ,

Fe Wt. fraction Cr-- Cr Fe Wt. fraction Cr-- Cr

(o) (b)

--- Tie Lines ,/- ~


,. ZA
Phase,oooOa, \
cc \

,~l /-.,. ",,.\~o~

Fe Wt fraction Cr- Cr
(c)
FIG. 55. Calculated isothermal sections of Fe-Cr-Ni ternary phase diagrams. Locations of several
commercial stainless steels are shown in the diagram. (a) 600°C. (b) 500°C. (c) 450°C. (After Watkin
et al., 1977.)

Lee et al. (1981) categorize the phases as follows:


7' phase: Radiation induced; (Brager and Garner, 1983d) disappears on post-
irradiation annealing.
G-phase: Radiation induced.
r/-phase: Radiation enhanced.
M23C6: Neither radiation enhanced nor modified.
Laves phase: Radiation enhanced, modified by absorbing extra Ni and Si during
irradiation.
MC phase: Radiation enhanced but not modified; may be suppressed by G-phase
formation.
a and FeEP phases: Insufficient data to categorize, although results suggest that tr is
enhanced by irradiation in HFIR and Fe2P is enhanced and modified
by irradiation in EBR-II.
PHASE STABILITY UNDER IRRADIATION 323

Table 9. Summary of crystallographic data of precipitate phases in type 316 stainless steel
Lattice
Crystal Parameter Solute Atoms Typical Orientation
Phase Structure (nm) per Unit Cell Morphology To ?-Matrix
7 Cubic, AI, Fm3m ao = 0.36 4 Matrix --
7' Cubic, LI 2, Fm3m ao =0.35 4 Small Sphere Cube-on-Cube
G Cubic, A1, Fm3m ao = 1.12 116 Small Rod Random
Fe2P Hex., C22, P321 ao = 0.604 6 Thin Lath (l]10)ppt//(011)~
Co= 0.36 (0001)ppt//(001)r
~/ Cubic, E93, Fd3m ao = 1.08 96 Rhombohedral Cube-on-Cube or
Twin
Laves Hex., C14, P63/mmc ao = 0.47 12 Faulted Lath Many Variants
co = 0.77
M23C6 Cubic, D84, Fm3m ao 1.06
= 92 Rhombohedral Cube-on-Cube or
Platelet Twin
MC Cubic, BI, Fm3m ao = 0.433 4 Small Sphere Cube-on-Cube
a Tet., D8b, P4/mnm ao = 0.88 30 Various Many Variants
co = 0.46
Z Cubic, AI2, l~13m ao = 0.89 58 Various Many Variants
TiN Cubic, B1, Fm3m ao = 0.425 4 Large Cuboid
Zr4C2S2 Hex., P63/mmc ao = 0.34 Globular
co = 1.21
ZrO Cubic, B1, Fm3m ao = 0.46 4 Globular
After Lee et al. (1981)

Table 9 summarizes the crystallographic and morphological characteristics o f the phases


f o u n d by Lee et al. The TiN, Zr4C2S2, and Z r O phases had formed during fabrication o f the
alloy, so that a total o f nine phases were observed to f o r m during irradiation.
The ordered 7' phase has a stoichiometric composition o f N%Si, but part o f the Si m a y
be replaced by A1, Nb, or other elements. Increasing the Si content increases both the a m o u n t
o f ? ' and the temperature range over which it is formed.
The G phases have the general formula T6Ni16Si7, where T is a transition element. Y a n g
et al. (1981) f o u n d the transition elements to be Cr and M n in n e u t r o n irradiated type 316
stainless steel.
The ~/-phase is o f the type M6X or MsSiX, where in the latter case Si occupies a metal site
and the X sites are vacant.
M23C6, also k n o w n as a z phase, m a y have Cr, Fe, M o , W, and possibly Ni and other metal
a t o m s in the M position.
Laves phases have the general formula A2B; Fe2Mo containing significant a m o u n t s o f Si
and Cr is often f o u n d in thermally treated M o - b e a r i n g stainless steels. U n d e r irradiation the
Laves phase also contains significant additional a m o u n t s o f Ni and Si.
Lee et al. (1981) note that the X phase, a F e - C r - M o c o m p o u n d , is chemically and
structurally related to the tr phase.
The M C carbide m a y form from a variety o f transition metals, including Ti, V, Nb, or Ta,
and as such would form in type 316 stainless steels only if modified with Ti or Nb.
The Fe2P phase was stated to be one o f a variety o f binary transition metal phosphides.
Figures 56-63 are electron micrographs o f the phases which a p p e a r in irradiated type 316
stainless steel. The crystal structures and compositions o f the phases are generally in
agreement with the conclusions o f Lee et al. (1981) except for the p h o s p h o r o u s - r i c h phase.
Lee et al. identified this phase as hexagonal Fe2P, whereas on the basis o f the diffraction
patterns in Fig. 63b Y a n g identified it as the o t h e o r h o m b i c FeP. The two investigations m a y
have involved two different Fe and P-rich phases.
324 P R O G R E S S IN MATERIALS SCIENCE

13.2.2. Temperatures of appearance of phases


The temperature and order of appearance of the various phases which appear under
irradiation have been found sensitive to a number of variables, including:
Small variations in composition of the steel,
Pre-irradiation condition of the steel,
Neutron flux and spectrum,
Applied stress,
Temperature changes during irradiation.
A detailed description of irradiation altered phase stability in stainless steels under all the
conditions in all the various studies will not be attempted here. Those interested in details
are referred to the original papers, many of which appeared in the Journal of Nuclear
Materials during the past decade, and to the numerous Proceedings of symposia on
irradiation effects in metals and alloys listed among the references.
Here we will give the results of Yang et al. (1981), who studied phase stability in three heats
of type 316 stainless steel which had been cold worked 20~ and irradiated to up to
14 x 1022n/cm 2 (E > 0.1 MeV) or ~ 7 0 d p a in EBR-II at temperatures of 467-650°C.
Table 10 shows only minor variations in composition between the three heats of steel.
Table 11 shows the response to irradiation to be very different, although samples of all three
heats were irradiated in the same capsule in the same experiment.
In general, the 7' and M23C6 phases were found to dominate at the lower temperatures and
~/-silicide, G, and Laves phases at the high temperatures. Striking heat-to-heat variations were
observed as, for example, no G-phase was observed in the R-lot heat. The R-lot steel also
had both the least 7' and the smallest range of stability for that phase of any of the three
heats. The authors also noted that both the intermetallic and 7' phases form relatively more
slowly in the N-lot steel than in the other two heats.
Thus, one is faced with the question of why three steels, of virtually identical compositions,
irradiated in the same capsule, would behave so differently. The mechanism for the differences
is not known, and at the moment it may be best to consider these heat-to-heat differences
in irradiation response as an illustration of the dictum by Landauer (1978) that such driven
systems are highly susceptible to noise. In this case the "noise" is probably the minor
heat-to-heat variations in composition and presumably microstructure which cause far
greater differences in precipitation behavior than would be observed under purely thermal
conditions.

13.2.3. Microchemical evolution


Garner (1981) and Brager and Garner (1981a) have constructed a scenario in which the
matrix chemistry changes in a complex way while various precipitation processes occur.
Matrix chemistry is of particular importance in dimensional stability of reactor materials

Table 10. Composition of various heats of 20~o cold worked AISI 316 stainless steel
Lot No. C Mn Si P S Cr Ni Mo Cu N Co B
N-Lot 0.056 1 . 6 4 0.46 0.013 0.006 16.52 13.66 2.41 0.08 0.006 0.05 0.0008
R-Lot 0.04 1.36 0.52 0.011 0.009 17.76 13.27 2.53 0.08 0.012 0.032 0.001
CN-13 0.053 1.63 0.50 0.002 0.003 17.26 13.72 2.25 <0.01 0.004 0.01 <0.0005
Data from Yang et al. (1981).
PHASE STABILITY U N D E R I R R A D I A T I O N 325

Table 11. Precipitate phases* observed in 20~ cold worked AISI 316 stainless
steel irradiated in the RS-1 experiment to 14 x 1022n/era 2 (E >0.1 MeV)
Irradiation
Temperature
(°C) CN- 13 N-Lot R-Lot
467 )" ~' M23C6
M23C6 M23C6 17
G ~/
t/ G

500 7' ~' M23C6


M23C6 M23C6 y'
G ~/ ~/
G

533 M23C6 ~' ~'


T' M23C6 + M6C** M23C6
G ~/ Y/
G

567 G M23C6 .4- M6C** M6C + M23C6"*


r/ ),' t/
M23C6 t]
G

600 M23C6 q- M6C** M6C -I- M23C6"* M6C + M23C6"*


Laves Laves Laves
t/ t/ t/
G G

650 Laves M6C + M23C6"* Laves


M6C + M23C6"* Laves Small amounts
~/ ~7 of M23C6 +
G G M6C** and

*Phases are listed in order of predominance.


**Both phases appear simultaneously and usually in association with each
other.
Data from Yang et al. (1981).

since increasing levels of Ni and Si in solid solution are known to temporarily suppress void
swelling (Garner, 1984).
At low temperatures carbon delays the onset of swelling in type 316 stainless steel by
perhaps interfering with the formation of the Ni-Si-Mn-rich G phase. By contrast, ternary
Fe-Ni-Cr alloys near the composition of type 316 stainless steel begin to swell very quickly.
At high temperatures, carbon promotes swelling by promoting a precursor to the formation
of such intermetallic phases as the Laves, which also becomes rich in Ni and Si during
irradiation.
Maziasz (1982), however, argues against carbon-rich precursors on the basis of obser-
vations on type 316 stainless steel irradiated in EBR-II.
Garner notes that except for MC (TiC in Ti-modified steels) and possibly M23C6 all phases
in type 316 stainless steel become progressively enriched with Ni and Si during irradiation.
Okamoto and Rehn (1979) postulated that slowly diffusing Ni concentrates at sinks by both
inverse Kirkendall effect and interstitial binding. Garner (1981) proposed mechanisms for
some of the compositional changes which occur during microchemical evolution. The 7'
326 PROGRESS IN MATERIALS SCIENCE

(Ni3Si) phase forms on dislocation lines and loops, but not on voids. Garner stated that,
unlike dislocation lines and loops, void surfaces are not enriched in Si. This difference in
behavior between the two sinks was taken to show that fast diffusing Si segregates not by
the inverse Kirkendall effect, but by Si:interstitial binding (see Section 5). The excess in
vacancy over interstitial flux at voids (the reverse occurs at dislocations) would then result
in segregation of the fast-diffusing Si to dislocations but not to voids. However, Okamoto
and Wiedersich (1974) and L. Thomas (1982) have found segregation of Si to void surfaces.
As such, there is doubt as to the mechanism (or mechanisms) of irradiation-induced
segregation of Si in type 316 stainless steel.
It has also been shown that the solute removal and phase evolution processes are sensitive
to temperature history (Garner et al., 1982; Yang and Garner, 1983), displacement rate
(Garner and Porter, 1983), stress (Garner and Porter, 1983; Garner, 1981b) and pre-
irradiation thermal mechanical treatment (Garner and Porter, 1983; Brager and Garner,
1979).
Dislocations may alter phase evolution by interaction with carbon (Garner, 1981a).
Interstitial carbon atoms tend to segregate to the stressed region near the dislocation core;
cold working may thus tie up a large fraction of the carbon atoms in the alloy and prevent
their participation in the formation of the carbon-rich precipitates which play an important
part in microchemical and microstructural evolution.

13.2.4. Neutron spectrum and helium effects


The High-Flux Isotope Reactor (HFIR) has a much larger thermal neutron flux than does
EBR-II (see Fig. 4). This thermal flux in HFIR has the effect of producing ca. 100 times more
He per dpa, burning out most of the manganese, and producing significant amounts of
vanadium during typical irradiations (Brager and Garner, 198 l b,c). There are also significant
differences in the flux-temperature combinations available in the two reactors.
Maziasz (1979a,b) reported on HFIR irradiation of both solution annealed and 20~o CW
type 316 stainless steel. On comparing HFIR and EBR-II results on the solution annealed
material, Maziasz found that for the same irradiation times and temperatures, the HFIR
samples differed in containing M6C and Laves phases and in not containing an unidentified
rod-shaped phase. In the case of 20~ CW material irradiated at 380-610°C, Maziasz found,
contrary to expectations, that r/-phase was only carbide type phase.
Maziasz (1979b) noted that temperatures in HFIR are not measured but are instead
calculated from nuclear heating values, and that new results indicate that the actual
temperatures might be as much as 75 ° higher than stated in his work. Brager and Garner
(1982) used the rounded shape of voids, characteristic of high temperatures, and the presence
of the high temperature X and a phases to suggest that the actual HFIR temperature in the
Maziasz study was 100°C or more above the calculated value. Upon accepting the higher
temperature, Brager and Garner concluded that there was no clear effect of neutron flux
spectrum effect on phase stability in type 316 stainless steel and subsequently (1983a,b) they
reported measurements on both annealed and cold worked specimens to support their
conclusion.
Maziasz et al. (1981) made comparison studies of the response of 20~o CW type 316
stainless steel (with and without Ti) to irradiation in EBR-II and HFIR. They also
pre-injected some of the EBR-II specimens with 110 appm He to determine the effects of high
He concentrations and the associated injection damage on microstructural evolution in
EBR-II. It was found that He pre-injection triggered copious ~/and Laves phase precipitation
in EBR-II after 8.4 dpa at 625°C. Samples irradiated at 500 or 625°C without pre-injected
rv

-]
>.

,<

>.

FIG. 56. Type 316 stainless steel neutron irradiated to 25dpa at 650°C showing M23C6 phase particles. (a) Bright field; (b) dark field; (c) electron diffraction
patterns. (Electron micrographs and diffraction patterns courtesy W. J. S. Yang.) --a
328 PROGRESS IN MATERIALS SCIENCE

(001)7" tl (001)M23C 6 (01 1 )7. II (011 )M23C 6

(c)
FIG. 56(C). For legend see previous page.
t~

C
Z
=

(b) (212)y II (O01).r/ (~1) 7, II (110)~


FIG. 57. Type 316 stainless steel neutron irradiated to 26 dpa at 400°C showing ~ (M6C) phase particles. (a) Bright field; (b) electron diffraction patterns. (Electron
micrograph and diffraction patterns courtesy W. J. S. Yang.)
0

~n

(001)T II (011 )G
>
-1

(b)(111)¥ti (011) G (o11) 7~ (OOl) G

FIG. 58. Titanium modified type 316 stainless steel neutron irradiated to 25 dpa at 600°(; showing G-phase particles. (a) Bright field; (b) electron diffraction
patterns. (Electron micrograph and diffraction patterns courtesy W. J. S. Yang.)
FIG. 59. Titanium modified type 316 stainless steel neutron irradiated to 61 dpa at 600°C showing Laves phase particles. (a) Bright field; (b) electron diffraction
patterns. (Electron micrograph and diffraction pattern courtesy W. J. S. Yang.)

ta..}
332 PROGRESS IN M A T E R I A L S SCIENCE

Fxo. 60. Titanium modified type 316 stainless steel neutron irradiated to 50 dpa at 430°C showing
7' phase particles. (a) Dark field; (b) electron diffraction pattern showing cube-cube orientation
relationship between 7 and 7' phases. (Electron micrograph and diffraction pattern courtesy W. J.
S. Yang.)
t"
-]

Z
~7

~r

FIG. 61. Titanium modified type 316 stainless steel neutron irradiated to 15 dpa at 600°C showing MC phase particles. (a) Dark field; (b) electron
diffraction pattern. (Electron micrograph and diffraction pattern courtesy W. J. S. Yang.) t.~
334 P R O G R E S S IN M A T E R I A L S SCIENCE

F16. 62. Type 316 stainless steel neutron irradiated to 31 dpa at 710°C showing sigma phase. (a)
Bright field; (b) electron diffraction pattern. (Electron micrograph and diffraction pattern courtesy
W. J. S. Yang.)
PHASE S T A B I L I T Y U N D E R I R R A D I A T I O N 335

500 nm

FIG. 63. Titanium ,modified type 316 stainless steel neutron irradiated to 25 dpa at 650°C showing
FeP phase particles. (a) Bright field, showing FeP particles on cube planes of grains of three different
orientations; (b) electron diffraction patterns. (Electron micrograph and diffraction patterns courtesy
W. J. S. Yang.)
0

r~

>.
-t

>

FIo. 64. NbC precipitation and void formation in FV 548 steel solution annealed and irradiated to 20 dpa with 22 MeV C 2÷ ions after pre-injection
with 10 appm He. (Electron micrograph courtesy T. M. Williams.)
PHASE S T A B I L I T Y U N D E R I R R A D I A T I O N 337

FIG. 65. NbC precipitation in stainless steel irradiated in the 1 MV HVEM to 1.8 dpa. (a) 20/25/Nb
steel irradiated at 400°C; (b) 20/28/Nb steel irradiated at 450°C showing precipitates and helices;
(c) EN58G steel irradiated at 450°C showing grain boundary and matrix precipitation. (Electron
micrographs courtesy S. B. Fisher.)
0

~J~

>

r~

FIG. 65 (b-c). See previous page for legend.


PHASE STABILITY UNDER IRRADIATION 339

He showed sluggish precipitation kinetics. H F I R irradiation of CW material produced more


r/and Laves phases, and produced them at a lower temperature than did thermal aging. It
was found that y' formed more rapidly at 375°C in H F I R than in EBR-II. However, the y'
which formed after 8.5 dpa in H F I R disappeared after 13 dpa and was replaced by the t/
phase. The authors noted that irradiation-induced ),' is stable in EBR-II to at least 70 dpa.
Maziasz (1982) made a comparative study of precipitation in solution annealed and 209/o
CW type 316 stainless steel under H F I R and EBR-II irradiation, and under long term thermal
aging. He found that, as in earlier studies, precipitation generally was much faster in H F I R
than in EBR-II. However, calculations of the displacement rate in H F I R did not include those
displacements due to recoils from 5qXlitransmutation reactions. When such displacements are
included, the microstructural evolution--at a given dpa level--is slower than was originally
thought.
Maziasz noted that H F I R irradiations produced a large number of small helium bubbles
which have a high sink strength for point defects, and proposed that this increased sink
strength reduced the rate of solute segregation at any particular sink and thereby reduced
the degree of segregation of Ni and Si.
There thus exist two points of view on the possibility of neutron flux spectrum effects on
phase stability in austenitic stainless steel. Maziasz (1979a,b, 1982), Maziasz et al. (1981) and
Farrell et al. (1983) claim strong effects due to a helium-induced change in the solute
segregation and provide extensive results to support their claim. Brager and Garner (1982a,b,
1981b, 1983a,b) claim that the observed effects cannot be simply explained by differences in
He/dpa ratio but are the combined result of variations in flux dose, solid and gaseous
transmutants, temperature history and other variables. Both groups agree that the precipitate
phases produced in H F I R occur on a finer scale than in EBR-II. Brager and Garner conclude
that the correlation between matrix evolution and swelling is unperturbed by minor
differences in phase evolution (Brager and Garner, 1983a,b; Garner, 1984). They also
conclude that the eventual void swelling rate is totally insensitive to pre-irradiation or
irradiation-induced variations in microstructure and matrix composition (Garner, 1984;
Garner and Wolfer, 1984).
13.3. Precipitation of NbC
The volume per Nb atom in NbC is much larger than the atomic volume in austenitic
stainless steel (6 = 0.88; see Section 4.1), so that co-precipitation of vacancies is required for
formation of intragranular NbC without prohibitive strain energy. A vacancy supersaturation
should then contribute to the thermodynamic driving force for NbC formation and give a
large increase in the rate of homogeneous nucleation of NbC (Section 4.1).A number of such
studies has been made, in quenched or irradiated alloys. Of technological interest, T. M.
Williams et al. (1982) have found that a high density of fine intragranular NbC particles is
effective in suppressing void swelling.
Shepherd (1976) studied precipitation in two Nb-bearing alloys. The compositions were
chosen so that NbC would dissolve on annealing at temperatures above ~ 1200°C.
An anneal at 1300°C followed by a quench to 700°C or below, where the alloy was held
for 50-1000 hr, gave ,,, l015 NbC particles/cm 3 in the phosphorous bearing alloy (Table 12).
Lower annealing temperatures or slower quenches gave lower particle number densities. The
other alloy gave an order of magnitude lower particle number density than the phosphorous
bearing alloy under all conditions.
These results were explained in terms of co-precipitation of vacancies with Nb and C.
Higher annealing temperatures and quench rates are known to give higher vacancy
340 PROGRESS IN MATERIALS SCIENCE

Table 12. NbC precipitation in austenitic stainless steel


Alloy*
(compositions in w/o) Treatment Result References
26Ni, 20Cr, 0.59Nb, Quench from 1300, 1250, ~ 1015NbC/cm3 for Shepherd (1976)
0.05C, 0.008P; 1200 to 700°C high P alloy fast-
24Ni, 20Cr, 0.04C, quenched from 1300°C.
0.08P Lower densities under
other conditions and for
low-P alloy.
24.8Ni, 20.7Cr, 0.7Nb, 50°C, 7.4 x 10 t7 f a s t Copious NbC precip- Higgins and Roberts
0.025C n/cm2, ~ 4 x 10-4dpa, itation in matrix, ten- (1966)
followed by 850°C dency for alignment.
anneal
12Ni, 16Cr, 1.4Mo, ~600°C, 1-2 x 102°n/cm2, Copious NbC precip- Shepherd (1969)
0.9Nb, 0.1C ~0.05-0.1 dpa itation in matrix.
~6 x 10-gdpa/s
SA FV 548 stainless steel, 500, 600, 700°C, 22 MeV Copious NbC precip- T. M. Williams et al.
C2÷, 20 dpa itation. (1977)
20~ CW FV548 stainless 500, 600°C, 46.5 MeV Irregular distribution of T. M. Williams and
steel Ni6÷, 20--180dpa; 500- NbC in matrix. Arkell (1979)
600°C, 22 MeV C2+,
20-40 dpa
25.5Ni, 20.4Cr, 0.6Nb, 350-500°C, 1 MV e-, Precipitation of NbC in K.R. Fisher and Will-
0.1C; 9.9Ni, 17.9Cr, 10-3 dpa/s, 1-2 dpa matrix and at hetero- iams (1972)
0.75Nb, 0.07C geneities.
SA FV 548 stainless steel, 200-650°C, 1 MV HVEM, Sparse NbC precip- T. M. Williams and
up to 78 dpa itation at most. Eyre (1976)
*SA = Solution annealed.
CW = Cold worked.

concentrations and supersaturations, and phosphorous helped maintain the vacancy super-
saturation by forming V - P complexes and preventing vacancy loss to sinks.
Irradiation has been found, in a number of cases, to give similarly high densities of N b C
particles in stainless steels. While irradiation-induced vacancies m a y enhance N b C nucleation
as effectively as if from a quench, other mechanisms for irradiation-altered phase stability,
such as solute segregation and nucleation on vacancy-rich cores of displacement cascades,
may also come into play.
Higgins and Roberts (1966) subjected a Nb-stabilized stainless steel to 7.4 x 1017 fast n/cm 2
(,,~4 × 10 -4 dpa) at 50°C. The samples were heated to 850°C, held 30 min, and subjected to
tensile tests.
Copious N b C precipitation occurred during the warmup and hold at 850°C, with a
tendency for the particles to be aligned along ( 1 1 0 ) directions on { 111 } planes or along (112)
on {110} planes. Such alignment did not occur in control samples. Higgins and Roberts
explained the banding as being due to irradiation induced segregation to dislocations,
followed by N b C precipitation. Particles not associated with dislocations probably formed
by co-precipitation of supersaturated irradiation-induced vacancies with N b and C atoms.
Shepherd (1969) irradiated a Nb-bearing alloy at ,-,600°C. Irradiation was to
1-2 × 102°n/cm 2 (,-,0.05-0.1dpa) over a period of 500hr, for a displacement rate of
--,6 × 10-s dpa/s.
Particles of N b C formed homogeneously in the matrix and on stacking faults in irradiated
specimens but not in non-irradiated control specimens with the same thermal history. The
highest N b C particle number densities were observed in samples with the highest pre-
irradiation annealing temperature (1300°C). Such behavior is expected, since these samples
would have more N b and C in solid solution to precipitate during irradiation and subsequent
PHASE S T A B I L I T Y U N D E R I R R A D I A T I O N 341

Table 13. NbC precipitation in irradiated


FV548 stainless steel
T dpa diam (A) density (cm -3)
500°C 20 70 1.3 × 1016
600°C 20 130 7.4 x 1015
700°C 20 440 2.2 x 1015

annealing. In samples given a 1300°C pre-irradiation anneal, either no, or a very narrow,
precipitate-free zone was observed near the grain boundaries, which were decorated with large
M23C6 particles. Precipitate-free zones were observed near grain boundaries in specimens
which had received lower temperature pre-irradiation anneals. At the lower annealing
temperatures more NbC and M23C 6 could form on grain boundaries, depleting the nearby
matrix of Nb and C and giving NbC-depleted zones on irradiation.
Shepherd stated that at 600°C, vacancies are lost too rapidly to be present in super-
saturation. Recent calculations by Parker and Russell (1982) have shown, however, that
substantial vacancy supersaturations may persist at 600°C in irradiated austenitic stainless
steels. As such, the enhanced matrix precipitation of NbC was probably due to reduction in
the nucleation barrier by co-precipitation of supersaturated vacancies with solute atoms.
T. M. Williams et al. (1977) irradiated solution annealed FV548 steel (Table 6) with 22 MeV
C 2÷ ions after pre-injection with 10 appm He. High densities of fine NbC precipitates were
observed, as shown in Fig. 64 and summarized in Table 13. The particles are most numerous
at low temperatures where high solute and vacancy supersaturations would give the highest
nucleation rate. T. M. Williams et al. (1977) believed that NbC nucleated on faulted
dislocation loops or the centers of displacement cascades.
T. M. Williams and Arkell (1979) studied the effects of 46.5 MeV Ni 6+ and 22 MeV C 2+
heavy ion irradiation on NbC precipitation in a FV548 stainless steel. The primary purpose
of the experiment was to study void swelling, so all samples were pre-injected with 10 appm
He before irradiation. All samples were solution annealed and cold worked 20% prior to
irradiation; some samples were then also given a second anneal at 650°C for 1000 hr.
Although intragranular particles of NbC formed during irradiation, the distribution was
less uniform than in similarly irradiated solution annealed samples (T. M. Williams et al.,
1977). The absence of intragranular NbC particles in regions of high dislocation density is
consistent with these sinks absorbing the vacancies needed to catalyze such NbC nucleation.
Fisher and K. R. Williams (1972) bombarded two solution-annealed, Nb-bearing stainless
steels denoted 20/25/Nb and EN58G (see Table 6) with 1 MV electrons in the HVEM.
Irradiations were in the 350-500°C temperature range at a displacement rate of 10 -3 dpa/s.
Figure 65 shows that copious matrix precipitation of NbC was observed as was hetero-
geneous precipitation on grain boundaries and dislocations. Fisher and K. R. Williams noted
that since HVEM irradiation does not produce displacement cascades, their results showed
that such cascades are not needed to nucleate high densities of intragranular NbC particles
in irradiated materials. This conclusion is consistent with Shepherd's (1976) observation of
copious NbC precipitation after quenching.
Fisher and K. R. Williams (1972) concluded that irradiation-enhanced diffusion was
responsible for the observed enhanced precipitation of NbC. However, although the
precipitate nucleation rate does depend somewhat on the diffusion coefficient, the con-
tribution of supersaturated vacancies to the thermodynamic driving force for nucleation
would have a far more powerful effect.
T. M. Williams and Eyre (1976) irradiated solution annealed FV 548 stainless steel
342 PROGRESS IN MATERIALS SCIENCE

(Table 6) in the 1 MV H V E M at temperatures of 200-650°C. All samples were pre-injected


with 10 appm He before irradiation to 28-78 dpa at 5 x 10 -3 dpa/s.
In no case was NbC precipitation observed on the fine scale found under neutron and heavy
ion irradiation and in samples quenched without irradiation. Fine NbC precipitation only
under neutron and heavy ion precipitation would suggest that the residues of displacement
cascades are needed for nucleation sites. However, copious precipitation of NbC in quenched
alloys (Shepherd, 1976) and in the H V E M (Fisher and K. R. Williams, 1972) showed that
displacement cascades are not required as nucleation sites. As such, the absence of fine NbC
precipitation in this study of electron irradiated FV548 stainless steel is puzzling.

13.4. Ferrite Formation


Figure 55 shows that the common stainless steels are metastable with respect to ferrite
formation at temperatures below 500°C, and some are metastable even at 600°C. Thermal
diffusion of substitutional atoms is very sluggish at these temperatures, so that ferrite
formation is usually very slow. Irradiation, however, has been found to produce significant
amounts of ferrite or of a magnetic phase which is sometimes identified as ferrite.

13.4.1. Proton irradiation


Keefer et al. (1971, 1972) and Keefer and Pard (1972, 1973) subjected 12.7/lm foils of type
321 stainless steel to a 0.75 or 1 MeV proton flux. Irradiation was to 1-50 dpa at 400 and
500°C. Displacement rate depended on position in the foil, and varied from 1.8 x 10 -4 to
7.2 x 10 -4 dpa/s.
Large areas of ferrite grains, ~ 1/~m diameter were found in the peak damage region of
the samples irradiated to 5 or 50 dpa at 500°C. Considerably less ferrite was found in the
off-peak region irradiated to 1 dpa at 1.8 x 10 -4 dpa/s and in foils irradiated at 400°C. Voids
were found in the ferrite region. Voids do not form readily in ferrite, so their presence was
taken as indicative of the ferrite forming late in the irradiation after the voids had already
nucleated.
Keefer et al. (1972) irradiated 12.7 # m thick foils of type 316 stainless steel with 0.75 or
1 MeV protons at 500 and 600°C. The foils were pre-injected with 2-9 appm He before proton
irradiation; damage was to 0.46-100dpa at displacement rates from 1.5 x 10 -4 to
12 x 10 -4 dpa/s. Ferrite formed at ~ grain boundaries in a sample irradiated at 600°C to
,-,20 dpa at 7-9 x 10 -4 dpa/s, but was not observed after 500°C irradiation. Voids were not
observed in the ferrite, which suggests that it formed early in the irradiation. The stated
displacement rate indicates that the ferrite formed in the peak displacement region of the foil.
In the study of Keefer and Pard (1972, 1973) foils of type 316 stainless steel were
pre-injected with 0-28 appm He, and irradiated at 500°C with protons of unstated energy,
but probably in the 0.75-1 MeV range. Damage was to 11-13 dpa at displacement rates of
0.74 x 10 -3 to 1.2 x 10 -3 dpa/s. Areas of ferrite ,-, 1 # m in diameter formed in a sample which
had received three incremental injections of helium (to a total of 3 appm) during irradiation
to 24 dpa at 1.2 x 10 -3 dpa/s. The ferrite formed primarily between 6.9 and 24 dpa, after the
third increment of helium. It was noted that more ferrite formed in this sample than in the
study of Keefer et al. (1972).
In interpreting the results of these and other proton irradiation experiments, it is well to
keep in mind the results of the calculations of Lam et al. (1981). The extremely non-uniform
displacement rate in the peak damage region in proton irradiation was found to give very
strong transient solute segregation effects. In particular, Si, which promotes ferrite formation,
PHASE STABILITY UNDER IRRADIATION 343

was found to be strongly segregated to the peak damage region and may thus be at least partly
responsible for ferrite formation in proton-irradiated type 316 stainless steel.

13.4.2. Neutron and heavy ion irradiations


Changes in magnetic properties are a sensitive indicator of formation of ferromagnetic or
superparamagnetic phases. Several investigators have used magnetic measurements to study
irradiation-induced phase transformations.
Reynolds et al. (1955) subjected up to 50~ cold worked type 347 stainless steel to a neutron
flux at T < 100°C. Exposures were up to 100 kW-hr/cm 2 of neutrons of unstated energy. If
one assumes the neutrons to have an energy of ~ 1 MeV, 100 kW-hr/cm 2 would produce
~ 2 dpa. The samples contained some ferrite before irradiation. Changes in saturation
induction were taken to be due to irradiation-induced ferrite. Up to 0.06~o additional ferrite
was formed by irradiation, with the amount generally increasing with neutron fluence and
degree of cold work. The small amounts of ferrite could well have formed due to cold work
which may occur during irradiation experiments.
J. L. Baron et al. (1974) subjected a type 316L stainless steel to neutron irradiation in the
Rapsodie fast test reactor. Both temperature and fluence varied over the length of the sample.
Temperatures were 400-500°C, and fluences were from < 1 to 7 × 1022n/cm 2 (presumably
most of E > 0.1 MeV). The largest change in magnetic permeability was at 490°C where the
neutron fluence was ~ 4 × 1022n/cm 2 (~20 dpa). The authors suggested that the ~ 800 appm
of transmutation-induced hydrogen helped stabilize the ferrite.
In several studies, magnetic measurements were used to obtain the number density of
magnetic particles. Stanley and Garr (1975) subjected type 316 stainless steel to
1.8 × 1022n/cm2 (,-~9dpa) at 425°C and 3.5× 1022n/cm2 (,-~18dpa) at 500 and 600°C
(E > 0.1 MeV). The samples were cut from tensile specimens which had been tested at various
temperatures, and had thus received post-irradiation anneals. The beginning microstructures
were:

Solution annealed, no precipitates.


Solution annealed, carbides at grain boundaries.
Mostly recrystallized, carbides at grain boundaries, a phase in matrix.
No precipitates, cold worked.
Carbides at grain boundaries, cold worked.

The shape of the magnetization versus field curves indicated large numbers of fine,
super-paramagnetic particles. Magnetic measurements indicated particle number densities of
0.4-4.3 × 10 TM per c m 3 with diameters of 15-30 ~. A maximum magnetization equivalent to
3.6~o ferrite was found in a specimen which originally contained carbides and a phase, and
was given a 1 hr post-irradiation anneal at 500°C after irradiation at 425°C. Stanley and Garr
noted that the change in magnetic properties could result from causes other than ferrite
formation.
Stanley (1979) subjected solution annealed type 316 stainless steel to neutron fluences of
1-4.5 x 1022n/cm2 (E > 0.1 MeV), or about 5-22dpa, at temperatures between 698 and
890 K. Figure 66 shows magnetization versus temperature-reduced field for two of the
samples. The low-fluence sample shows the linear behavior characteristic of a paramagnetic
solid, while the high fluence behavior is that of a super-paramagnetic solid which contains
many small magnetic particles. A particle number density of some 3 × 1017/cm.3 w a s calculated
from the magnetic measurements.
344 PROGRESS IN M A T E R I A L S SCIENCE

40

r~
3O

2C
E

e.o I I
2 4 6 8 ib IJ2
M.0H/T, miIli Tes4(]/*K

FIG. 66. Magnetization vs. temperature reduced field measured at 300K for irradiated type 316
stainless steel. (A) Irradiated to 4.5 x 1022n/cm 2 (E > 0.1 MeV) at 773 K, showing supeq)arama~etic
behavior; (B) Irradiated to 1.90x 1022n/cm 2 ( E > 0 . 1 M e V ) at 873K showing paramagnetic
behavior. (After Stanley, 1979.)

Stanley argued that for two reasons the particles were neither martensite nor ferrite. Firstly,
the austenite matrix did not contribute to the high field magnetic behavior. Since only ~ 2~o
ferrite would have been needed to produce the magnetic particles, the composition of the
austenite matrix would have been such as to produce a contribution. Secondly, the Curie
temperature of the particles was too high for either ferrite or martensite. On this basis, Stanley
identified the magnetic phase as Fe and Ni-rich clusters. Such clusters have a relatively high
Curie temperature (800 K for equiatomic Fe and Ni) and their formation would have so
depleted the austenite matrix of Fe and Ni as to give the observed magnetic behavior.
D. L. Porter and Wood (1979) subjected samples of solution annealed type 304L, and of
both solution-annealed and 20~o cold worked type 316 stainless steels to fast neutron fluences
of up to 7.6 x 1022/cm2 (~40dpa) at temperatures of 400-550°C. A TEM study revealed
extensive ferrite formation in all three steels, with the greatest amount of transformation
occurring at the higher temperatures and greater fluences. The type 304L sample irradiated
at 502°C to 7.6 x 1022/cm2 had formed about 15~ ferrite by volume. An example is shown
in Fig. 67. The ferrite was in blocky particles ~ 1 #m in diameter, a far different micro-
structure than the ~ 1017 fine particles/cm 3 deduced by magnetic measurements by Stanley
(1979) and Stanley and Garr (1974) in similar steels under comparable irradiation treatments.
In type 304L steel the ferrite was observed to nucleate on faulted irradiation induced
dislocation loops.
Mazey et al. (1979, 1980) studied the effect of up to 1.4w/oSi additions on the irradiation
response of a Fe-12w/oCr-15w/oNi alloy which had been implanted with 10 appm He.
Irradiation was with 46.5 MeV Ni 6+ ions to 60 dpa at 1-3 x 10 -3 dpa/s at temperatures of
525, 575 and 625°C. These alloys are somewhat more resistant to ferrite formation than types
304 or 316 stainless steel (see Fig. 55) and showed no ferromagnetic phase prior to irradiation.
Grains of a BCC phase with the 0t-Fe lattice parameter were observed in almost all alloys
after irradiation at 525 or 575°C and in some alloys irradiated at 625°C. The ferrite grains
were the order of 1/~m in diameter; the amount of ferrite increased with Si content, to a
maximum of 50~o in the 1.4w/oSi alloy irradiated at 525°C. The BCC and FCC phases
showed the Kurdjumov-Sachs crystallographic orientation relationship and had the same
composition.
Figure 68 shows electron micrographs and an electron diffraction pattern from this study.
From the orientation relationship it was concluded that the BCC phase was martensite which
PHASE S T A B I L I T Y U N D E R I R R A D I A T I O N 345

FIG. 67. Dark field electron micrograph showing formation of blocky ~, in neutron irradiated type
304L stainless steel. (a) Irradiated to 18 dpa at 499°C, showing transformation interfaoe; (b)
irradiated to 33dpa at 518°C, with diffraction patterns showing orientation relationship between
ferrite and austenite phases. (Electron micrographs courtesy D. Porter.)
346 P R O G R E S S IN MATERIALS SCIENCE

FIG. 68. Fe-15Ni-12Cr-l.4Si irradiated to 60dpa at 525°C with 46MeV Ni + ions. (a) Ferrite
formation from austenite matrix; (b) electron micrographs and diffraction patterns showing relative
orientations of adjacent bcc and fcc grains. (Electron micrographs and diffraction patterns courtesy
D. J. Mazey and Journal of Nuclear Materials.)
PHASE STABILITY UNDER IRRADIATION 347

(a.) UNIRRADIATED E42-10

b,) Ni-IONIRRADIATED AT750°CTO lOOdpa (c.) Ni-ION IRRADIATED AT 70OOC TO 100dpa

d) N~-ION IRRADIATED AT 650°C TO100 dpa (e.) Ni-ION IRRADIATED AT 600OC TO 100dpa

FIG. 69. Dark field electron micrographs showing ?' morphology in alloy E-42 irradiated to 100 dpa
with Ni ions at various temperatures after pre-ageing at 815°C for 10hr. (a) Unirradiated; (b)
irradiated at 750°C; (c) irradiated at 700°C; (d) irradiated at 650°C; (e) irradiated at 600°C. (Electron
micrographs courtesy A. L. Chang and Journal of Nuclear Materials.)
348 PROGRESS IN MATERIALS SCIENCE

I000
Dynam~Scaling ~(t)Rmo,(f):3oC~*Q
)B'C'B Theory)
Thermol Coarsening: ~(t)Rmol(l): T(=
3
o.~ (L-S-W Theory)
800
rt~
I Dynamic Scollng
- ~C-B Theory)
600
E-42-36 ot 5~0"C, 247 x 1022n/cm 2
/
) E-42-36 at 593°C, 2.7 11022 n/cm 2
400 - 0 0 E-40-36 at 593°C, 27 x IO2z n/cm2 --
E
E I E-42-1q ~E
~3000 h -42-10 tit 600°C Ni ions I00 dpo I
W °°c - \ /
20(] -\--The,.°, -...~ ,~ ~ ° , ,~o'c ,, o,, ,~,,Po-1
\ Co=.eninq ~ /
.. jooo h. = ,oo -'--1
I-- I t I
1.0 2.0 3.0 4.0
Supersoturofion zx(l)in otomic fraction (xlO 2)

FIG. 70. Experimental verification of dynamic scaling theory on 7' redistribution under irradiation.
Alloy compositions are given in Table 16. (After Chang and Baron, 1979.)
PHASE STABILITY U N D E R I R R A D I A T I O N 349

had formed on cooling from the irradiation temperature. Martensite formation was attributed
to Ni loss to M23C6 and 7' precipitates.
Boothby and Williams (1981) irradiated a 12w/oCr-15w/oNi-0.44w/oSi steel at 600°C with
46.5 MeV Ni 6+ ions to 60 dpa. Like Mazey et al., they observed large amounts of ~'
martensite and concluded on the basis of Kurdjumov-Sachs and Nishiyama-Wassermann
orientation relationships that the martensite formed on cooling.
The various studies (see Table 14) have shown that at least two types of ferromagnetic
particles formed in irradiated austenitic stainless steels. One is the high density of extremely
fine particles thought by Stanley (1979) to be Fe-Ni clusters, the other is the blocky
~ l-/am-diameter grains ferrite which have been observed in the TEM. Mazey et al. (1979)
presented evidence that the ferrite phase formed martensitically during cooling, while other
investigators concluded that a diffusional transformation occurred during irradiation.
Brager and Garner (1981b) studied the irradiation response of type 316 stainless steel
modified with up to 2w/oSi. Irradiation in EBR-II was to 2-10 x 1022n/cm 2 (E > 0.1 MeV),
or 10-50 dpa, at temperatures ranging from 400 to 600°C. Up to 50~o or more of the matrix

Table 14. Ferrite formation in austenitic stainless steel


Alloy
(Compositions in w/o) Treatment Result References
Type 321 stainless 400, 500°C, 0.75 MeV protons, Copious ~ formation Keefer et al. (1971)
steel 1.8 × 10 4 to 7.2 × 10-4 in peak damage re-
dpa/s, 1-50 dpa gion at 500°C, much
less at 400°C.
Type 316 stainless 500, 600°C, 0.75-1 MeV Ferrite formed at Keefer et al. (1972)
Steel protons, 1.4 × 10-4 to 600°C, not at 500°C.
12 × 10-4dpa/s, 20dpa
Type 316 stainless 500°C, ~ 1 MeV protons, Significant ferrite Keefer and Pard
steel 24dpa, 1.2 × 10-3 formation. (1972, 1973)
dpa/s, incremental He
to 3 appm
Type 347 stainless T < 100°C, neutrons, up to Small amount of fer- Reynolds et aL
steel, CW up to 50% 2 dpa rite formed at higher (1955)
dpa and levels of
CW.
Type 316L stainless steel 400-500°C, up to 7 × 1022 Ferrite forms, max- J. L. Baron et al.
n/cm2, ~ 35 dpa; imum at 490°C, (1974)
4 x 1022n/cm2
Type 316 stainless steel, 425°C, 1.8 x 1022n/cm 2 "] Many fine super- Stanley and Garr
various pre-irradiation (E > 0.1 MeV) ~ 9 dpa; paramagnetic par- (1975)
treatments 500 & 600°C, 3.5 × 1022 ticles
n/cm 2 (E > 0.1 MeV),
18 dpa
SA Type 316 stainless steel 698-890 K, 1-4.5 x 1022 ~3 x 1017/cm3 mag- Stanley (1979)
n/cm2 (E > 0.1 MeV)5-22 netic particles at
dpa higher fluences
SA Type 304L stainless 400-550°C, up to 7.6 x Ferrite formation, D. L. Porter and
steel; 20% CW, SA 1022 fast n/cm 2, ~ 4 0 d p a greatest at higher Wood (1979)
Type 316 stainless fluences and tem-
steel peratures.
20~ CW Fe-12Cr-15Ni 525--625°C, 46.5 MeV Ni 6+, Up to 50~ ct formed Mazey et al. (1979)
plus up to 1.4Si 60 dpa, 1-3 x 10-3 dpa/s martensitically.
12Cr-I 3Ni~.44Si stainless 600°C, 46.5 MeV Ni 6+, Over 50~ ct formed Boothby and Will-
steel 60 dpa martensitically. iams (1981)
Type 316 stainless steel 400-600°C, EBR-II, Up to 50% ferrite Brager and Garner
plus up to 2w/o Si 2-10 x 1022n/cm2 formed in high Si (1981b)
(E > 0.1 MeV) or alloys.
10-50 dpa
350 PROGRESS IN MATERIALS SCIENCE

Table 15. Miscellaneous observations in austenitic steels


Alloy
(compositions in w/o
unless otherwise noted) Treatment Result References
Fe-50a/oNi 250°C, up to 10~8n/cm% Anisotropyof ordering Chamberod et al. (1971)
~5 x 10-4dpa in
magnetic field
Fe-26.4, 29, and 30Ni 80-183°C,up to 0.8 × Ms depressed by up to Porter and Dienes (1959)
101Sn/cm2 (E > 1 MeV) 14.6°C (fine grains) or
~4 x 10-4dpa 34°C (coarse grains)
Fe-7AI-1.8C 50-70°C, 4 × 109R(106C/kg) K phase partly loses co- Kurdyumov et al. (1978)
from 6°Co herency, less tetrago-
nality of martensite.
Fe-Cr-Ni based alloys 510-750°C, various Enhanced matrix solute Chang and Baron (1979)
neutron and Ni ion concentration.
irradiations
Fe-35.4Ni-7.5Cr 593°C, 7.6 × 1022n/cm2 SpinodaMike com- Brager and Garner (1983c)
(E > 0.1 MeV), ~ 38 dpa position modulations of
~400 nm wavelength.

was found to have transformed to ferrite in the 2~o Si alloy. The transformation occurred
because of segregation of Ni and Si, which became so depleted in regions of the sample that
the 7--,~ transformation occurred.

13.5. M i s c e l l a n e o u s O b s e r v a t i o n s
The Fe-Ni system shows ordered FeNi 3 and FeNi phases. Chamberod et al. (1971)
subjected an equiatomic Fe-Ni alloy to neutron irradiation at 250°C in the presence of a
magnetic field. They gave the critical temperature of the Fe-Ni phase in their alloy as 320°C.
Irradiation to up to 10~8n/cm z produced marked magnetic anisotropy, due to the
development of the preferred variant of the ordered phase. The effect of irradiation in
increasing the ordering was interpreted as nucleation of ordered regions on the residues of
displacement cascades.
L. F. Porter and Dienes (1959) studied the effect of neutron irradiation on alloys of 26.4,
29, and 30w/oNi. Irradiation was at 80, 107 and 183°C, to up to 0.8 x 10~8n/cm 2
(E > 1 MeV) or approximately 4 x 10 -4 dpa. The martensite start (Ms) temperature for these
alloys in the absence of irradiation was in the vicinity of 0°C. Alloys with either fine (0.01 mm)
or coarse (0.1 mm) diameter grains were irradiated.
Irradiation lowered M s for the fine-grained samples by up to 14.6°C, and of course-grained
samples by up to 34°C. A lowered M s occurred even in samples which had been decarburized
prior to irradiation to avoid any possible thermal stabilization of the austenite.
The lowering of Ms was attributed to irradiation strengthening of the austenite which
would make the subsequent shear transformation to martensite more difficult. Such an
explanation is consistent with accepted characteristics of both irradiation hardening and the
martensite transformation.
Kurdjumov e t al. (1978) irradiated an austenitic steel of 7w/oA1 and 1.8w/oC with 7-rays
from a 6°Co source to a dose of 4 x 109 R(106C/kg). Irradiation was at 50-70°C; subsequent
transformation was at 78 K.
Irradiation was found to reduce the coherency of the K-phase (Fe3AIC) with the matrix
and to reduce the tetragonality of the martensite, which formed at 78 K. It was concluded
PHASE STABILITY UNDER IRRADIATION 351

Table 16. Compositionsof alloysinvestigatedby Chang and Baron (1979)


Alloy Composition (wt%)
E48 Fe-14.6Cr-25.0Ni-3.04Ti-l.51A1-0.006C
E40 Fe-14.9Cr-34.8,Ni-l.98Ti-l.50A1--0.004C
E54 Fe- 14.5Cr-45.7Ni-1.60Ti-1.05A1-0.001C
E42 Fe-I 5.ICr-34.6Ni-l.99Ti-l.70AI-0.04C-0.69Si-0.015B
A286 Fe-15.01Cr-25.51Ni-2.16Ti-0.18A1-0.05C-0.55Si-l.4Mn-0.03V

that the reduction in coherency strains around the K-phase particles caused the lower
tetragonality of the martensite. The irradiation-induced loss of coherency is consistent with
results on Cu-Co discussed in Section 11.4. The lowered martensite tetragonality could
indeed be due to the coherency strains of the K-phase no longer being present to
accommodate the transformation strains of the martensite.
M. Baron et al. (1977) and M. Baron (1981) developed a theory for particle coarsening
under the influence of irradiation-induced disordering (Section 9). The theory predicted a
maximum particle radius, rmax(O0), in the presence of irradiation. Prior to approaching this
equilibrium state, a relationship between maximum particle size and matrix supersaturation
was predicted, such that:

rmax(t)A(t)= ~ (13.1)
where rmax(t)=maximum particle size at time t; A(t)=difference between actual and
equilibrium solute concentrations in matrix, in atomic fraction; and ct = a constant the otder
of the lattice parameter. Under thermal conditions, mean particle radius, f, and super-
saturation are related by the familiar Gibbs-Thomson equation:
2yI)
F = kT In C/Ce' (13.2)
where C and Ce are actual and equilibrium solute concentrations, respectively.
Chang and M. Baron (1979) studied thermal and irradiation-enhanced coarsening of the
~,' phase in a series of experimental and commercial Fe-based alloys. Table 16 shows the alloys
to be based on the Fe-Ni-Cr ternary.
Figure 69 shows the evolution of the 7' phase after irradiation to 100 dpa at various
temperatures.
Figure 70 summarizes the results of the study. The matrix concentration is seen to obey
Eq. (13.2) under thermal conditions and Eq. (13.1) under a range of neutron and Ni-ion
irradiation conditions. Agreement between dynamic scaling theory and experiment is
remarkably good in view of the theory taking no account of irradiation-induced solute
segregation, which is known to be important in this class of alloy. The dissolution parameter
(Section 9.1) was set at 2.74 x 10-9cm to give the best fit between theory and experiment.
Brager and Garner (1983c, 1984) irradiated a Fe-35.5w/oNi-7.5w/oCr alloy to
7.6 x 1022n/cm2 (E > 0.1 MeV) or ~ 3 8 d p a at 593°C. After irradiation the alloy showed
spinodal-like composition modulations with a wavelength of approximately 200 nm. The
modulations were attributed to irradiation accelerated spinodal decomposition.

14. NICKEL-BASEDALLOYS*
Nickel-based alloys often have attractive strength and corrosion resistance properties, and
also tend to be resistant to irradiation-induced void swelling, all of which make them
*See also notes in proof (pp. 430-434).
352 PROGRESS IN MATERIALS SCIENCE

30

20
~° /,~.~.-___~.~___o~: ~., _

._o 10

:[~ I */,S,
5o0 ,ooo ~soo 2000
Depth (,~)

FIG. 71. Peak-to-peak ratio of solute-to-Ni lines as a function of depth, obtained by Auger
spectroscopy of irradiated Ni-based alloys. The samples were irradiated as follows: Ni-Si, 8.5 dpa
at 560°C; Ni-A1, 10.7dpa at 620°C; Ni-Ti, 8.5dpa at 575°C; Ni-Mo, ll.2dpa at 617°C. (After
Potter et al., 1977.)

attractive for fast reactor and controlled fusion reactor applications and have led to numerous
studies of irradiation effects. As in the case of austenitic steels, irradiation-induced solute
segregation plays a key role in irradiation altered phase stability and will be discussed first.
Table 17 gives the compositions of some commercial Ni-based alloys discussed in this section.

14.1. Solute Segregation


Potter et al. (1977) studied solute segregation under 3.5 MeV Ni ÷ ion irradiation in alloys
of 1 a/o A1, Be, Mo, Si, or Ti. After irradiation at temperatures between 560 and 690°C, Auger
electron spectroscopy (Fig. 71) showed the oversized AI, Mo, and Ti atoms to be segregated
away from the surface of the foil, and the undersized Si and Be to be enriched at the surface,
in agreement with the misfit theory of solute segregation under irradiation.
Segregation of A1 from the surface was illustrated in a 1 MeV Ni ÷ ion irradiation
experiment by Potter (1978) on a pre-thinned 1000,~ foil. After irradiation to 4.8 dpa at
600°C a layer in the center of the foil was highly enriched in 7', Ni3A1 precipitates which
formed from A1 rejected from the surface.

2
/,"--,AI Ti
/ ,,r',~Mn ' ~
;/ ' ,
.g=

/// /,, ~ i .

3
i
Ii • l •

-I t I , l ,
300 400 500 600
Irradiation Temperature (*C)
FIG. 72. The segregation parameter plotted against irradiation temperature in a Ni-based alloy. Note
that while AI, Mn and Cr have the same relative ranking at all temperatures, Ti moves from having
a dip in solute concentration below 400°C to the strongest enhancement in the defect-rich region
for irradiation at 600°C. (After Marwick et al., 1979.)
PHASE STABILITY UNDER IRRADIATION 353

Rehn et al. (1978) confirmed the results of Potter et al. (1977) for Ni alloys of 1 a/o At,
Mo, Si, or Ti irradiated to 10 dpa in the 400-650°C temperature range.
Marwick et al. (1979) irradiated Ni-0.33a/oSi and an alloy with small amounts of A1, Cr,
Mn, Si, and Ti with 75 keV Ni ÷ ions, which penetrate only about 200/~ into the foil. Figure
72 shows some of the results of their study. At 500°C, Si segregated toward the foil surface
while all the other solutes moved toward the interior, as predicted.
However, at temperatures near 25°C, where vacancies and divacancies are immobile and
all segregation is due to interstitial migration, all elements segregated to the surface. Marwick
et al. suggested that segregation of the oversized AI, Cr, Mn, and Ti atoms to the surface
is due to a weak interaction between the solute and the interstitial. The authors conclude that
at higher temperatures the AI, Cr, Mn, and Ti atoms segregate from the surface by an inverse
Kirkendall mechanism, which results from their being fast diffusers in Ni. Thus, either size
misfit or the inverse Kirkendall effect serves to explain the same results. It is at present unclear
which mechanism has primary responsibility for determining the direction of segregation.
Rehn et al. (1981) irradiated alloys of 10 a/o and 60 a/o Cu with 2.2 MeV Ar ÷ ions at a
peak damage rate of 3 x 1 0 - 3 dpa/s. Irradiation was to a maximum surface dose of 5 dpa at
500°C. The oversized Cu atoms were found to move from the sample surface to the interior
as predicted by the size misfit theory.
Takahashi et al. (1981) irradiated a Ni-2a/oCu alloy in a 650 kV HVEM at a displacement
rate of ~5 x 10-4dpa/s. After irradiation to 5dpa at 623 K, Cu was depleted near void
surfaces.
Brimhall et al. (1981) irradiated the Nimonic alloy PE-16 (Table 17) with 5 MeV Ni 2÷ ions
at a displacement rate of 3 x 10 -3 dpa/s. After 2 dpa at 775 K, phosphorous was strongly
segregated to the free surface.
Yang (1982) irradiated the Nimonic alloy PE-16 to 6-10 x 1022n/cm ~ (E > 0.1 MeV) or
about 30-50 dpa in the temperature range 425-650°C. At temperatures above 510°C, a 300 ,~
thick layer of 7', enriched in Ni, Ti, AI, and Si, formed at grain boundaries. It was not clear
which of these elements arrived at the boundary due to irradiation-induced solute segregation
and which simply diffused down a solute concentration gradient to precipitate in the ),' phase.

Table 17. C o m p o s i t i o n s o f s o m e n i c k e l - b a s e d a l l o y s
Nimonic Alloy* Inconel 625t Inconel 706t
P E - 16 (nominal (nominal
Element (in w t % ) c o m p o s i t i o n , %) composition, ~)
C 0.10 m a x 0.05 0.03
Si 0.3 m a x 0.2 0.2
Mn 0.2 m a x 0.2 0.2
Cr 15.0-18.0 21.5 16
Ti 0.9-1.5 0.2 1.8
AI 0.9-1.5 0.2 0.2
Co 2.0 m a x -- --
Mo 2.5--4.0 9.0 --
B 0.005 m a x -- --
Ni -- 61.0 41.5
Nb -- 3.6 2.9
Zr 0.05 m a x -- --
S 0.015 m a x -- --
Ni + Co 42.0-45.0 -- --
Fe balance -- --
* F r o m Publication 3349, January 1968, H e n r y W i g g i n & C o . , L t d .
t F r o m Metal Progress 1982 Materials and Processing Databook, A m e r -
i c a n Society f o r M e t a l s , M e t a l s P a r k , O H (1982).

J . P . M S 28,3-4 G
354 P R O G R E S S IN MATERIALS SCIENCE

Marwick et al. (1978) irradiated solution annealed samples of PE-16 with 46.5 MeV Ni +
ions to 60 dpa at 590 and 650°C. Scanning transmission electron microscopy showed the
nickel concentration at the void:matrix interface to be enriched by over 54~ over the
background concentration. The authors did not state whether or not the void were decorated
with 7' particles, as may happen in this alloy. As such, it is uncertain whether the Ni was
present in solid solution in the matrix or in 7' precipitates.

14.2 N i - A I Alloys
Nickel will dissolve ca. 20a/oAl at high temperatures, which on cooling precipitates as the
coherent, ordered, 7' phase. The so-called 7-7' alloys are prototypical of such more complex
Ni-based engineering alloys as Nimonic PE-16 and the Inconel alloys. In that the high
temperature strength of these alloys requires maintenance of a stable particle size distribution,
the coarsening behavior of the 7' phase is of major interest. Use of these materials in a nuclear
environment requires knowledge of the effects of irradiation on particle stability, whether
enhanced diffusion, solute segregation, disorder-dissolution, or some other mechanisms are
operative.
Nelson et al. (1972) studied the behavior of 7' precipitates in a 13.5a/oA1 alloy under
100 keV Ni + ion irradiation at 10 -2 dpa/s. During 550°C irradiation the 7' particles, originally
250/~ in diameter, took on a fragmented appearance and finally, by 1.1 × 1017 ions/cm2,
(~ 100 dpa) were replaced by a new fine particle distribution. Figure 73 illustrates the inverse
coarsening of the irradiated alloy. This result was attributed to the disorder-dissolution model
discussed in Section 8. Irradiation of this alloy to 0.1 dpa at 25°C disordered the 7' particles.
Ro and Mitchell (1978) irradiated Ni-13a/oA1 and 16a/oAl alloys in the 650 kV HVEM
at temperatures of 300-750°C. Displacement rates were calculated as being between
2.26 x 10-3/s and 5.77 x 10-4/s depending on crystal orientation; maximum damage levels
were a few dpa. Figure 74 shows the effects of high flux irradiation at 750°C on a specimen
of Ni-8w/oA1 pre-aged at 650°C for 2 hr.
Ro and Mitchell found, rather surprisingly, the same ? 3 versus t rate law observed in
thermal coarsening. Figures 75 and 76 show that coarsening was greatly accelerated by
irradiation below about 700°C, an enhancement attributed to an irradiation-enhanced
diffusion coefficient. The enhanced diffusion coefficient was calculated from coarsening data
to be approximately 5 x 10-~Scm2/s between 300 and 650°C, which is consistent with
estimates made in Section 2.7.
Potter and Hoff (1976) and Kirchner et al. (1976) obtained different results than did Ro
and Mitchell in Ni + ion irradiation of similar alloys. Potter and Hoff irradiated Ni-13a/oAl
alloys with 3.2 MeV Ni + ions at a displacement rate of 2.7 × 10-3/s. The starting particle size
was ~ 3 5 A . Coarsening occurred at low doses, followed by fragmentation into smaller
particles as irradiation proceeded to ,-~20 dpa.
Kirchner et al. obtained results similar to those of Potter and Hoff in irradiation of a
Ni-14a/oA1 alloy with 2.8MeVNi + ions at the relatively high displacement rate of
4.4 x 10-2 dpa/s. Irradiation to 20 dpa at 725°C reduced a 400-A average particle size to 85/~;
at lower temperatures, precipitates appeared fractured, and a fine distribution of new
precipitates formed.
Potter and Ryding (1977) and Potter and McCormick (1979) showed that Ni + ion-
irradiated Ni-A1 alloys show irradiation-enhanced coarsening up to about 20 dpa. After that
damage level, larger particles dissolved and a fine distribution of new particles was nucleated,
giving a decrease in average particle size, as shown in Fig. 77.
PHASE STABILITY U N D E R I R R A D I A T I O N 355

uni rrad i a t e d
1.6x 1016cm" 2

4"8 x 101acm"2 1"1 x ld7cm "2

0"Ip
FIG. 73. Dark field electron micrographs of changes in y' precipitates in Ni-13.5a/oAl under 100 kV
Ni ÷ ion irradiation at 550°C to a maximum displacement level of ~ 100 dpa. (Electron micrographs
courtesy D. J. Mazey and Journal of Nuclear Materials.)
356 P R O G R E S S IN MATERIALS SCIENCE

000 200 IToy' ~/2~'0


\\/

(a) (b)

(c) (d) '


OI~rn
FIG. 74. Effects of 650 kV e-, 2.26 x 10-3dpa/s irradiation at 750°C on Ni-8w/oAl. (a) 0dpa; (b)
0.4dpa; (c) 1.4 dpa; (d) 3.5 dpa. Dark field electron micrographs imaged with (lI0) 7' superlattice
reflection, foil normal [001], ~ = [1T0]. (Electron micrographs courtesy D. H. Ro.)
PHASE STABILITY UNDER IRRADIATION 357

38 x I0~

35
/
.26x 10-3 dpo/s
~ 3o e / a t 750%

~25

~2o
/
'.26 xlO'3 dpa/s
at 650*Cj,~e / 2.26xlO-3dpaJ
.,~.e/ at 500%
~|/ Aged/h s#u at 650%
o rl i = ~ I I
30 60 90 120 150 180 210
Irrodiotion Time (minutes)

FIG. 75. Electron irradiation enhanced coarsening o f y' particles in 650 kV e - irradiated Ni-8w/oA1.
(After Ro and Mitchell, 1978.)

The inverse coarsening was attributed to segregation of oversized A1 atoms away from
dislocation loops, which serve as point defect sinks. The Ni-rich areas become biconvex lenses
up to ~ 4000/~ in diameter, denuded of Ni3AI precipitates. Finally, A1 diffused back into the
center of the lenses, away from the defect sink at the circumference, and new y' particles were
nucleated.
Butler and Orchard (1981) irradiated a series of fully ordered alloys, including Ni3A1, in
the 1 MV H V E M at temperatures between 10 and 300 K where the vacancy is immobile. The
results of the experiments were described in terms of:
S = Sac-A%
where S and So are actual and initial order parameters, ~bt is fluence, and A is the effective
disordering cross section. The disordering cross section, A, was found to be dependent on

~. Rod- Enhanced •
Coarsening
IO-P6
k Thermal Coarsening •

A
10-=7
to

.~ io-le
i

I0"re

10-m

iO-Z I

1.0
,

1.2 1.4
,\ 1.6
I I

1.8
1/T x I000 (°K) -I
FIG. 76. Arrhenius plot or irradiation enhanced rate constant in 650 kV e - irradiated Ni-8w/oAl.
k = rate constant, T = temperature, Ce = equilibrium weight fraction o f A1. (After Ro and Mitchell,
1978.)
358 PROGRESS IN M A T E R I A L S SCIENCE

A
4
O o%%%
3

-: 2 ~ o~
i0ted
,
%\\
\\
',,\

2/0 Conventional Aging_ ~o


I I I ' ~ I
2 4 6 8 10 12
Time (103s)
FIG. 77. Cube diameter of 7' vs. Ni ÷ irradiation time at 650°C, compared to conventional aging
kinetics. Particle size in the irradiated Ni-6.35w/oAl alloy starts to decrease after about 20dpa.
(After Potter and Ryding, 1977.)

temperature, accelerating voltage, and crystal orientation. At 300 K and 1 MV, a (100)
oriented Ni3A1 crystal had A = 5.1 x 10 -27 m 2, which implied 1.1 atoms disordered per atom
displaced. The authors concluded that disordering was due to replacement collision se-
quences. Figure 78 shows the results of 1 MV irradiation at two different temperatures.
Liu et al. (1981) irradiated fully ordered Ni3A1 and NiAI alloys in a 650 kV HVEM. Below
190 K a dose of 0.9 dpa fully disordered the Ni3A1. From 200 to 240 K, irradiation of Ni3A1
to 2dpa produced a residual order of S = 0.1-0.15, which was attributed to enhanced
reordering by interstitial dumbbells. Irradiation of initially fully ordered NiA1 to 6.6 dpa at
150 K left a residual order of S ~ 0.5. This unexpected residual order was interpreted in terms
of a high ordering energy in NiAI which induced point defect recombinations which tended
to reorder the alloy.
Liu et al. found no orientation dependence of disordering rate at 150 K, in conflict with
the observations of Butler and Orchard. The lack of orientation dependence caused them to
attribute disordering to uncorrelated "wrong" recombination of vacancies and interstitials.

14.3. N i - B e Alloys
Beryllium solubility in nickel is 15.3 a/o at the 1157°C eutectic temperature, and decreases
to approximately 5 a/o at 650°C (Hansen, 1958). The precipitation sequence in Ni-Be alloys

Ni3AI , 1MV e-,<ll0>


£n S IOK

-I.0 __

/o/ .j..1.
o 300 K - -

-0.5
f.//
o~/ 10
I I
2.0
I
3.0
Fluence x IO~e-/cm 2

FIG. 78. HVEM irradiation induced disordering of [110] orriented Ni~A1 as a function of temperature.
(After Butler and Orchard, 1981.)
PHASE STABILITY UNDER IRRADIATION 359

under thermal conditions is (Merrick, 1978):

solid solution~GP zones~/3'~fl


The GP zones are monolayers of Be atoms that form on { 100} matrix planes. The metastable
fl' phase has the body centered tetragonal structure and is also coherent. The equilibrium
equiatomic NiBe/3 phase is CsCI type.
Okamoto et al. (1975) subjected alloys containing 0.1, 0.3, and 0.7 a/o Be to 3.2 MeV Ni ÷
ions at 425, 525, and 625°C. Dose levels of 1-30 dpa were obtained at a displacement rate
of 3 x 10-3/s.
Precipitates formed during irradiation at all temperatures, in spite of the alloys being
thermally single phase. The precipitate phase was not identified. Precipitation was attributed
to irradiation-induced segregation of the undersized Be solute to point defect sinks until the
concentration exceeded the solubility limit. Beryllium has been observed to segregate to defect
sinks in irradiated alloys (Table 18).
Mukai and Mitchell (1982) subjected a Ni-la/oBe alloy to irradiation at 3.7 x 10 -4 t o
9.3 x 10 -4 dpa/s in a 650 kV HVEM. Irradiation was at temperatures of 390-610°C, all of
which are above the 360°C solvus temperature of the alloy.
Irradiation at 390°C to 2.2 dpa at the higher displacement rate gave plate-like 13' and/or
GP zones on {I00} planes, as seen in Fig. 79. Precipitation was homogeneous and no
dislocation loops were observed. At 470°C at the low displacement rates homogeneous
nucleation of/3' particles occurred after less than 0.02 dpa, a remarkably low damage level
to produce precipitation in an undersaturated alloy. Figure 80 shows that further irradiation
at 470°C produced platelets of the equilibrium/3 phase on {211} planes. Figure 81 shows the
effects of progressive irradiation at 470°C on the precipitate microstructure of Ni-la/oBe. At
610°C precipitation was less obvious and irradiation produced mostly dislocation loops.
Mukai and Mitchell proposed a precipitation mechanism whereby undersized solute
clusters attracted interstitial dumbbells, which contained a high percentage of undersized Be
atoms. This mechanism is something of a restatement of that for coherent precipitation in
undersaturated solutions discussed in Section 4.2.
Kernohan et al. (1956) irradiated a Ni-2.5a/oBe alloy to 4 x 1017 fast n/cm z ( ~ 2 x 10 . 4
dpa) at 300°C. Measurements of the ferromagnetic Curie temperature showed that 1.3a/o
more of the Be precipitated with irradiation than during a thermal anneal of the same
duration. The authors assumed that precipitates were of the equilibrium/3 phase.
The enhanced precipitation was tentatively attributed to irradiation enhanced diffusion.
The 2 × 10 -4 dpa were sufficient to create a vacancy concentration far in excess of the thermal
value, so enhanced diffusion is a feasible mechanism. It is, however, also possible that
displacement cascades or interstitial loops provided nucleation sites for the semicoherent fl
phase.
Grinchuk and Loboda (1974) irradiated Ni-2w/oBe to fast neutron fluences from
4.6 x 1019/cm2 to 1.2 x 102l/cm2 (E >0.8 MeV) or --~0.02-0.5dpa at 70°C. Precipitation
during post-irradiation annealing was then studied by x-ray, resistivity, and microhardness
measurements and by TEM. GP zone formation in the 400-500°C range was found to be
significantly accelerated by irradiation, with the effect increasing with neutron fluence. The
acceleration was attributed to irradiation-enhanced diffusion. The lifetime of the irradiation-
produced vacancies was thought to be prolonged by formation of multi-defect complexes so
as to maintain a vacancy supersaturation at the relatively high annealing temperatures, since
free vacancies would escape very rapidly.
360 P R O G R E S S IN M A T E R I A L S SCIENCE

Table 18. Solute segregation in nickel-based alloys


Alloy Treatment Result References
Ni-la/oAl 3.5 MeV Ni ÷, 2.5 x 10 -3 AI segregated from Potter et al. (1977)
dpa/s, 620°C, 10.7 dpa surface
Ni-la/oBe Same as above except Be segregated to void Potter et al. (1977)
690°C, 1.5 dpa surfaces.
Ni-la/oMo Same as above except Mo segregated from Potter et al. (1977)
617°C, 11.2 dpa surface
Ni-la/oSi Same as above except Si segregated to sur- Potter et al. (1977)
560°C, 8.5 dpa face
Ni-la/oTi Same as above except Ti segregated from Potter et aL (1977)
575°C, 8.5 dpa surface.
Ni-7.1w/oAl 600°C, 1 MeV Ni ÷, 3.4 Dense layer of ?' par- Potter (1978)
x 10-4dpa/s, 4.8 dpa ticles in center of thin
foil.
Ni-la/oAl 3.5 MeV Ni ÷, 2.5 x 10 3 A1 segregated from Rehn et al. (1978)
dpa/s, 10.3dpa, 510°C surface.
and 10.7dpa, 620°C
Ni-la/oMo Same as above except Mo segregated from Rehn et al. (1978)
11.6 dpa, 530°C, surface.
and ll.2dpa, 615°C
Ni-la/oSi Same as above except Si segregated to sur- Rehn et al. (1978)
5 dpa, 385°C, face.
3.6 dpa, 480°C,
4.0 dpa, 530°C,
8.5 dpa, 560°C,
3.9 dpa, 600°C,
4.4 dpa, 660°C,

Ni-la/oTi Same as above except Ti segregated from Rehn et al. (1978)


11.2 dpa, 515°C, surface.
8.5 dpa, 575°C,

Ni + small amounts of 74 keV Ni ÷, 1.6 × 10~6/cm2, AI, Cr, Mn, Ti segre- Marwick et al. (1979)
AI, Cr, Mn, Si and Ti 500°C, ~ 15 dpa gated from surface, Si
segregated to surface
Ni + small amounts of Same as above except AI, Cr, Mn, Si, Ti all Marwick et al., (1979)
AI, Cr, Mn, Si, and Ti ~25°C segregated toward sur-
face.
Ni-0.33a/oSi Same as above except Si segregated to sur- Marwick et al. (1979)
500°C, ~ 15 dpa face.
Ni-10a/oCu 2.2MeV Ar ÷, 3 x 10 -3 Cu segregated from Rehn et al. (1981)
dpa/s, 5dpa, 500°C surface.
Ni-60a/oCu Same as above. Cu segregated from Rehn et al. (1981)
surface.
Ni-2a/oCu 623 K, 650 kV HVEM, 5 Cu segregated from Takahashi et al.
dpa, ~5 x 10 4dpa/s voids. (1981)
Nimonic PE-16 775 K, 5 MeV Ni 2÷, 3 P segregated to free Brimhall et al. (1981)
x 10 3dpa/s, 2 dpa surface
Nimonic PE-16 425-650°C, 6-10 x 1022 Ni, Ti, AI, and Si seg- Yang (1982)
n/cm 2 (E > 1 MeV) regated to 7" at grain
~ 30-50 dpa boundaries.
Nimonic PE-16 590 and 650°C, 46.5 Ni concentration near Marwick et al. (1978)
MeV Ni ÷, 60dpa voids enhanced by
over 54%.

G r i n c h u k a n d K i r s a n o v (1974) e x p o s e d a 2 w / o B e a l l o y w h i c h c o n t a i n e d G P z o n e s to fast
n e u t r o n f l u e n c e s f r o m 7 x 1016/cm 2 to 1.2 x 1021/cm 2 ( ~ 4 x 1 0 - 5 - 0 . 5 d p a ) at 5 0 - 7 0 ° C . T h e
lattice p a r a m e t e r o f the N i - r i c h m a t r i x w a s f o u n d to d e c r e a s e p r o g r e s s i v e l y d u r i n g i r r a d i a t i o n .
T h i s d e c r e a s e w a s i n t e r p r e t e d to m e a n t h a t at t h e h i g h e s t d a m a g e level o v e r 90~o o f the G P
PHASE S T A B I L I T Y UNDER IRRADIATION 361

Fie;. 79. Electron micrograph of/~' GP zones and/or precipitates which formed in Ni-la/oBe after
irradiation at 390°C in the 650KV HVEM. ~ = [1TI], Z---[011]. (Electron micrograph courtesy T.
Mukai, T. E. Mitchell and Journal of Nuclear Materials.)
362 P R O G R E S S IN MATERIALS SCIENCE

FIG. 80. Electron micrograph of ~' images in Ni-la/oBe after 5min HVEM irradiation at
5.5 × 10-4dpa/s. ~ = [200], ~ = [001]. (Electron micrograph courtesy T. Mukai, T. E. Mitchell and
Journal of Nuclear Materials.)
,<

S
Z
7

O.Ipm

FIG. 81. Electron micrographs showing microstructural changes in Ni-la/oBe irradiated in the HVEM at 470°C and 3.7 x 10 4 dpa/s for various
periods of time. (a) 0 min; (b) 1 min; (c) 4 min. The 1 min irradiation induces homogeneous precipitation of fl'; plate-like fl' forms along [2"i0]
directions after 4min. (Electron micrographs courtesy T. Mukai, T. E. Mitchell and Journal of Nuclear Materials.)
364 PROGRESS IN MATERIALS SCIENCE

10-3 p 0 0 °O °==
I

oeeoo
t
1

lff* • 7"'Present
o ),' Absent
I , I = I = I , I
0 200 40O 600 800
T, °C
FIG. 82. Displacement rate vs temperature precipitation diagram for a Ni-6a/oGe alloy under 1 MV
H V E M irradiation. (After Barbu, 1980.)
PHASE STABILITY UNDER IRRADIATION 365

zones had been dispersed back into solid solution in the matrix. Even at 10 -4 dpa, some 30~o
of the GP zones had been returned to solid solution. Disappearance of the GP zones was
verified by TEM studies of the irradiated alloys
Dissolution of the GP zones was attributed to substitution of Ni for Be in the monolayer-
thick zones during the passage of dynamic crowdions. A computer modeling study supported
this mechanism.

14.4. N i - C Alloys
Shriver and Wuttig (1972) exposed a Ni-0.3w/oC alloy to fast neutron fluences of
5 x 1013-2 × 1017n/cm 2 ( ~ 2 × 10-8-10 -4 dpa) at 40°C. Carbon clustering and precipitation
during post-irradiation annealing were followed by magnetic measurements, the authors
concluded from permeability and disaccommodation measurements that up to 40~ of the
carbon left solid solution during irradiation, but returned during annealing between 150 and
200°C. Loss of carbon was attributed to interaction with mobile defects, which at 40°C would
be self intersitials. Above 200°C irradiation-enhanced decomposition of the solid solution was
interpreted as due to the provision of additional precipitate nucleation sites.

14.5. N i - C u Alloys*
Hansen (1958) shows complete solid solubility in the Ni-Cu system, although a low
temperature miscibility gap has been predicted (Mozer et al., 1968).
Wagner et al. (1980, 1982) used diffuse and small-angle neutron diffraction to study an
electron-irradiated Ni-41a/oCu alloy. The alloy was irradiated with up to 1.6 x 102°, 2-3 MV
e-/cm 2 ( ~ 1.6 × 10-2 dpa) at temperatures between 373 and 640 K.
Wagner et al. found a periodic solid solution decomposition with a wavelength of about
45 ~ between 373 and 480 K. The decomposition kinetics agreed with Cook's (1970) model
for short range clustering and spinodal decomposition within a limited momentum transfer
range. The decomposition structure was stable during thermal annealing up to 600 K.
Wagner et al. interpreted their results in terms of irradiation-enhanced diffusion, but also
considered the possibility of an irradiation-induced transformation. The low damage level
( ~ 5 × 10-3 dpa) is unlikely to give an irradiation-induced transformation but could certainly
increase the atomic mobility by orders of magnitude over the thermal values.

14.6. N i - G e Alloys
Barbu (1980) irradiated a Ni-6a/oGe alloy in the 1 MV HVEM at temperatures between
125 and 650°C. The solid solubility of Ge in Ni is only weakly temperature-dependent, being
approximately 12 a/o at 600°C (Hansen, 1958).
Barbu observed precipitation of the equilibrium fl, Ni3Ge phase, which has the Cu3Au
structure, below 600°C at 3.3 × 10 -4 dpa/s and below 550°C at 1.5 x 10 -3 dpa/s. Figure 82
presents these results as a displacement rate versus temperature precipitation diagram. In the
figure precipitation of the 7' phase was observed in region II, but not in regions I and III.
The alloy is thermally single phase at both temperatures, so the precipitation was induced
by irradiation. Germanium is oversized in Ni (D~f= 4.5~o), so irradiation-induced solute
segregation to point defect sinks was probably not responsible for inducing precipitation. The
actual mechanism is, however, unknown.
*See also Cu-Ni alloys.
366 PROGRESS IN MATERIALS SCIENCE

14.7. N i - M o Alloys
The compound Ni4Mo shows long range order (LRO) up to approximately 1130K.
Banerjee et al. (1983, 1984) irradiated Ni4Mo in the 1 MV HVEM at 10-3-10-2dpa/s in the
temperature range 50-1050 K. Irradiations were to a maximum of ~ 2 dpa. Samples were
irradiated in the quenched condition, with only short range order (SRO), and in the aged,
fully ordered condition, with a LRO parameter S --- 1.
Three different responses to irradiation were observed, depending on the temperature:
(i) Low temperatures (50 K < T < 200 K). Irradiation destroyed both LRO and SRO.
Figure 83 shows destruction of LRO at 170 K.
(ii) Intermediate temperatures (200 K < T < 550 K). From 200-400 K, samples initially
ordered showed a complete decay of LRO, followed by development of SRO. Samples
which began with SRO retained it even after prolonged irradiation. At 450-550 K the
final structure depended on the initial structure. Samples initially in the SRO state and
samples initially in the LRO state retained their structure and no transition to the other
structure occurred. Figure 84 shows destruction of LRO and development of SRO
during a 300 K irradiation.
(iii) High temperatures (550 K < T < 1050 K). In this range LRO increases with irradi-
ation dose. Up to 800 K LRO developed continuously and above 800 K occurred by
nucleation and growth. Up to 720 K, the SRO did not disappear completely. Figure
85 illustrates formation of LRO and destruction of most SRO during a 670K
irradiation.
The authors interpreted their results as competition between irradiation-induced dis-
ordering (most effective at low temperatures) and ordering reactions assisted by irradiation-
enhanced diffusion. Good agreement (Banerjee and Urban, 1984) was found between the
measured order parameters and calculated values of the kind shown in Fig. 31.
Carpenter and Kenik (1977) irradiated Ni4Mo with 1 M V e - in the HVEM at
,,,4 x 10 -3 dpa/s to up to 41 dpa at temperatures of 400-900 K. They found that specimens
which initially showed SRO transformed to LRO; the transformation was faster at the higher
temperatures. At 900 K SRO was replaced by LRO after only 2.7 dpa.
The results of Carpenter and Kenik are consistent with those of Banerjee et al. except in
the 400-500°C range, where the latter did not observe the SRO-~LRO transition. Differences
in estimating the temperature in the HVEM could account for much of this discrepancy.

14.8. N i - N b Alloys
Amorphous or microcrystalline alloys are thought to be highly resistant to displacement
damage, and are considered as candidates for use in advanced fusion reactor systems.
The Ni-Nb system shows eutectics at 85 and 60 a/oNi; rapid solidification renders alloys
of the former composition microcrystalline and the latter amorphous (Chernock and Russell,
1984).
Amorphous Ni~,0a/oNb crystallizes at approximately 925 K during thermal annealing.
Rechtin et al. (1978) irradiated amorphous and partly crystalline Ni-40a/oNb alloys with
3.5 MeV Ni 2+ ions at 3 x 10 -3 dpa/s. Irradiation of amorphous samples to 6-20 dpa at 300
or 900 K produced no changes in the microstructure or electron diffraction pattern.
Irradiation of partly crystalline material at 300 or 900 K induced nearly complete reversion
of the Ni-Nb crystallites to the amorphous state after ~ 1-2 dpa and complete reversion after
6 dpa. No voids formed in either material.
PHASE STABILITY UNDER IRRADIATION 367

(a) (b)

Cc) (d)
FIG. 83. Electron diffraction patterns of an initially long-range ordered sample of Ni4Mo taken in
situ during 1 MV electron irradiation in a HVEM at 170 K, at a defect production rate of
3 x 10 -3 dpa/s. Zone axis: [001]. Faint spots in (a) are due to double diffraction. After (a) 0 s; (b)
45 s; (c) 165s; (d) 300s. (Electron diffraction patterns courtesy S. Banerjee, K. Urban and M.
Wilkens.)
368 PROGRESS IN M A T E R I A L S SCIENCE

Ca) Cb)

Co) Cd)
FIG. 84. Destruction of LRO and formation of SRO in Ni4Mo during irradiation at 300K at
5 × 10 -3 dpa/s. Long, range order is destroyed without any evidence for formation of short-range
order. Only after LRO is destroyed virtually completely does SRO form. After: (a) 0 s; (b) 60 s;
(c) 120 s; (d) 240 s. (Electron diffraction patterns courtesy S. Banerjee, K. Urban and M. Wilkens.)
PHASE STABILITY U N D E R I R R A D I A T I O N 369

Ca) Cb)

(c) (d)
FIG. 85. Formation of LRO from SRO at 670 K. After prolonged irradiation the Ni4Mo specimen
reaches a metastable state characterized by simultaneous presence of SRO and LRO (although LRO
is eventually most pronounced there is still some SRO intensity which persists). Only at T > 720 K
does the system proceed to complete LRO. Defect production rate: 5 x 10-3dpa/s. After: (a) 0 s;
(b) 15 s; (c) 90 s; (d) 270 s. (Electron diffraction patterns courtesy S. Banerjee, K. Urban and
M. Wilkins.)

FIG. 86. Electron micrograph of NissNb~5, as rapidly solidified showing internal cellular structure
with fine precipitates the cell boundaries. (Electron micrograph courtesy R. S. Chernock.)
370 PROGRESS IN M A T E R I A L S SCIENCE

FIG. 87. Electron micrograph of rapidly solidified NiasNb15 after Ni 2÷ ion irradiation to 20 dpa at
700 K. (a) Bright field; (b) dark field. (Electron micrograph courtesy R. S. Chernock.)
PHASE STABILITY UNDER IRRADIATION 371

Chernock and Russell (1984) used 5.5 MeV Ni 2+ ions in irradiation of partly crystalline
Ni--40a/oNb and microcrystalline Ni-15a/oNb alloys. Irradiation of the 40 a/o alloy to 2 or
20 dpa at 800 or 900 K halted the evolution of the microstructure which would have occurred
under thermal annealing. The difference between this and the result of Rechtin et al. (1978)
was explained on the basis of Rechtin et al. beginning with a microstructure consisting of
NiNb in an amorphous matrix, whereas the crystalline phase in the Chernock and Russell
samples was identified as Ni3Nb. The two phases clearly responded differently to irradiation.
Irradiation of the 15a/oNb alloy to 20 dpa at 800 K gave partial dissolution of the existing
NiaNb particles and precipitation of an array of small particles surrounding them. The
microstructure suggested that a recoil-resolution process of the type discussed in Section 8
might be operative. Irradiation to 20 dpa at 700 K produced a few small voids, the only
occurrence of void formation in either study.
Figures 86 and 87 show the 15a/oNb alloy as solidified and after irradiation to 20 dpa at
700 K. Irradiation is seen to cause partial breakup of the original cell structure, although the
dark field view shows that a number of the grains have the same orientation.
In both investigations, as far as could be determined, amorphous and microcrystalline
alloys indeed proved to be highly resistant to displacement damage.

14.9. N i - S i Alloys
The ordered, coherent y' phase Ni3Si has a solubility in the Ni-rich terminal solid solution
of 17.6 a/o at 1125°C which does not decrease to below approximately 10 a/o, even at 400°C
(Hansen, 1958). The y' phase is of the Cu3Au (LI2) type structure, which also occurs in the
Nimonic alloy PE-16, and in some irradiated austenitic stainless steels. As such, the Ni-Si
system is attractive for studies of the behavior of y' precipitates without the complications
inherent in use of a complex, commercial alloy.
Silicon is 5.8~ undersized in Ni, and under irradiation segregates to internal and external
point defect sinks (Table 18). This behavior is the basis for most of the observed
irradiation-altered phase stability effects in Ni-Si alloys.
The 7' phase forms readily in undersaturated Ni-Si solid solutions under a variety of
radiation conditions.
Silvestre et al. (1975) exposed a Ni-8a/oSi alloy to 1.4 x 1020n/cm 2 (~0.07 dpa) at 400°C,
to 4.3 x 10 22 n/cm 2 (~20 dpa) at 450°C, and to 1.4 x l02° n/cm 2 (~0.07 dpa) at 500°C. In all
cases Ni3Si plates formed on {ll l} planes. Precipitation of Ni3Si was less pronounced in a
la/oSi alloy.
Barbu and Martin (1977) bombarded alloys of 2, 4, 6 and 8a/oSi with 500 keV Ni + ions.
Dose rates of 2.08 x 10-3/s and 2.08 × 10-4/s were used; total doses were l and 10 dpa. The
7' particles lay on {l 1l} planes, and it was deduced that precipitation was occurring on
prismatic dislocation loops, which serve as point defect sinks and become enriched in Si. The
larger particles were found to be toroids and the smaller ones were discontinuous discs.
Barbu and Martin (1977) irradiated 2,4,6 and 8a/oSi alloys in the 1MV HVEM.
Displacement rates were from 8.7 x 10-5/s to 1.45 x 10-3/s, with total doses of 0.1-1 dpa at
temperatures from ,-~25°C to ~ 700°C. Figure 88, which summarizes the results of the study
on the 2 and 6 a/oSi alloys, shows that a dose-rate-dependent threshold for precipitation exists
at both high and low temperatures. The authors note that the high temperature threshold
is consistent with a solute segregation mechanism, but the low temperature threshold is not.
Precipitate morphologies were generally similar to those found by Barbu and Ardell (1975).
372 P R O G R E S S IN M A T E R I A L S SCIENCE

el o

:!i
i0-~

8.
"o

6 o/o Si
ey' o No.t"

2 oloSi
i0-~ i X ' aNo 7"

i i I , I , I ,
200 4~ 6~
T, °C

FIG. 88. Precipitation diagram for y' formation in Ni-Si alloys under 1 MV HVEM irradiation.
(After Barbu and Martin, 1977.)

Barbu et al. (1980) used e- irradiation in the van de Graaff generator to study
irradiation-induced precipitation at a very low displacement rate. Ni-4a/oSi was irradiated
at temperatures of 130-630°C at a displacement rate of 5.5 x ]0-9/s.
Particles of Y' nucleated at internal defect sinks between 250 and 500°C after an incubation
dose of approximately l0 -3 dpa. The authors note that higher incubation doses are required
at higher displacement rates, such as l0 -2 dpa at l0 -4 dpa/s. The high temperature limit for
electron irradiation-induced precipitation was found to decrease by approximately 100°C as
the displacement rate decreased from 1.5 x 10-3/s to 5.5 x 10-9/S, as is shown in Fig. 89.
Barbu et al. proposed that Ni3Si precipitation occurred when a certain critical Si
emrichment was reached at point defect sinks. Lower displacement rates would give less direct
recombination of point defects, hence more solute segregation and faster Ni3Si precipitation
than would high rates
Janghorban and Ardell (1979) irradiated Ni-2, 4, 6 and 8 a/oSi alloys with 400keV
protons. Damage levels were between 0.3 and 0.25dpa at a dose rate of 2.37 x 10-5/s;
temperatures were 400-600°C. In addition to Y' formation at point defect sinks, regions of
fine coherent particles formed in sink-free regions in the 8a/o alloy at 550 and 600°C and in

a~
t • eel. /

" 7

i6 ~ !
I
i66 I
I
167 I
i
168
ol- • .4= o
i , , , I i
0 500
T, °C
• ),' Precipitation
o No ~" precipitation

FIG. 89. Precipitation diagram for formation of y' in Ni-4a/oSi under electron irradiation at various
displacement rates. (After Barbu el al., 1980.)
PHASE STABILITY U N D E R I R R A D I A T I O N 373

the 6a/o alloy at 600°C. The occurrence of irradiation-induced precipitation in the absence
of nearby point defect sinks was taken to rule out solute segregation as a mechanism.
Lam et al. (1981) performed an analysis of solute segregation under the irradiation
conditions of Janghorban and Ardell (1977), and found that the large depth dependence of
displacement damage production by 400 keV protons and a strong coupling of interstitial and
solute fluxes gave a net transient flux of Si into the peak damage region. Calculations showed
that in the 6 and 8a/o alloys the Si concentration in this central region would exceed the
solubility limit, and could therefore give 7' formation. The irradiation-induced precipitation
results of Janghorban and Ardell are thus entirely consistent with a mechanism of solute
segregation.
Robrock and Okamoto (1980) irradiated Ni-6a/oSi to 3 dpa with 3.5 MeV Ni + ions at
displacement rates of 4.5 x 10-4/s and 2 x 10-2/s. At the lower displacement rate V'
formed at 420, 530 and 610°C but not at 710°C. At the higher rate, 7' formed at 580 and
614°C, but not at 480 and 725°C. The ),' phase formed at dislocation loops and at the sample
surface, indicating that irradiation-induced solute segregation was responsible for precip-
itation in the thermally single phase alloy.
Potter et al. (1979) irradiated a Ni-12.7a/oSi alloy with 3.5 MeV Ni ÷ ions at temperatures
of 400-700°C and displacement rates of 2.7 x 10-4/s, 2.7 x 10-3/s, and 2.1 x 10-:/s. For-
mation of V' occurred on dislocation loops, stacking faults, the external surface, and grain
boundaries at 650°C and below. The alloy is thermally two-phase over the range of irradiation
temperatures.
Potter and Wiedersich (1979) studied precipitate restructuring in a Ni-12.7a/oSi alloy
under 3.0 MeV Ni ÷ irradiation. At temperatures of 450-650°C and low displacements, the
mean 7' particle size obeyed the F3 versus t relationship predicted by the Wagner-Lifshitz-
Slyozov coarsening theory, as shown by Fig. 90. Above 5-8 dpa, ~' particles which had
nucleated at dislocation loops consumed the intragranular particles and made further
coarsening measurements impossible. The coarsensing rate constant was found to agree with
the theoretical value if an irradiation-enhanced diffusion coefficient were assumed.
Potter and Hernandez (1981) studied irradiation-induced disordering in a Ni-12.7a/oSi
alloy. Surface films of ~' were formed during higher temperature irradiation; subsequent
disordering was then detected through changes in thermal emissivity or by transmission
electron microscopy.

20

oc:~
/5-
600%/ ,i

e o

~8
?o

I 3 4 5
Dose(dpo)
FIG. 90. Average precipitate diameter in Ni-12.7a/oSi vs. displacement dose from 3 MeV Ni + ions.
Irradiation temperatures are indicated. Coarsening rates are orders of magnitude above thermal
values. (After Potter and Weidersich, 1979.)
374 PROGRESS IN M A T E R I A L S SCIENCE

10

0.6 "-.

°\°
JJ o.z

005 Predicted " \ "


~ . 252'C
0,03 I I I I I I
2 4 6 8 I0 12 14
Time (min)

FIG. 91. Fractional emissivity change vs. Bombardment time for Ni-12a/oSi under Ni ÷ ion
irradiation at 10-3dpa/s. (After Potter and Hernandez, 1981.)

Irradiation at 10-3 dpa/s with Ni + ions of 2 MeV energy at temperatures of 337°C or below
produced disordering, as shown in Fig. 91.
The disordering curves were modeled by assuming creation of disordered zones of volume
at a rate N. Then,

where et, e~, and e0 are thermal emissivities at times t, oo, and 0, respectively. Transmission
electron microscopy studies showed ~ 2 × 10 -s zones/~a.s were produced. A fit of the
preceding equation to the results in Fig. 91 gave a disordered zone diameter of 86A,
somewhat larger than the ~ 50/~ zone observed in the electron microscope. Potter and
Hernandez attributed the discrepancy to partial disordering between the zones.
The expression for irradiation-induced disordering (Section 8) is:

dS
- eKS,
dt

where S = long range order parameter, K = displacement rate, and e = atoms disordered/
atom displaced. Application of this equation to the 252°C data in Fig. 91 gives e - 7, which
is somewhat smaller than the value of 10-100 originally estimated by Liou and Wilkes (1979).
Butler and Orchard (1981) studied irradiation-induced disordering of Ni3Si in the HVEM
and obtained much the same results as described earlier for Ni3A1. The disordering cross
section for Ni3Si at 300 K and 1 MV was, however, larger: 100 x l0 -28m 2 for Ni3Si as
compared with 51 × l0 -28 m 2 for Ni3A1.
Potter (1981) found that 1 MeV Kr + ion irradiation of Ni3Si in a 12.7a/oSi alloy at 200 K
first produced disordering, but that by 1.3 dpa the disordered Ni3Si had transformed to NisSi2,
which was calculated to have a lower free energy than disordered Ni3Si, but a higher free
energy than ordered Ni3Si. This study illustrates the ability of irradiation to drive a normally
stable phase into a higher free energy state, from which it may transform spontaneously into
a phase not normally stable.

FIG. 92. Transmission electron micrographs of 5 MeV Ni 2+ irradiated N i - T h O 2 showing the distribution of small ThO 2 particles. Displacement level is 50 dpa. (a)
Bright field; (b) dark field using an arc of the (111) ThO 2 diffraction ring. (Electron micrographs courtesy R. H. Jones and Journal of Nuclear Materials.)
376 PROGRESS IN MATERIALS SCIENCE

)000 N ICKEL MATRIX


OoOoOooooo
0 0 0 0
0 0 000000
0 O00 0 0
0 O0 0
0000 0
0 ~ 0 0 ~ 0 0
O0 o

.l°o°o
RA0,0s, r I~r I
FIG. 93. Schematic diagram of irradiated Ni-ThO 2 microstructure. (Courtesy R. H. Jones and
Journal of Nuclear Materials.)
PHASE STABILITY UNDER IRRADIATION 377

14.10. N i - T h Alloys
Thoria (ThO2) is a very stable oxide which has little solubility in nickel. This low solubility
affords fine thoria particles a high resistance to thermal coarsening, even at temperatures near
the melting point of nickel.
Jones (1978) studied the stability of a Ni-l.8volg/o thoria alloy under 5 MeV Ni 2+ ion
bombardment. The alloy was irradiated to 50 dpa at 3 x 10 -2 dpa/s over the temperature
range 525-750°C.
After irradiation the Ni/ThO2 interface became irregular and the particles became
surrounded by a halo of extremely fine ThO2 particles, as seen in Fig. 92. This halo effect
is shown schematically in Fig. 93.
Jones attributed this halo formation to recoil resolution (see Section 7) and estimated the
maximum range of energetic Th atoms recoiling from the ThO2 surface as 550 A, which is
considerably greater than the ~ 100 A, which separates the halo from the original particle
surface. However, due to lower energy recoils and recoils from inside the particle, the
maximum deposition of Th would be at a smaller distance. As such, the halo would be
nearer the particle, and agreement between measured and estimated recoil distances is not
unreasonable.

14.11. N i - T i Alloys
Near equiatomic Ni-Ti alloys have shape memory properties which result from a
martensitic transformation.
G. Thomas et al. (1982) irradiated two Ni-Ti alloys with 0.5-2 MV electrons in the HVEM
at damage rates between , - , 1 0 -4 and ~10-2dpa/s. Samples were irradiated at room
temperature in the high temperature B2 state, and at -80°C in the low temperature
martensitic state. Figure 94 shows that during irradiation at room temperature a 50a/oTi
sample became amorphous in 30 min at 10202 MeV e-/cm2, s ( ~ 20 dpa). At -80°C a 49a/oTi
sample became amorphous after a 13-min dose of 3.5 x 10TM 1MV e-/cm2.s (,,-0.3dpa).

14.12. Nimonic P E - 1 6
Nimonic PE-16 is a 7' Ni3(A1,Ti) precipitate-strengthened alloy of interest in fast reactor
structural applications. The nominal composiiton of this alloy is given in Table 17.
In a classic experiment, Nelson et al. (1972) studied the effects of 46 MeV Ni ion irradiation
on several alloys, including PE-16. After an unstated (but probably high) number of
displacements at 600°C, 7' precipitates were found to have taken on lenticular shapes in
association with the dislocation structure.
Sharpe and Whapham (1974) studied the effects of neutron irradiation on the 7' particle
distribution in PE-16 which had been solution annealed or solution annealed and aged. After
irradiation in the Dounray fast reactor to up to 27 dpa at temperatures of 530-640°C chains
of 7' precipitates were found to decorate the dislocation network. Precipitates were also found
on grain boundaries. Irradiation-enhanced coarsening of matrix 7' was not observed.
Gelles (1979, 1981a,b, 1983) made an extensive study of precipitate redistribution in
irradiated Nimonic PE-16. Irradiation was of solution annealed, of aged, and of overaged
samples, at 400-650°C, to 5.4-10 × 1022n/cm2 (E > 0.1 MeV) or -,,27-50 dpa in EBR-II.
Very extensive 7' decoration of edge and screw dislocations and of voids was observed after
irradiation. Figure 95 is a dark field electron micrograph of a solution annealed and irradiated
sample showing fine matrix precipitation and precipitation on dislocations. The zigzag
378 PROGRESS IN MATERIALS SCIENCE

configurations are Archimedes spirals which have formed by 7' precipitation on screw
dislocations which were climbing due to point defect absorption. Gelles noted that the ?'
particles in aged samples were larger after irradiation, so that inverse coarsening of the type
discussed in Section 9 was not operative. This observation is not surprising since the relatively
high solubility of 7' in this alloy would make inverse coarsening observable only at low
temperatures. Gelles concluded that precipitate redistribution in his samples was governed
by irradiation-induced solute segregation to point defect sinks and by the same kinds of
nucleation, growth, and coarsening processes observed under thermal conditions. Agreement
was found between his coarsening measurements and the Wagner-Lifshitz-Slyozov coarse-
ning theory (Eq. 9.2) when the modeling incorporated an irradiation-enhanced diffusion
coefficient.
14.13. Inconel Alloys
Table 17 gives compositions of some of the Ni-based Inconel alloys of interest in fast
reactor applications. L. Thomas (1981) describes the phases which appear in irradiated
In-706. These are:
7' L12--ordered FCC Ni3(Ti,Nb,A1).
?" DOz2--ordered BCT Ni3Nb.
q DO24--ordered HCP, nominally Ni3Ti, although Nb and AI may replace about half
the Ti.
6 Orthorhombic Ni3Nb.
? FCC matrix phase.
Thomas irradiated solution treated and solution treated and aged In-706 in EBR-II to up
to 15 x 1022n/cm 2 ( E > 0 . 1 MeV) or ~ 7 5 d p a at temperatures from 400 to 735°C. The
pre-irradiation microstructure of the solution annealed alloy (Fig. 96) was single phase solid
solution apart from MC carbides, whereas the aged alloy contained isolated ?" discs and
cuboidal ?' particles enclosed by 7" discs, and 7' on dislocations, as seen in Fig. 97.
Irradiation produced phase stability effects in In-706 at the level of complexity observed
in type 316 stainless steel (Section 13). Figure 98 summarizes these effects for a solution
annealed and aged alloy irradiated to 5.7 x 1022n/cm 2 (E > 0.1 MeV) at 510°C.
Irradiation effects include Frank loop formation in ?' particles, partial ?" dissolution, 7'
redistribution at voids and Frank loops, and ?"--*?' re-precipitation. Coatings of ?' on
stacking faults give (111 ) relrod satellites around ?' fundamental and superlattice reflections
in the accompanying diffraction pattern. Stacking faults which formed within ?' particles in
the aged alloy obtained a coating of ?' as they grew beyond the ?' particle boundaries. An
extrinsic stacking fault in ?' gives seven planes of HCP stacking, hence constitutes incipient
q-phase.
The original ?' particles in the aged material were redistributed to Frank loops, voids, and
matrix precipitates. Extensive precipitation of ?' and ?" from solution occurred in the solution
annealed alloy. Precipitation of 7' at dislocations accompanied by denucfing of the sur-
rounding area of 7" occurred at 400-500°C, and was taken to suggest that elements in ?"
segregate to point defect sinks and precipitate as 7%
Above 650°C q-phase formed rapidly at the expense of 7' and 7" in both irradiated and
thermally aged In-706. Irradiation was found to affect this transformation little, if at all.
No 6-phase formed during thermal aging or irradiation.
Bell and Lauritzen (1982) irradiated aged samples of In-706 and 718 in EBR-II at
temperatures between 400 and 650°C to a peak fluence of 6 x 1022n/cm 2 (E > 0.1 MeV), or
iiI,III
i i /

>

~v

~v
>

>

Z
FIG. 94. Electron micrograph and diffraction patterns of equiatomic NiTi alloy. Foil orientation []~1] of austenite. (a) Before room
temperature irradiation, in fully martensitic state; Co) in austenitic state after 5 min of 2MV e- at 1019e-/cm2/s; (c) after a further room
temperature irradiation for 30rain at 102°e-/cm2/s, about 20dpa. Austenite has become amorphous in irradiated region. (Electron
micrographs courtesy G. Thomas, H. Mori, H. Fujita, R. Sinclair and Scripta Metallurgica.)
380 PROGRESS IN M A T E R I A L S SCIENCE

FIG. 95. Dark field electron micrograph of solution treated PE-16 irradiated to 5.4 x 1022n/cm 2
(E > 1 MeV) at 620°C. Precipitation on the climbing screw dislocations has produced Archimedes
spirals. (Electron micrograph courtesy D. S. Gelles.)
PHASE S T A B I L I T Y U N D E R IRRADIATION 381

Fro. 96. Dark field electron micrograph and diffraction pattern of solution treated and annealed
Inconel 706 showing ),'-?" microstructure. Only one y" variant is seen. (Electron micrograph and
diffraction pattern courtesy L. E. Thomas.)

FIG. 97. Bright field view, selectively imaged dark field views, and electron diffraction pattern of
solution treated and annealed Inconel 706 irradiated to 5.7 x 10Z2n/cm2 (E > 0.1 MeV) at 510°C.
Foil orientation is near [001]. Relrod image is from ),' coated stacking faults. (Electron micrograph
and diffraction pattern courtesy L. E. Thomas.)
382 PROGRESS IN M A T E R I A L S SCIENCE

IRRADIATIONTEMPERATURE
400 450 500 550 600 650 700 *C

VOID FORMATION
y' PRECIPITATESFROM
SOLUTION
y" PRECIPITATESFROM
SOLUTION
PRECIPITATESFROM
SOLUTION
)," DISAPPEANCE
(~,'-- ~,')

y' REDISTRIBUTION
STACKINGFAULTSIN y'
(INCIPIENT~/)
y' REDISTRIBUTIONTO
DISLOCATIONS
y'AND ~'" COARSENING
STABLE 'r/ FORMATION

FIG. 98. Microstructural instabilities in neutron irradiated Inconel 706 at fluences from 2 to
15 × 1022n/cm 2 ( E > 0 . 1 MeV). (After L. E. Thomas, 1981.)
PHASE S T A B I L I T Y UNDER IRRADIATION 383

Table 19. Phase stability in nickel-based alloys


Alloy Treatment Result References
Ni-13.5a/oAl 550°C, 100 keY Ni + Inverse coarsening of y'. Nelson et aL (1972)
10-2dpa/s, up to 100dpa
Ni-13.5a/oA1 25°C, 100 keV Ni +, Disordering of y'. Nelson et al. (1972)
10 -2 dpa/s, 0.1 dpa
Ni-13a/oA1, 16a/oAl 300-750°C, 650 kV e-, Irradiation enhanced Ro and Mitchell (1978)
2.3-5.8 x 10-3 coarsening.
dpa/s, a few dpa
Ni-13a/oA1 300-650°C, 3.2 MeV Ni + Coarsening followed by Potter and Hoff (1976)
2.7 x 10-3dpa/s, up inverse coarsening.
to 34dpa
Ni-14a/oAl 525-725°C, 2.8 MeV Ni +, Invers~ coarsening at Kirchner et al. (1976)
4.4 x 10-2dpa/s, 725°C, nucleation of new
4-125 dpa ~,' at other temperatures.
Ni-13a/oA1 650°C, 3.2 MeV Ni +, Irradiation enhanced Potter and Ryding
2.7 x l0 -3 dpa/s, up coarsening to ~ 20 dpa (1977)
to 30 dpa followed by renucleation.
Ni-12.8a/oA1 450-700°C, 3.5 MeV Ni +, Irradiation enhanced Potter and McCormick
2.7 x 10-3 dpa/s, up coarsening. (1979)
to 12 dpa
Ni3AI 10-300K, up to 1 MV Disordering dependent on Butler and Orchard
in HVEM T, voltage, and orien- (1981)
tation.
Ni3A1 < 190 K, 650 kV HVEM, Complete disordering. Liu et al. (1981)
0.9 dpa
Ni3A1 200-240 K, 650 kV ~0.1-0.15 residual order. Liu et aL (1981)
HVEM, 2 dpa
NiA1 150 K, 650 kV HVEM, ~0.5 residual order. Liu et al. (1981)
6.6 dpa
Ni-0.1, 0.3, 0.7a/oBe 425-625°C, 3.2MeV Ni +, Precipitation in a ther- Okamoto et al. (1975)
3 x 10-3dpa/s, 1-30dpa mally single phase alloy.
Ni-la/oBe 390, 470°C, 650 kV e-, Homogeneous precip- Mukai and Mitchell
3.7 x 10 -4 or 9.3 x 10 -4 itation in under-saturated (1982)
dpa/s, up to 2.2 dpa alloy.
Ni-2.5a/oBe 300°C, 4 x 1017 fast More precipitation than Kernohan et al. (1956)
n/cm 2, ~ 2 x 10 -4 dpa during thermal anneal.
Ni-2w/oBe 70°C, 4.6 x 10~9-1.2 Enhanced GP-zone for- Grinchuk and Loboda
X l021n / c m 2 (E > 0.8 MeV) mation on post-ir- (1974)
~ 0.02-0.5 dpa radiation annealing.
Ni-2w/oBe 50-70°C, 7 x 1016- Dissolution of GP-zones. Grinchuk and Kirsanov
1.2 x l021 n/cm 2 (E > (1974)
0.8 MeV) ~ 4 x l0 -5 - 0.5 dpa
Ni-0.3w/oC 40°C, 5 x 10t3-2 x l017 Accelerated solid solution Shriver and Wuttig
fast n/cm 2, ~ 2 x l0 - s - decomposition on post- (1972)
10-4dpa irradiation annealing.
Ni~,la/oCu 373-640K, up to 1.6 x 102° Spatially periodic decom- Wagner et al. (1980,
2-3 MV e-/cm 2, ~ 1.6 x 10 -2 position. 1982)
dpa
Ni-6a/oGe 125-4150°C, l MV e-, Precipitation in single Barbu (1980)
3.3 x 10-4-1.5 x 10-3 phase region under some
dpa/s. conditions.
Ni-20a/oMo 50-1050 K, 1 MV HVEM, Complete disorder at low Banerjee et al. (1983,
with SRO or LRO 10-3-10-2dpa/s, up to T, mixed LRO and SRO 1984)
~ 2 dpa at intermediate T, LRO at
high T.
Ni-20a/oMo 400-900 K, 1 MV e-, SRO--~LRO. Carpenter and Kenik
with SRO or LRO 4 x 10-3 dpa/s, up to 41 dpa (1977)
Ni-40a/oNb 300, 900 K. 3.5 MeV Ni 2+, No crystallization of Rechtin et al. (1978)
6-20 dpa amorphous samples, re-
version of partly crys-
talline to amorphous.
Ni-40a/oNb 800, 900 K, 5.5 MeV Ni 2+, Thermal evolution of mi- Chernock and Russell
2, 20 dpa crostructure halted. (1984)
384 PROGRESS IN M A T E R I A L S SCIENCE

T a b l e 19. Continued
Alloy Treatment Result References
Ni- 15a/oNb 800 K, 5.5 MeV Ni 2+, Formation of precipitate Chernock and Russell
20 dpa haloes around existing (1984)
particles.
Ni-8a/oSi 400-500°C, up to 4.3 x 1022 7' plates from on {111} Silvestre et al. (1975)
n/cm 2 ( ~ 20 dpa) planes.
Ni-2, 4, 6, 8a/oSi 500°C, 500 keV Ni +, 2.08 y' discs and toroids form Barbu and Ardel| (1975)
× l0 3 and 2.08 x 10-4 on { 111 } planes.
dpa/s, l and 10dpa
Ni-2, 4, 6, 8a/oSi 25-700°C, 1 MV e-, 1.45 Dose-rate-dependent Barbu and Martin
x 10-3-8.7 x 10 -s threshold for precip- (1977)
dpa/s, 0.1-1 dpa itation in under-saturated
alloys.
Ni-4a/oSi 250-500°C, 2 MV e-, 7' formation at sinks after Barbu et al. (1981)
5.5 x 10-gdpa/s, 10 -3 dpa.
10-3 dpa
Ni-2, 4, 6, 8a/oSi 400-600°C, 400 keV Fine coherent particles Janghorban and Ardell
protons, 2.37 x 10 -5 not associated with sinks (1979)
dpa/s, 0.13-0.25 dpa form in the 6a/o alloy at
600°C, in the 8a/o alloy at
550 and 600°C.
Ni-6a-oSi 420, 530, 610°C, 3.5 7' formation at sinks. Robrock and Okamoto
MeV Ni ÷, 4.5 × 10 -4 (1980)
dpa/s, 3 dpa
Ni-6a/oSi 580, 614°C, 3.5 MeV Ni +, 7' formation at sinks. Robrock and Okamoto
2 x 10-2dpa/s, 3dpa (1980)
Ni-12.7a/oSi 400-700°C, 3.5 MeV Ni ÷, 7' formation at internal Potter et al. (1979)
2.7 x 10-4-2.1 x 10 -2 and extenals sinks.
dpa/s, up to ~ 12 dpa
Ni-12.7a/oSi 400-650°C, 3.0 MeV Ni ÷, Irradiation enhanced Potter and Wiedersich
2.7 x 10 -3 dpa/s, up coarsening following ~3 (1979)
to ~ 8 dpa versus t law.
Ni-12.7a/oSi 80-337°C, 2 MeV Ni ÷, Disordering of 7' phase. Potter and Hernandez
10-3dpa/s, up to 0.7 (1981)
dpa
Ni3Si 300K, up to 1 MV in Disordering, dependent Butler and Orchard
HVEM on T, voltage, and orien- (1981)
tation.
Ni-12.7a/oSi 200K, 1 MeV Kr ÷ Ni3Si disorders, then is Potter (1981)
1.3 dpa replaced by NisSi2.
Ni-1.8 vol.~ThO2 525-750°C, 5 MeV Ni 2+, Formation of ThO2 Jones (1978)
3 x 10-2dpa/s, 50dpa haloes around pre-
existing particles.
Ni-50a/oTi ~25°C, 1020 2 M V e /cm2/s, Sample becomes amor- Thomas et al. (1982)
30 min, ~ 20 dpa phous.
Ni--49a/oTi -80°C, 3.5 x 10Is 1 MV Sample becomes amor- Thomas et al. (1982)
e-/cm2/s, 13min ~0.3 phous.
dpa
Nimonic PE-16 600°C, 46 MeV Ni ions Lenticular plates of 7' Nelson et al. (1972)
decorate dislocations.
Nimonic PE-16 530-670°C, DFR, up to 7' decorates dislocations Sharpe and Whapham
27 dpa and grain boundaries. (1974)
Nimonic PE-16 450-650°C, EBR-II, 5.4- 7' precipitation at dis- Gelles (1979, 1981b,
10 x 1022n/cm 2 (E >0.1 locations and voids, par- 1983)
MeV), 27-50 dpa ticle coarsening.
Inconel 706, 400-735°C, up to 15 Complex interplay be- Thomas (1981)
solution annealed, and × 1022n/cm 2 (E >0.1 MeV) tween 7', 7", and r/
solution annealed and 75 dpa phases.
aged
Inconel 706 and 718, 400-650°C, EBR-II, < 6 Formation of thin plate- Bell and Lauritzen
solution annealed and x 102~n/cm 2 (E > 0.1 MeV) lets of q phase. (1982)
aged ~ 30 dpa
SA Inconel 625 600-650°C, EBR-II, 5 Formation of metastable Appleby et al. (1973)
× 1022n/cm 2 (E >0.1 MeV) 7", Ni3Nb.
~25 dpa
PHASE STABILITY UNDER IRRADIATION 385

about 30 dpa. Thin platelets of precipitate appeared in both alloys over much of the
temperature range. Although the thinness of the platelets made identification difficult, after
extensive study it was concluded that they were of r/, Ni3Ti rather than 6, Ni3Nb.
The tetragonal, y" Ni3Nb phase is metastable with respect to the orthorhombic, 6 Ni3Nb
phase in In-625 (Appleby et al. 1973). However, Appleby et al. observed the metastable y"
phase in solution annealed In-625 which had been irradiated in EBR-II to 5 x 1022n/cm 2
(E > 0.1 MeV), or ,-~25 dpa at 600-650°C. The stable 6 phase formed after thermal ageing
of unirradiated In-625 heat treated for ~ 1000 hr at 650°C (Comprelli and Lewis, 1965). Thus,
irradiation produced a different precipitate phase than appeared thermally.

15. REFRACTORYALLOYS*
Refractory alloys are of interest for fusion reactor first walls because their high melting
points would permit high operating temperatures. In addition, niobium-based super-
conducting alloys may be used in fusion reactor magnets and, as such, must withstand some
irradiation damage. Radiation effects on superconductivity were recently reviewed at length
(B. S. Brown et al. 1978) so that only a few illustrative examples need be cited here.

15.1. M o - T i Alloys
A. Wagner and Seidman (1979) made an atom-probe field-ion microscope study of solute
segregation to voids in a (nominally) Mo-la/oTi alloy which had previously been irradiated
at ~ 700°C to ~ 1022n/cm 2 (E > 1 MeV), or about 5 dpa. The Ti concentration was found
to be the same at the void surface as in the bulk of the material. Titanium is only 0.56~o
oversized in Mo (Table 1) so the size misfit theory of solute segregation would predict that
any Ti segregation would be away from the voids. The amount of such segregation was clearly
small at the most.

15.2. M o - T i - Z r Alloys
Sprague et al. (1979) studied the behavior of TZM (Mo-0.5w/oTi-0.1w/oZr) irradiated in
EBR-II to 22 dpa at 650°C. A low density of globular precipitates was observed; these were
believed due to interstitial impurities in the alloy.

15.3. M o - Z r Alloys
Liou et al. (1979) studied the effect of 14 MeV Cu + ion irradiation at 700-900°C on a
Mo-9. la/oZr alloy which had been aged to give the equilibrium Mo-rich solid solution with
a dispersion of y-phase Mo2Zr precipitates.
After irradiation to approximately 6 dpa at 900°C, incoherent MozZr precipitates formed
at the grain boundaries, with no apparent dissolution of the matrix precipitates. The grain
boundary precipitation was attributed to irradiation-induced solute segregation. However, as
they note, the Zr atom has an effective atomic volume 26~o greater than that of Mo, so that
on a size misfit basis Zr should be segregated away from the grain boundaries and other point
defect sinks. There are, nevertheless, other mechanisms (see Section 5) which could give
irradiation-induced segregation of Zr to the grain boundaries. Heterogeneous nucleation and
growth, assisted only by irradiation enhanced diffusion, could also be responsible for
formation of Mo2Zr at the grain boundaries.
*See also notes in proof (pp. 430-434).
J P M S 2g:~ 4 H
386 PROGRESS IN M A T E R I A L S SCIENCE

15.4. N b - S n Alloys
Karkin et al. (1976) studied the effects of neutron irradiation on the superconducting
properties ofNb3 Sn. The alloys were irradiated to up to 8 x 1019 fast n/cm 2 (,-,0.04 dpa) and
T < 70°C.
The superconducting critical temperature decreased rapidly with fluence, dropping from
17 K before irradiation to 3 K after 6 × 10~9n/cm z, as seen in Fig. 99. The upper critical field
dropped from 230 kG before irradiation to 40 kG after. The long range order parameter, S,
decreased from an initial value of 0.98 to 0.6 after 4 x 1019 fast n/cm 2. The authors attributed
this deterioration in superconducting properties to irradiation-induced disordering of the
Nb3Sn due to local melting. Although local heating due to displacement cascades is too
short-lived ( ~ 1-2 atomic vibration times) to give melting, cascade-induced disordering of the
type discussed in Section 8 is a plausible mechanism for the change in superconducting
properties. The superconducting properties recovered following a 900-950°C anneal, which
is the usual range of ordering temperatures for Nb3Sn.

15.5. N b - T i Alloys
Tsubakihara et al. (1975) studied the effect of neutron irradiation on the superconducting
properties of alloys of Nb-47.6a/oTi and Nb-59.8a/oTi. Irradiation was to up to
1.3 × 1018n/cm 2, in a university research reactor, with temperature and neutron energy
unspecified. The temperature was probably in the vincinity of 25°C, and even if the fluence
were all fast neutrons, fewer than 10 -3 dpa were produced.
Irradiation was found to decrease the critical current density and superconducting
transition temperature in the 59.8a/oTi alloy by the order of 10-20~o, depending on prior heat
treatment. These decreases were attributed to cascade dissolution of the Ti-rich precipitates
which contribute to the superconducting properties of the alloy. Appreciable irradiation
dissolution would not occur at a 10 -3 dpa; however, the residues of displacement cascades
could degrade superconducting properties by disturbing the regularity of the crystal lattice.
H. W. Weber (1982) reviewed irradiation effects in alloy superconductors, particularly
Nb-Ti alloys, and noted that alloys which depend on precipitates for flux pinning are more
radiation resistant than those which depend on dislocation cell walls. He attributed the
greater radiation sensitivity of the latter class of materials to irradiation induced dislocations
in the cells. These dislocations reduce the difference in pinning force between the cells and
cell walls, and degrade the superconducting properties of the alloy.

2O
I,k

, e 10

2 4 6 8 x 10=9
Fluence, n/cm2
FIG. 99. Decrease in superconducting critical temperature o f Nb3Sn with fast neutron fluence. (After
Karkin et al., 1976.)
PHASE STABILITY UNDER IRRADIATION 387

15.6. T i - A l , T i - A I - M o , Ti-AI-V, T i - V Alloys


Erck et al. (1979) irradiated a single phase Ti-14.4a/oA1 alloy with 3 MeV Ni + ions at
600°C to 1.3-14.7 dpa. A TEM study showed strong segregation of the undersized AI atoms
to voids, grain boundaries and the irradiated surface. The equilibrium ~2 (Ti3AI) phase
precipitated at these sinks as seen in Fig. 100.
Potter (1981) irradiated a Ti-14.4a/oA1 alloy to 1, 5, and 12 dpa with 3 MeV Ni + ions at
690°C. The displacement rate was presumably the order of 10 -3 dpa/s. The pre-aged and
unirradiated microstructure had a uniform dispersion of 250/~, ~2 (Ti3A1) particles. The Ti3A1
phase is ordered HCP, and has much the same relationship to the HCP ~ matrix phase as
does the ?' phase in FCC Ni or Fe-based alloys. Unlike 7-7' alloys, however, irradiation
produced little, if any change in particle size, shape, or number density.
Wang et al. (1982) studied solute segregation of V, A1, and Mo in Ti-based alloys irradiated
at 650°C with 2.1 MeV Ar + ions. In this study Ti-6w/oA1-4w/oV, Ti-8w/oAl-lw/oMo,
Ti-8.7w/oAl and Ti-3w/oV were irradiated to surface doses of ~0.75 or ~ 3 dpa at a
displacement rate of ~ 8 x 10 -4 dpa/s. Auger depth profiling was used to measure the degree
of solute segregation to the irradiated surface.
Segregation of all three solutes to the surface was observed. The peak concentration of Mo
was ~2.2 times the bulk concentration and that of V was ~ 6 times that of the bulk.
Aluminum segregation occurred under irradiation, but to a lesser degree. Figure 101 shows
the degree of V and AI segregation. All three solutes are undersized in ~-Ti and on the basis
of the size misfit theory of solute segregation were expected to move toward the irradiated
surface.
Brimhall et al. (1981) studied solute segregation in a Ti-6.35w/oA1--4.0w/oV alloy under
5 MeV Ni 2+ ion bombardment. Irradiation was to a surface dose of 2 dpa at 3 x 10 -3 dpa/s
at temperatures between 725 and 825 K. No irradiation induced solute segregation was
observed, which is in contrast to the results of Erck et al. (1979) and Wang et al. (1982) who
found strong solute segregation effects.
Agarwal et al. (1979) irradiated a Ti-6w/oA1-4w/oV alloy with 2.4 MeV V + ions to a dose
of from 3-25 dpa. The displacement rate was presumably in the 10-2-10 -3 dpa/s range. In
some cases the sample was also irradiated with 0.95 MeV ~ particles to inject 10 appm
He/dpa. Profuse precipitation of BCC/~ platelets in the HCP ~ matrix occurred at 450. 550,
and 650°C, as seen in Fig. 102. Precipitation was not observed in unirradiated control samples
which had the same thermal history.
Peterson (1982) irradiated duplex annealed, B-annealed and mill-annealed
Ti-6.4w/oA1-3.9w/oV in EBR II to 4 . 2 x 1022n/cm 2 ( E > 0 . 1 M e V ) at 395°C and to
5.0 x 1022n/cm 2 (E > 0.1 MeV) at 450 and 550°C.
The pre-irradiation microstructure of the duplex annealed alloy consisted of equiaxed
grains with regions of 8. Irradiation at 550°C gave ~2 precipitation on voids and some coarse
precipitates.
The//-annealed alloy had a pre-irradiation microstructure of acicular ct plates. Irradiation
at 450 and 550°C produced ~2 and fl precipitates.
Prior to irradiation the mill-annealed alloy had a microstructure of equiaxed ct grains with
intergranular particles of fl phase. After the 550°C irradiation, voids in the ~ phase were
coated with ~t2.
Ayrault (1981) irradiated mill annealed Ti-8w/oAl-lw/oV-lw/oMo with 2.4 MeV V ÷ ions
or with 2.4 MeV V ÷ and 0.95 MeV He + at a ratio of 10 appm He/dpa. Irradiation was to
from 3 to 25 dpa at 5 x 10 -3 dpa/s at nominal temperatures of 450, 550, and 650°C. The
388 PROGRESS IN M A T E R I A L S SCIENCE

pre-irradiation microstructure was mostly equiaxed ~ grains with some small untransformed
fl grains where three or more ~ grains came together. The ~ grains contained a uniform
distribution of fine ( < 10 nm) ct2 particles.
Precipitation of fl-phase was observed under all irradiation conditions. A fine dispersion
of particles formed at 450°C; precipitation was on a coarser scale and tended to occur on
subgrain boundaries at higher temperatures. Matrix and grain boundary precipitation of fl
phase is seen in Fig. 103. No dose dependence of precipitation was observed. Precipitation
of fl-phase did not occur in unirradiated control specimens.
Irradiation at 450 and 550°C gave dissolution of all or nearly all of the %. Irradiation at
650°C reduced the volume fraction of ~2 and redistributed the rest to fl precipitates, voids
and dislocation loops.
Ayrault noted that V, Mo and A1 are all undersized in Ti, and therefore should segregate
to point defect sinks during irradiation. Dissolution and redistribution of the ~2 phase and
precipitation of the fl phase were then attributed to irradiation induced segregation of V, Mo,
and AI to defect sinks.

15.7. V - B Alloys
Irradiation studies of vanadium and its alloys have been motivated largely by the potential
use of these materials as fusion reactor first wall materials. Good high temperature strength
makes this class of alloys attractive.
Rau and Ladd (1969) studied the effects of low fluence neutron irradiation on vanadium
alloys containing 10ppm B by weight and somewhat larger quantities of C, O, N, and Fe.
The samples were irradiated to 5.4 x 1019n/cm 2 (E > 1 MeV) and to 3.5 × 1020n/cm 2 (ther-
mal) at ,-~70°C. The visible damage was found to be in the form of concentric circular
cylinders with inner shells 1.3/~m in radius and the outer shells 2.6 ~tm in radius. The cores
of the cylinders were fibrous V3B2 precipitates aligned in (100) directions; these were present
at the start of irradiation. The fast neutron flux would give less than 0.1 dpa, so that neutron
displacement damage was probably relatively unimportant. However, boron has a large
transmutation cross section for thermal neutrons for the following n,~ reaction:
B 1° + n---~Li 7 + ct.
The Li and ct particles have recoil energies which transport then about 1.4 and 2.8/~m,
respectively, from the site of the transmutation event. These distances are close to the
measured shell radii. So it was concluded that the inner shell was due to damage from Li
recoil and the outer from ~t recoils. The damage is shown schematically in Fig. 104.
Subsequent annealing at temperatures to 1400°C gave fine pores in the outer damage layer
and precipitates in the inner. The pores were identified as He-filled bubbles and the diffraction
pattern of the precipitates was consistent with orthorhombic LiV2Os.

15.8. V - C Alloys
Takeyama et al. (1982a) irradiated vanadium with 0.026 to 0.886a/oC at 973 K to 48 dpa
with 200 keV C ÷ ions. Carbon was found to segregate to the irradiated surface. Takeyama et
al. (1982b) neutron irradiated the same alloy to 5 x 1020 fast n/cm 2 at 773 K and 102~ fast
n/cm 2 at 740 K. These fluences correspond to approximately 0.25 and 0.5 dpa. The thermally
stable V2C phase disappeared during irradiation and was replaced by thin plates of a new,
unidentified C-rich phase. Takeyama et al. referred to this phase as a quasi-carbide, and
PHASE STABILITY UNDER IRRADIATION 389

o b

I 4
0.1~m
c d
FIG. 100. Precipitation o f ~t2-Ti3Al at point defect sinks in Ti-14.4a/oA1 irradiated with 3 MeV Ni +
ions at 690°C. (a) Bright field; (b) dark field views of the same area show ~ coatings on voids;
(c) shows a2 film on irradiated surface; (d) ct2 film at grain boundary. Foil planes are near (0001).
(Electron micrographs courtesy D. I. Potter and Journal of Nuclear Materials.)
390 PROGRESS IN MATERIALS SCIENCE

0.25 Ti-6A1-4V 0.75dpa --


at 650 °C
0.20 \ ! ~ OCv/CT,
0.15 0 CAI ICTi
0.10 -o-o c o o---

0.05 - %-0 o c o--


i-= , I I 1 I I
0.30 Ti-6AI-4V As Prepared
,.9
0.25i - 0 CV/CTI
0.20~- a CAI/CTi

0.rsk~,
O.lO~-~O-o-o-o o--AI
005 Ldp¢,Oo"-'o-o-c o-- v

o I I I 1 I I
500 I000 1500 2000 250q
Sputlering time (s)

FIG. 101. Solute segregation to the surface in Ti-based alloys irradiated with 2.1 MeV Ar + ion.~
(a) Not irradiated; (b) irradiated to 0.75 surface dpa at 650°C. (After Wang et al., 1982.)
PHASE S T A B I L I T Y U N D E R I R R A D I A T I O N 391

FIG. 102. Precipitation of fl phase in Ti-6w/oAl-4w/oV during 2.4 MeV V + ion irradiation. (a) Dark
field view, 25 dpa at 465°C; (b) bright field view, 12 dpa at 547°C; (c) bright field view, 25 dpa at
660°C. (Electron micrograph courtesy G. Ayrault and Journal of Nuclear Materials.)
392 PROGRESS IN M A T E R I A L S SCIENCE

FIG. 103. Dark field electron micrograph of Ti-8w/oAl-lw/oV-lw/oMo irradiated to 25dpa at


450°C with 2.4 MeV V + ions. fl-phase precipitation has occurred in the grain and in the grain
boundary. (Electron micrograph courtesy G. Ayrault.)
PHASE S T A B I L I T Y UNDER IRRADIATION 393

FlG. 104. Schematic drawing of cylindrical damage shells in neutron irradiated V-10wt. ppmB
showing V3B2 core, Li damage shell and ~t particle damage shell. (After Rau and Ladd, 1969.)

concluded that it consisted of C atoms and vacancies. It thus appears that the formation of
the quasi-carbides took place by a solute-supersaturated vacancy coprecipitation process of
the type discussed in Section 4.1.
Weber et al. (1975) irradiated V-50 ppm C (by weight) with 18 MeV Cu ÷ ions at 600-750°C.
Total damage was from 1 to 5 dpa at 3 x 10 -4 dpa/s. Agarwal and Taylor (1976a) irradiated
V-120ppm C (by weight) with 3.25 MeV Ni ÷ ions at 650-700°C. Irradiation was to a
maximum dose of 55 dpa. In both studies irradiation gave a dispersion of fine VC precipitates
even though the carbon content is below the solubility limit.
Vanadium carbide has a larger volume per V atom than does the matrix (f = 0.29), and
must absorb vacancies to form. The irradiation conditions in both studies were such as to
give a high vacancy supersaturation; incorporation of these vacancies into the VC precipitates
would provide a huge increase in the thermodynamic driving force for precipitation.
Accordingly, it appears that the excess vacancy effect discussed in Section 4.1 may have been
responsible for the irradiation-induced precipitation of VC. However, segregation of C to
defect sinks resulting in local solute super-saturations may also be the mechanism for carbide
precipitation.

15.9. V - C r Alloys
Agarwal et al. (1978) and Rehn et al. (1979) irradiated a V-15w/oCr alloy with V + ions.
Heavy Cr segregation to the foil surface was observed in both studies. Chromium atoms are
undersized in the V lattice, and by the size-misfit model for solute segregation would be
expected to be preferentially incorporated into dumbbell interstitials and migrate to sinks,
as observed.
Bentley and Wiffen (1976) neutron irradiated V-10w/oCr alloys to 1.5 × 1022n/cm 2
(E > 0.1 MeV), or approximately 10 dpa at 490-580°C. Fine precipitates were observed, but
as is often the case in such studies, were not identified. Vanadium and chromium form a
continuous BCC solid solution at 700°C (Hansen, 1958) so the V-15Cr should have been
thermally single-phase at irradiation temperatures. As such, the precipitates were induced by
irradiation, but since they were not identified, it is not possible to specify the mechanism for
their formation.

15.10. V - F e , V - M o , V - N b Alloys
Takeyama et al. (1982b) irradiated alloys of vanadium plus 3 a/o of Fe, Mo, or Nb to fast
neutron fluences of 102~/cm 2 at 740 K and 5 x lO2°/cm 2 at 773 K. These fluences correspond
to approximately 0.5 and 0.25 dpa. Undersized Fe atoms segregated toward grain boundaries
394 PROGRESS IN MATERIALS SCIENCE

while oversized Mo and Nb atoms were rejected from grain boundaries. These results are as
expected on the basis of size misfit.

15.11. V - N i Alloys
Lott et al. (1979) studied the effects of 14 MeV Cu + ion irradiation on V-l~oNi alloys.
Irradiation was to 1-5dpa at 6 × 10-4dpa/s.
Nickel segregation to dislocations was observed at both 550°C and 650°C. Observation of
strain contrast in the transmission electron microscope was taken to indicate the onset of
coherent or semi-coherent precipitation at both temperatures.
The phase boundary in the V-Ni system in V-rich alloys is not known at these temperatures
due to sluggish precipitation kinetics. In view of this, the authors suggest that irradiation-
enhanced diffusion may be responsible for the observed precipitation. Other effects, including
irradiation-induced solute segregation, could also be operative.

15.12. V-Si Alloys


Type AI5 superconductors, in particular Nb3Sn are of interest in fusion reactor applica-
tions. It is therefore important to know the effects of neutron irradiation on the super-
conducting properties of this class of materials.
Meier-Hirmer and Kfipfer (1982) both studied and reviewed the effects of irradiation on
the A15 compound V3 Si, which is not of technical interest but unlike Nb3Sn, may be
produced in bulk in relatively defect free single or poly-crystalline form.
Irradiation to ~ 3 x 1019 fast n/cm 2 ( ~ 0.015 dpa) at 150°C or lower temperatures reduced
the usual 16-17 K critical temperature by up to 90~. Irradiation at higher temperatures
allowed more point defect recombination and led to a smaller decrease in critical temperature.
The upper critical field at 10 K increased by 20~o over the unirradiated value at 10 ~8fast n/cm 2,
then declined at higher fluences. Irradiation to 1.2 x 1019 fast n/cm 2 at 240°C gave a factor
of 10t increase in the critical current density at 13.5 K. The increase was attributed to
irradiation causing the formation of dislocation loops with a size the order of the
superconducting coherence distance. All superconducting properties reverted to their pre-
irradiation state after 400-700°C anneals. Figure 105 shows the recovery of critical
temperature and critical current to their pre-irradiation values on annealing of a sample which
had been irradiated to 1.2 x 10 ~9fast n/cm 2 at 240°C.

I.O
-103

~1,-= o9 e•/ " '0' t ._o1~

0.8 I I e-,l...-e,~e,~ e_ 100


500 600 700 800
Ta ( ° C ) ~

FIG. 105. Recovery of critical temperature and critical current for V3Si irradiated to 1.2 x 1019 fast
n/cm 2 at 240°C and then annealed. Jc is measured at T = 13.5 K, H = 0.5 He2. (After Meier-Hirmer
and Kfipfer, 1982.)
PHASE STABILITY UNDER IRRADIATION 395

1 5 . 1 3 . V - T i - O Alloys
Santhanam et al. (1975) irradiated a V-lw/oTi-200 ppm O (by weight) alloy with 3 MeV
V ÷ ions at 700°C. Irradiation was to from 2.4 to 54dpa at a displacement rate of
5.5 x 10-3 dpa/s. Agarwal and Taylor (1976b) irradiated a V-10w/oTi alloy with a small
amount of oxygen, with 2.7 MeV V + ions; irradiation was at 650°C to 2-60 dpa. In both cases
the precipitates were thought to be TiO, which also forms thermally in V-Ti alloys.
Sprague et al. (1979) found TiO2 precipitates in V-20~Ti alloys irradiated to
8.4 × 1022n/cm 2 (E > 0.1 MeV), or ~40 dpa at 650°C in EBR-II. Precipitates were either
blocky, ~ 150 nm in diameter or rod shaped, 100 nm in cross section by 200-400 nm long.
Figure 106 shows the elongated TiO2 precipitates in the V-20w/oTi alloy.

15.14. W - R e Alloys
Tungsten-rich W-Re alloys show extensive terminal solid solubility ranging from 28a/oRe
at 1500°C to approximately 22a/oRe at 500°C. These solid solution alloys are used in high
temperature thermocouples.
Studies by R. K. Williams et al. (1983) and Sikka and Moteff (1974)* found that several
of these alloys did not remain single-phase under neutron irradiation as is predicted by the
phase diagram.
In the Sikka and Moteff (1974) study, W-25a/oRe was irradiated at 1000°C to
3.7 × 1022n/cm 2 (E > 1 MeV) in EBR-II, which produced 4.2 dpa. Electron diffraction studies
showed the presence of a second phase with a diffracted spot pattern which matched much
better with that of the Z (WRe3) phase than that of the equiatomic tr phase.
R. K. Williams et al. (1983) studied the response of 5, 11, and 25 a/oRe alloys to EBR-II
neutron irradiation at 600-1500°C to fluences of 4.3 x 1021-3.7 x 10 22 n/cm2(E > 0.1 MeV) or
0.5-4.2 dpa. Figure 107 shows the W-Re phase diagram along with alloy compositions and
irradiation temperatures. Precipitates formed in all samples, even the low Re alloys irradiated
at the highest temperatures. Figure 108 shows precipitate particles adjacent a grain boundary
in a 25~o Re alloy irradiated at 1500°C. Electron diffraction patterns of samples irradiated
at 1100°C and above showed the precipitates to be the Z, WRe3 phase rather than the tr phase.
At lower temperatures the diffraction patterns were too smeared to be indexed. Whereas the
phase diagram may be sufficiently uncertain that the 25~ Re alloys of Sikka and Moteff were
in the two-phase region, most of the alloys in this study were well into the (thermally)
single-phase region.
The question then arises as to what mechanism is responsible for irradiation-induced
precipitation in a single-phase region of a phase which is not even the least unstable. R. K.
Williams et al. (1983) found spherically symmetric strain fields around the Z precipitates (see
Fig. 109), which persisted even after the particles had been etched out. The strain fields were
attributed to enrichment of shells around the particles by undersized Re atoms. This
enrichment is opposite to the effect in precipitation under thermal conditions, where a
growing Re-rich particle would be surrounded by a Re-poor diffusion field. R. K. Williams
et al. (1983) infer from the enriched zones that irradiation-induced segregation of the
undersized Re atoms to point defect sinks is responsible for x-phase formation.

*The two sets of alloys were irradiated in the same experiment. R. K. Williamset al.'s manuscript was delayed
in publication.
396 PROGRESS IN MATERIALS SCIENCE

The question remains as to why the g phase formed and the more stable a phase did not.
Wilkes (1979) has shown that phases which are equiatomic or nearly so have a high ordering
energy and have lower defect mobilities than the adjacent phases are particularly susceptible
to irradiation induced disordering. He notes that the tr W-Re phase satisfies the first two
conditions and probably the third. On this basis he suggests that irradiation-induced
disordering is responsible for the absence of the a phase in irradiated W-Re alloys.
If the solute segregation and irradiation disordering hypotheses are correct, the W-Re
results represent the intriguing combination of two mechanisms to alter phase stability. Solute
segregation produces Re-rich regions in the sample where precipitation is possible, and
irradiation-disordering prevents formation of the most stable phase.
Very recently Herschitz and Seidman (1984a, b) reported the results of atomic probe field
ion microscope studies of a neutron irradiated W-10a/oRe alloy. Irradiation in EBR-II was
to ~ 4 x 1022n/cm2 ( E > 0 . 1 M e V ) or 8.6dpa at 575, 625 and 675°C at an average
displacement rate of 1.4 × 10 -7 dpa/s. Extrapolation of the solvus to 575°C gave a solubility
of 16.5a/oRe, so the alloy was thermally single phase at all irradiation temperatures.
The FIM study revealed coherent, semi coherent, and possibly incoherent disc shaped
precipitates of tr phase which were 1-2 atomic planes thick and had a mean diameter of
~ 57/~. The precipitates formed at a number density of ~ 1016/cm3, and were not associated
with point defect sinks, so that some mechanism other than solute segregation was responsible
for their formation.
Herschitz and Seidman proposed that the particles nucleated in the interstitial-rich regions
surrounding displacement cascades by reaction of two mobile mixed dumbbells to form an
immobile Re cluster. Growth was hypothesized to occur by an irreversible vacancy-self
interstitial atom annihilation reaction of the type discussed in section 4.2.
It thus appears that in W-Re alloys irradiation induces the ~ phase at high temperatures
and the a phase at low temperatures. Different mechanisms may well be responsible for the
appearances of these irradiation induced phases.

15.15. Z r - A I Alloys
Zirconium-based alloys are of great interest in connection with pressure tubes in the
Canadian CANDU heavy water reactor, and have been considered as potential CTR first wall
materials.
Howe and Rainville (1977) studied the effect of sequential irradiation by 0.5, 1.0, 1.5, and
2.0 MeV Ar ÷ ions on an ordered Zr3A1 alloy of interest for CANDU pressure tubes. The
sequential irradiation with different energy ions produced a uniformly damaged region
between 2000 and 8000 ~ beneath the foil surface but at the expense of a displacement rate
which varied with both time and location in the sample. Displacement rates were not given
but were presumably in the 10-3-10 -4 dpa/s range characteristic of heavy ion irradiations.
Irradiations were between 30 and 693 K; the self-interstitial is mobile over the entire
temperature range, while the vacancy is mobile only at the highest temperatures, if at all.
Irradiation initially produced 20-80/~ diameter black spots which were interpreted as
regions disordered by cascades. At 285, 573, and 673 K a completely disordered state was
produced by ~ 2 dpa.
Irradiation to 2 dpa at 30 K or to 20 dpa at 300 K rendered the alloy amorphous. The
greater number of displacements needed at 285 K is apparently needed to reverse some
annealing which takes place at this temperature. Figure 110 shows the amorphized Zr3A1.
PHASE S T A B I L I T Y U N D E R I R R A D I A T I O N 397

15.16. Zr-Nb Alloys


These alloys are of particular interest because of their use in pressure tubes in CANDU
heavy water reactors.
Figure 111 shows the low temperature portion of the Zr-Nb phase diagram. Both fl-Zr
and fl-Nb are BCC, whereas ~-Zr is HCP. The metastable, coherent co-phase may form either
thermally or athermally over a range of alloy compositions. The athermal transformation
may occur below T~s, shown in Fig. 111. The isothermal solvus and the temperatures for the
beginning of athermal co-phase formation are those given by Bremer (1974), and the
equilibrium portion of the diagram is from Hansen (1958).
Nuttall and Faulkner (1977) made a study of the effects of 1 MV HVEM irradiation on
the microstructures of Zr-2.5w/oNb and Zr-12w/oNb alloys. The alloys were quenched from
850°C and aged at 500°C for 1000 hr to produce coarse fl-Nb precipitates in an ~-Zr matrix.
The Zr-12w/oNb alloy was water-quenched from 900°C and aged at 450°C to produce a fine
dispersion of hexagonal, athermally produced co-phase in a fl-Zr matrix.
The HVEM used in the experiment had a displacement rate of 5.4 x 10 -3 dpa/s. Irradiation
of the 2.5w/oNb alloy to ~ 12 dpa at 465°C produced a significant enlarging of the fl-Nb
particles in the ~-Zr matrix. Particles of fl-Nb in a sample heated for the same length of time
but not irradiated were unchanged. This result is illustrated in Fig. 112. Had the matrix been
depleted to the equilibrium solute concentration at the start of the irradiation, the experiment
would have demonstrated irradiation-enhanced coarsening, which may occur at the high dpa
achieved. If not, the particles may have grown due to irradiation-enhanced diffusion.
Irradiation of the Zr-12w/oNb alloy at 430°C gave a dispersion of fine ( ~ 50 A diameter)
co-phase particles between the pre-existing particles. After ~ 6 dpa, a significant fraction of
the original co-phase particles had disappeared, and the remainder were visible only
indistinctly, while the fine co-dispersion had increased in both number density and diffraction
intensity. These effects were interpreted as recoil resolution from the large co-particles,
followed by re-precipitation of a fine dispersion of particles, as predicted by the theoretical
analyses discussed in Section 7. No calculations were performed to support this recoil
resolution interpretation.
Hernandez and Potter (1981) bombarded a Zr-12.5w/oNb alloy with 3.0 MeV Ni ÷ ions
at 425°C to doses from 0.4-10.8 dpa at a displacement rate of ,~3.7 x 10 - 3 d p a / s . The
material had been aged at 450°C to produce a distribution of ~ 400 A diameter, cubiodal,
isothermal HCP co-phase particles in a BCC, fl-phase matrix.
Restructuring and re-nucleation effects of the type observed by Nuttall and Faulkner (1977)
in an alloy of the same composition and by other workers in Ni-A1 and Ni-Si alloys were
expected. Suprisingly, however, as seen in Fig. 113, irradiation altered neither the size,
morphology, nor spatial distribution of the co-phase particles. These results are in sharp
contrast to those of Nuttall and Faulkner (1977) who found that irradiation of a similar alloy
in the HVEM at 340°C resulted in refinement of the co-phase particle size.
Hernandez and Potter suggested that the difference between the results of HVEM and
heavy ion irradiation could be due to one of four causes:
(i) Nucleation could have been easier under electron irradiation due to the slightly lower
temperature, which should decrease the activation barrier for co-phase formation.
(ii) Nucleation occurred preferentially on the uniform distribution of vacancies produced
by electron irradiation.
(iii) Chemical effects in the thin ( ~ 1000 A) TEM foil interior may have occurred due to
irradiation induced segregation and could cause precipitate redistribution.
398 P R O G R E S S IN M A T E R I A L S SCIENCE

(iv) Displacement cascades during Ni + ion bombardment may have dissolved small, newly
nucleated o~-phase particles and prevented their growth to observable size.
More studies will be needed to determine the reason for the difference in results between the
two investigations.
Bremer (1974) studied the effect of room temperature, low fluence neutron irradiation on
the subsequent co transformation in Zr-15w/oNb and Zr-30w/oNb alloys. Irradiation was
in the Union Carbide research reactor at a fast neutron flux of 1.5 x 10~3/cm2.s (giving about
10 -s dpa/s).
Irradiation to 1019 fast n/cm 2 gave a large enhancement in the rate of formation of tn-phase
on subsequent heat treatment at 400°C. This enhancement was attributed to reduction of the
strain energy of the ~o-phase particles by irradiation-induced vacancies. Such an explanation
is reasonable, as the 1019n/cm2 fluence would produce about 10 -2 atomic fraction of
vacancies, which would far outnumber the thermal vacancy concentration at the low aging
temperatures used, and could be very helpful in relieving transformation strains.
Urbanic et al. (1975) studied the effects of neutron irradiation on Zr-2.5w/oNb alloys
which had been quenched from 870°C, cold worked to 15~ reduction in thickness, and given
aging treatments from 450-550°C for 100-12,000min. Specimens were irradiated with
5.8 × 1013n/cm2/s or 1.2 x 1014n/cm2/s (E > 1 MeV) fluxes to 3.3 x 1020 or 1 × 102~n/cm 2
(E > 1 MeV) or 0.2 or 0.5 dpa at a temperature of 270-280°C. Specimens were also aged for
the same time and at the same temperature in an out-reactor autoclave, for comparison.
Except for one fully annealed specimen, in-reactor corrosion rates were lower than
out-reactor rates, often by several fold. The low corrosion rates continued even after the
irradiated samples had their oxide layer removed and were exposed out of reactor. The
decrease in corrosion rate was attributed to fl-Nb precipitation on irradiation-induced defect
clusters. No additional fl-Nb precipitates were observed in electron microscope studies, but
the occurrence of the optimum corrosion resistance of this alloy where the Nb concentration
of ~-Zr is close to the equilibrium value led to the belief that the matrix Nb concentration
had decreased during irradiation.
Coleman et al. (1981) irradiated a Zr-2.5w/oNb alloy in the Oak Ridge reactor (see Fig.
3) at 570 K to 0.8 dpa, 670 K to 0.74 dpa, and 770 K to 0.62 dpa at a displacement rate of
approximately 1.4 x 10 -7 dpa/s. The starting material was either in the annealed state or in
the quenched and aged state. Annealed material had a microstructure of large HCP 0~-Zr
grains with BCC fl-Nb phase at the grain boundaries. The quenched and aged material had
a microstructure of ~-Zr grains surrounded by a matrix of ~' martensite. The ~ phase also
contained some globular fl-Nb precipitates.
Irradiation in the 570-770 K range resulted in precipitation of fl-Nb in the ~ phase for both
annealed and quenched and aged material. Figure 114 shows precipitation of the fl phase,
which appeared to nucleate homogeneously at 570 and 670 K and heterogenously at 770 K.
Nucleation of fl-Nb did not occur in thermally aged control samples.
The authors attributed the precipitation of fl-Nb to irradiation enhanced diffusion and
solute segregation of the undersized Nb atoms.

16. MISCELLANEOUS ALLOYS

16.1. M g - C d Alloys
Butler and Orchard (1981) included MgaCd in their study of irradiation disordering
described in Ni-based alloys. A [0001] oriented specimen of Mg3Cd was irradiated in the
PHASE STABILITY UNDER IRRADIATION 399

FIG. 106. Electron micrograph of TiO 2 precipitates in a V-20w/oTi alloy irradiated to 8 x 1022n/cm 2
(E > 0.1 MeV) at 650°C. (Electron micrograph courtesy J. A. Sprague.)
400 PROGRESS IN MATERIALS SCIENCE

3500

3000

2500
to
o

a; 2000 I i .ll
HCP

T t

/ , , II I
1000 ; I / I'
tl I / ,
J !l i I II I
50O I I II I
/ 'I I' Fxl
I
I
0 I I I I
0 10 2O 30 40 50 60 70 80 90 I00
Alomic Percent Rhenium Re

FIG. 107. Tungsten-Rhenium phase diagram showing alloy compositions and temperatures of
irradiation in study of Williams et al. (1983). (After Williams et al., 1983.)
PHASE S T A B I L I T Y U N D E R I R R A D I A T I O N 401

FIG. 108. Electron micrograph of W-25a/oRe neutron irradiated at 1500°C showing precipitate
particles adjacent to grain boundaries and precipitate-free regions within the grains. (Electron
micrograph courtesy R. K. Williams and Metallurgical Transactions.)
402 PROGRESS IN MATERIALS SCIENCE

FIG. 109. Electron micrograph of W-I la/oRe neutron irradiated at 1500°C. Three views were taken
with the diffraction vector rotated 90 ° between each. Figures (a, b, c) show that the strain fields
around the particles are spherically symmetrical. (Electron micrograph courtesy R. K. Williams and
Metallurgical Transactions.)
PHASE S T A B I L I T Y U N D E R IRRADIATION 403

FIG. 110. Zr3AI bombarded at 40 K with 1 x 10Is Ar ÷ ions/cm 2. (a) Bright field electron micrograph.
No features characteristic of crystallinity are observed; (b) Same area as (a), imaged in dark field
from arc of main diffuse diffraction ring; amorphous regions appear white. (c) Diffraction pattern
characteristic of amorphous state of Zr3AI. (Electron micrographs and diffraction pattern courtesy
L. M. Howe and Journal of Nuclear Materials.)
404 PROGRESS IN MATERIALS SCIENCE

1200 I
,O-Zr, NbI
I000 f ~

(,~
o
80G [ ~

~oo\ \ \ . /
,./
\
---., 8 (metastable)
400
~\u~+~ (metastable)
200 \
\ T~,,
0
\
-200
0 I0 20 30 40 SO 60 70 80 90 I00
Zt Atomic per cent Niobium Nb

FIG. 11 1. Z r - N b phase diagram with metastable ~o field added. The atherrnal to-phase transformation
may occur below the To,, line.
r~

,-]

t'-
.q

Z
~7

~7
-]

7r

FIG. 112. Electron micrographs of fl-Nb precipitates in a Zr-2.5w/oNb alloy (a) Without irradiation; and (b) after 60min HVEM irradiation
(approximately 20dpa) at 465°C. (Electron micrographs courtesy K . Nuttall.) t~
406 PROGRESS IN MATERIALS SCIENCE

0 dpa I000~ 0.4 dpa


I I

4.2 dpa 10.8 dpa


FIG. 113. Electron micrographs of Zr-12.5w/oNb aged 5 h at 450°C and irradiated with 3 MeV Ni +
ions at 425°C to various displacement levels. Foil orientations are near (110) in (a) and near (100)
in (b), (c) and (d). (Electron micrographs courtesy O. G. Hernandez and The Metallurgical Society
of AIME.)
PHASE S T A B I L I T Y UNDER IRRADIATION 407

FIG. 114. Electron mmrograph of Zr-2.5w/oNb irradiated 1350 hr at 770 K (0.62 dpa) in the Oak
Ridge reactor and given a 7 d post irradiation anneal at 770 K. The sword-shaped precipitates are
irradiation-induced B-phase. (Electron micrograph courtesy C. E. Coleman.)
408 P R O G R E S S IN MATERIALS SCIENCE

25 -- ~o-o.~

~1>1o-5~ /~" I J
500 I000 1500 2000
Time in minutes
FIG. 115. Percent volume change vs. time curve for the white-to-grey tin transformation following
reactor irradiation of the white tin in liquid nitrogen. Temperature of transformation reaction is
-50.3°C. Annealed pure white tin shows no transformation and would be represented by the
abscissa in the figure. (After Fleeman and Dienes, 1956.)
PHASE S T A B I L I T Y U N D E R I R R A D I A T I O N 409

Table 20. Phase stability in refractory alloys


Alloy Treatment Result References
M o - 1a/oTi 700°C, 1022n/cm 2 No Ti segregation Wagner and Seid-
(E > 1 MeV) ~ 5 dpa to/from voids, man (1979)
Mo-0.5w/oTi, 0.1w/oZr 650°C, 5.4 x 10 22 n/cm 2 Low density of glob- Sprague et al. (1979)
(E > 0.1 MeV), 22 dpa ular precipitates.
Mo-9. I a/oZr 700-900°C, up to 12dpa, Mo2Zr at grain Liou et al. (1979)
14 MeV Cu ÷ boundaries.
Nb3Sn T < 70°C, up to 8 x 1019 Drastically de- Karkin et al. (1976)
fast n/cm% ~ 0.04 dpa creased To, He2, and
degree of long range
order.
Nb-59.8a/oTi 25°C(?), 1.3 x 1018n/cm 2, Critical current and Tsubakihara et al.
10-3 dpa transition tempera- (1975)
ture (Jc and T,,) low-
ered
Ti-14.4a/oAl 600°C, 1.3-14.7 dpa, Heavy AI segre- Erck et al. (1979)
3 MeV Ni ÷, 4.7 x 10 -3 gation and Ti3A1
dpa/s precipitation at sinks.
Ti-8.7w/oAl 650°C, 2.1 MeV Ar ÷, Strong segregation Wang et al. (1982)
Ti-6w/oAI-4w/oV 8 × 10-4dpa/s, 0.75-3 dpa of Mo and V, mod-
Ti-8w/oAl-lw/oMo at surface erate segregation of
.Ti-3w/oV AI to defect sinks.
Ti- 14.4a/oA 1 690°C, 3 MeV Ni ÷, No change in ct2 par- Potter (1981)
1, 5, 12dpa ticle number density,
size or shape.
Ti-6.35w/oAl~,.0w/oV 725-825K, 5 MeV Ni ÷ No irradiation in- Brimhall et al.
3 x 10-3dpa/s, 2dpa duced segregation. (1981)
Ti-6w/oA 1-4w/oV 450-650°C, 2.4 MeV V ÷ or Profuse precipitation Agarwal et al. (1979)
2.4 MeV V ÷ and 0.95 MeV He ÷, of fl platelets in ~t
3-25 dpa matrix.
Ti-6.4w/oAl, 3.9w/oV, 395-550°C, EBR-II Precipitation of ~t2 at Peterson (1982)
various anneals 4.2 x 1022 to 5.0 x 1022 defect sinks, fl pre-
n/cm 2 (E > 0.1 MeV) 21-25 dpa cipitation.
Ti-8w/oA 1-1w/oV-lw/oMo 450, 550, 650°C, 2.4 MeV V + Dissolution or redis- Ayrault (1981)
or 2.4 MeV V ÷ and 0.95 MeV tribution of ct2, pre-
He +, 3-25dpa, 5 x l0 -3 cipitation of ft.
dpa/s
V - ~ 10wt.ppm B 70°C 5.4 x 1019n/cm 2 Concentric haloes of Rau and Ladd
(E > 1 MeV) ~0.02 dpa, He bubbles and of (1969)
3.5 x 1020n/cm 2 (thermal) LiV205 around cy-
lindrical V3B2 pre-
cipitates.
V-0.026 to 0.886a/oC 973 K, 200 keV C ÷, C segregates to irra- Takeyama et al.
48 dpa diated surface. (1982a)
V-0.886a/oC 773 K, 5 x 1020 fast Precipitation of C-V Takeyama et al.
n/cm2; 740 K, 1021 fast aggregates. (1982b)
n/cm 2, or ~0.25 and 0.5dpa
V- ~ 50wt.ppm C 600-750°C, 1-5 dpa at VC precipitates be- Weber et al. (1976)
3 x 10-4dpa/s by 18MeV low solubility limit.
Cu ÷ ions
V - ~ 120wt.ppm C 650-700°C, 3.25 MeV Ni ÷ VC precipitates be- Agarwal and Taylor
ions, up to 55 dpa low solubility limit. (1976a)
V- 15w/oCr 650°C, 5-60dpa, 3.5 MeV Heavy Cr segre- Agarwal et aL (1978)
V ÷ ions, 10-3dpa/s. gation to free sur-
face.
V- 15w/oCr 450-700°C, up to 60dpa, Heavy Cr segre- Rehn et aL (1979)
3.5 MeV V ÷, 10 3 dpa/s. gation to free sur-
face.
V-10w/oCr 496, 580°C, 1.5 x 1022 Fine unidentified Bentley and Wiffen
n/cm 2 (E > 1 MeV), precipitates. (1976)
10 dpa
410 PROGRESS IN M A T E R I A L S SCIENCE

Table 20. Continued


Alloy Treatment Result References
V-3a/oFe 740 K, 102. fast n/cm2; Fe segregates to Takeyama et al.
773 K, 5 x 102o fast grain boundaries. (1982b)
n/cm 2, or ~0.5-0.25 dpa
V-3a/oMo 740K, 1021 fast n/cm2; Mo segregates away Takeyama et al.
773K, 5 x 1020 fast from grain bound- (1982b)
n/cm 2, or ~0.5-0.25 dpa aries.
V-3a/oNb 740K, 102~ fast n/cm2; Nb segregates away Takeyama et al.
773K, 5 x 1020 fast from grain bound- (1982b)
n/cm 2, or ~0.5-0.25 dpa aries.
V-la/oNi 550, 650°C, 1-5 dpa at Ni segregation to Lott et al. (1979)
6 x 10-4 dpa/s with 14 MeV dislocations, signs of
Cu ÷ ions coherent precipitates
V-lw/oTi, 200ppmO 100°C, 3 MeV V ÷ ions, Fine precipitates Santhanam et al.
(by weight) 2.4-54dpa, 5.5 × 10 -3 (TiO?) (1975)
dpa/s
V3Si T < 150°C, 1-3 × 1019 90~ decrease in T,., Meier-Hirmer and
fast n/cm 2 103 increase in J,, Kiipfer (1982)
V-10w/oTi 650°C, 2.7 MeV V ÷ ions, Fine precipitates Agarwal and Taylor
2-60 dpa (TiO?) (1976b)
V-20w/oTi 650°C, 8.4 x 1022n/cm 2 Fine TiO 2 precip- Sprague et al. (1979)
(E > 0.1 MeV), ~40 dpa itates
W-25a/oRe 1000°C, 1022n/cm 2 Formation of Z Sikka and Moteff
(E > 1 MeV) 4.2 dpa (WRe3) (1974)
in single phase re-
gion
W-5, 11, 25a/oRe 600-1500°C, 4.3 x 1021 to Formation of Z R.K. Williams et al.
3.7 x 1022 n/cm 2 (WRe3) in single (1983)
(E > 0.1 MeV), 0.5-4.2 dpa phase region I100°C
and above
W - 10a/o Re 575, 625, 675°C, 10j6 thin a-phase Herschitz and Seid-
4 x 1022n/cm 2 (E >0.1 MeV) discs/cm 3, not asso- man (1984a, b)
8.6 dpa. ciated with point de-
fect sinks
Zr3AI 285-673°C, ~ 2 d p a , 0.5-2 Complete disor- Howe and Rainville
MeV Ar ÷ dering (1977)
Zr3AI 30 K, ~ 2 dpa or 300 K, Total amorphization Howe and Rainville
~20dpa, 0.5-2 MeV Ar ÷ (1977)
Zr-2.5w/oNb 465°C, 1 MV e-, 4.4 × 10 -3 Growth of /~-Nb Nuttall and Faulk-
dpa/s, 12dpa particles, ner (1977)
Zr-12w/oNb 340°C, 1 MV e-, 4.4 x 10_3 Large ~o-phase par- Nuttall and Faulk-
dpa/s, ~ 6 d p a ticles replaced by ner (1977)
fine ones.
Zr-12.5w/oNb 425°C, 3.0 MeV Ni +, 0.4-10.8 No effect on size, Hernandez and Pot-
dpa, 3.7 x 10 3 dpa/s morphology, or spa- ter (1981)
tial distribution of
400 ,~ ~o-phase par-
ticles.
Zr-15,30w/oNb ~25°C, 1019 fast n/cm 2 Enhanced formation Bremer (1974)
~ 5 x 10-3dpa of o~-phase on sub-
sequent heat treat-
ment.
Zr-2.5w/oNb 270-280°C, 3.3 × 102o or Increased corrosion Urbanic et aL (1975)
1 x 1021n/cm 2 (E > 1 MeV), resistance.
0.2-0.5 dpa
Zr-2.5w/oNb, annealed 570, 670, 770 K, 0.6-0.8 dpa Heterogeneous pre- Coleman et al.
or quenched and aged in ORR, ~1.4 × 10-Tdpa/s cipitation of fl-Nb at (1981)
770 K; homogeneous
precipitation at
lower temperatures.
PHASE STABILITY UNDER IRRADIATION 411

HVEM at 400 kV at temperatures of 133-193 K. The effective disordering cross section, A,


obtained from the initial portion of a InS versus t curve was found to decrease from
270 × 10 -~sm 2 at 133 K to 165 × 10-58m2 at 193 K.
The irradiation temperatures were low enough to eliminate any vacancy contribution to
disordering and the loss of order during irradiation was attributed to replacement collision
sequences. The decrease in the disordering cross section, A, with temperature was believed
due to the attenuation of the replacement sequence length due to lattice vibrations.

16.2. Pb-Ni Alloys


Wohofsky and Waidelich (1968) studied the effects of neutron irradiation on precipitation
in alloys containing from 0.1 to 0.5w/oNi. Nickel is only sparingly soluble in lead, so all the
alloys were supersaturated at the 50°C irradiation temperature. Irradiation was to 5 × 1018n/
cmz (E > 0.1 MeV) or about 2 × 10 -3 dpa, in a fast neutron flux of 9.5 × 1012n/cm2/s.
Small angle x-ray diffraction showed that during irradiation supersaturated single phase
alloys developed 20/~ diameter Ni-rich zones. However, irradiation reduced the size of
pre-existing zones in more concentrated alloys. Both effects were attributed to disturbances
by displacement cascades. However, the number of dpa is too low for recoil resolution to
produce a significant mass loss. Formation of the new, small particles could be due to either
irradiation-enhanced diffusion or to displacement cascades serving as nucleation sites.

16.3. Pd-Si Alloy


Klaumiinzer et al. (1982) irradiated foils of the metallic glass PdsoSiz0 with 25-250 MeV O,
Ar, and Kr ions, which passed entirely through the samples thereby avoiding implantation
effects. Foils were irradiated after cold working, cold working plus annealing at 520 K, and
after annealing. Irradiation was to up to 0.07 dpa at temperatures between 50 and 140 K.
Sample resistance was measured after various increments of irradiation and during post-
irradiation annealing.
The sample resistivity increased during irradiation, and decreased sharply on annealing at
temperatures just below 500 K, as might have been expected.
Sample length and width increased by up to 10Y/ooduring irradiation, but with an
accompanying decrease in thickness so that sample volume remained virtually constant. No
mechanism was proposed for these dimensional changes which are far greater than can be
obtained in crystalline metals and alloys at such low displacement levels.

16.4. P t - C Alloys
Westmacott and Perez (1979) studied precipitation in quenched Pt-C alloys. Their results
are included here because processes in quenched alloys sometimes provide insight into
behavior under irradiation.
Carbon atoms are ,,~37~ too large to fit into 0, 0, 1/2 octahedral interstitial sites in Pt
without strain. The resulting strain energy may be reduced by formation of a vacancy:carbon
(V-C) complex.
Samples of an impure Pt containing 800 at.ppm C and of a more pure Pt with only
80 at.ppm C were quenched from the range 1000-1760°C (T,, = 1774°C). After quenching,
the impure Pt contained planar defects which lay on {100} planes, had a Burgers vector and
fault displacement vector of l/3a(100), and were intrinsic in nature. The defects were stable,
412 PROGRESS IN M A T E R I A L S SCIENCE

and did not unfault even when they intersected a free surface. The defects were interpreted
as a disc of C atoms, which had formed by precipitation of V-C complexes. Such precipitation
would remove a single layer of Pt atoms and replace them with a layer of C atoms in what
are normally substitutional sites. Inward collapse of the neighboring planes toward the
smaller C atoms would give the observed Burgers vector.
The pure Pt gave the vacancy loops on {111} planes as is usually observed in quenched
FCC metals. Whereas the impure Pt had one C atom per vacancy even at the melting point,
the ratio in the pure Pt was 1 : 10 at the most, so that at least 90% of the vacancies were free
and able to precipitate on {111} planes with the associated low-energy stacking fault.
The precipitation of V-C complexes is an example of the mechanism of co-precipitation
of free vacancies and solute described in Section 4.1. The quench from high temperature to
ca. 400°C produced a huge vacancy supersaturation, which persisted even with complex
formation. At the lower temperature, supersaturations in both V and C contribute to the
thermodynamic driving force and help to reduce the activation energy for platelet nucleation.

16.5. Sn Alloys
Fleeman and Dienes (1955) used neutron irradiation to accelerate the white-to-grey Sn
transformation. The high temperature white tin may transform to a 27% less dense form
known as grey tin at temperatures below 13.2°C. The transformation is usually sluggish, and
may be prevented by quenching or by small amounts of impurities, or it may be enhanced
by cold working. The progress of transformation is easily followed by volume change
measurements.
The samples were irradiated to 10~8n/cm 2 at 78 K, which would give about 5 × l 0 - 4 dpa.
Transformation rates were measured between - 2 0 and -59.5°C, and showed a maximum
at -30°C; induction periods were the order of 100-500min. Figure 115 shows the
transformation kinetics at -50.3°C. Pure, annealed white Sn would have shown no
transformation at all during the times studied, but would ultimately have also shown a
maximum rate at -30°C.
Enhancement of the transformation was interpreted as being due to defects introduced
during irradiation serving as nucleation embryos. Such an explanation is reasonable, for
although only a small fraction of atoms were displaced, the order of 1017 displacement
cascades were created per cm 3. The cascade centers are vacancy-rich regions ~ 100/~ in
diameter, which could serve as effective nucleation sites for the less-dense grey Sn. The ability
of cold work also to enhance the transformation also shows that the transformation is
catalyzed by structural defects in the white Sn lattice.

16.6. U-AI Alloys


Phase stability studies in U-based alloys are mainly in the context of their use as metallic
nuclear fuels. The dimensional stability of nuclear fuels is strongly affected by their
microstructural stability, hence there is interest in irradiation-altered phase stability.
The irradiation dose received by nuclear fuels is customarily given in terms of megawatt
days per ton of fuel (MWD/T), fission events per unit volume, or percent burnup of all
U-atoms, rather than in dpa. Each fission event gives the order of 200 MeV, mostly in very
high energy fission fragments which will lose much of their energy in inelastic interactions
which do not displace atoms. Taking one displacement per 200 eV fission energy gives very
roughly 1 dpa for each MWD/T, or for 10 t7 fissions per cm 3, or 10-4% burnup.
PHASE STABILITY UNDER IRRADIATION 413

B. Hudson et al. (1974) studied the microstructures of ~ uranium fuels with 585 and
1250 ppm of aluminum. Discs of the material were fl-quenched into water from 993 K and
given a 1-hr anneal at 823 K before being placed in a reactor and burned at 670 K to
520 MWD/T, or about 500 dpa in 1.8 x 1 0 6 S, for a displacement rate of ,~ 3 × 10 -4 dpa/s.
The displacement rate in the fuel is thus closer to that in heavy ion or HVEM irradiations
than that in reactor cladding. It was noted that thermal anneals produce little additional
precipitation in similar materials.
The samples contained a high density of fine (300-400A,) UA12 precipitates before
irradiation. Irradiation produced a very high density of very small ( ~ 100 A) UAlz particles
in both high and low A1 materials without appreciably affecting the pre-existing precipitate
distribution.
The investigators noted that precipitation of UAI2 was not complete prior to irradiation
and that nucleation of UA12 particles is difficult under thermal conditions. They then
interpreted their results in terms of irradiation enhancing diffusion and in addition producing
heterogeneous nucleation sites for UA12. Such an interpretation is reasonable, although a
number of other mechanisms could be operative at 500 dpa.

16.7. U - F e alloys
Bloch (1962) studied the effect of irradiation on U6Fe precipitates in alloys containing from
1 to 3.8w/oFe. Irradiation was to up to ~ 0 . 1 ~ burnup of all U atoms at temperatures from
265 to 400°C. The 0.1~ burnup is equivalent to ~ 1 0 3 MWD/T or ,,~ 103dpa.
During irradiation the x-ray diffraction lines which corresponded to U6Fe progressively
diminished and then disappeared entirely without any change in the lattice parameter of the
-U. The disappearance of the U6Fe lines was interpreted as each fission event rendering a
volume of ~ 10-17cm 3 amorphous. The ,,~103dpa received at the highest burnups are
certainly enough to render a material amorphous in the absence of annealing of the
displacement damage. Amorphization of the U6Fe but not the ~-U implies that the latter
underwent some annealing during irradiation but the former did not. It is reasonable to
expect that an intermetallic compound such as U6Fe would be slower to anneal than U-metal,
due to the former being slower-diffusing.

16.8. U - M o Alloys
Bleiberg et al. (1956) and Konobeevsky et al. (1956) made in-reactor studies of U - M o
alloys. The former studied natural U-9, 10.5, 12, and 13.5 w/o Mo alloys which were in the
stable et-U + 7' (UzMo) or quenched y states before irradiation. Figure 116 gives the relevant
portion of the U - M o phase diagram. Irradiation was to up to 0.09~ burnup ( ~ 103 dpa) over
6 weeks, for a displacement rate of ~ 3 x 10-4/s, at temperatures of 70-200°C. Density, x-ray
diffraction, electrical resistivity, and hardness measurements all showed that the alloys had
reverted to the metastable high-temperature y phase during irradiation. However, optical
microscopy showed the microstructures of the alloys to be unchanged by the intense
irradiation.
These results were interpreted in terms of mixing by displacement cascades, which
disordered the U2Mo y' phase to form the y phase. Calculations based on diffusion from
cascade mixing were presented to support this interpretation.
The theory of irradiation-induced disordering presented in Section 8 predicts that the high
displacement rates and low irradiation temperatures in this study would favor disordering of
the y' phase.
414 PROGRESS IN M A T E R I A L S SCIENCE

Weight per cent Uranium


75 80 85 90 95 97 99
900

800
o +l, ol /
70C

~" 60C J

500 Ii

y'+ (~1o) [4 ../~ ,

,oo II i
50 60 70 80 90 I00
Atomic per cent Uranium U

FIG. 116. Uranium-rich end of U - M o phase diagram. (After Elliott, 1965.)

Konobeevsky et al. (1956) exposed a U-9w/oMo alloy in both the metastable 7 state and
in the stable 0t + ~' state to 1019-1020n/cm 2 at 50°C. The authors stated that the U had
received " . . . 75% enrichment..." but did not make clear whether the enrichment was by
or to 75%. If the former were the case the alloy would have received about the same
number of dpa as the alloys of Bleiberg et al. and if the latter, about an order of magnitude
more dpa.
Diffraction measurements showed the alloys in the 7 + ? ' state reverted partially or
completely to 7 on irradiation, whereas the 7 alloys remained in that state.

16.9. U - N b Alloys
The U-Nb and U-Mo phase diagrams (Hansen, 1958) are very similar, and show eutectoid
transformations at approximately the same temperature and composition.
Bleiberg et al. (1956) subjected partially transformed U-10w/oNb to the same irradiation
conditions as the U-Mo alloys described earlier. The results of post-irradiation studies with
density, x-ray diffraction, electrical resistivity, and hardness measurements all indicated
reversion to the high temperature 7 phase.
As in the case of U-Mo alloys, optical microscopy showed no microstructural change from
irradiation. Resistivity measurements, however, indicated that the reversion may not have
been as complete as was the case with U-Mo alloys.

17. SUMMARY
A material under irradiation is a driven, irreversible system with a continual input of
energy. Nonetheless, such a system may (or may not) reach a steady state which has some
of the characteristics of equilibrium.
Irradiation of metals produces large numbers of vacancies, self interstitials, highly energetic
atoms and, except for electron and ~,-ray irradiation, displacement cascades. These species
interact with a metallic microstructure in various ways which alter the stability of actual and
potential phases. To date, each effect has usually been analyzed individually, whereas the
actual microstructure is altered by a number of effects occurring simultaneously.

17.1. Diffusion
At temperatures below about 0.5 Tin, fast reactor or charged particle irradiations give
substantial increases in the diffusion coefficient of substitutional species. Typically, between
PHASE STABILITY UNDER IRRADIATION 415

Table 21. Phase stability in miscellaneous alloys


Alloy Treatment Result References
Mg3Cd in [0001] 133-193 K, 400 kV e- Disordering cross sec- Butler and Orchard
orientation HVEM tion, A, decreases with (1981)
increasing temperature

Pb4).2 to 0.5w/oNi 50°C, 5 x l0 ta n/cm2 Precipitation in super- Wohofsky and Wa-


(E > 0.1 MeV) ~ 2 × 10 -3 d p a saturated alloys; reduc- idelich (1968)
9.5×10Un/cm2-s. tion in size of pre-
existing precipitates
Glassy Pds0Si20; 50-140 K, 25-250MeV Up to 10~o increase in Klaumiinzer et al.
CW, CW and annealed, O, Ar, or Kr ions, up lateral sample dimen- (1982)
or annealed to 0.07 dpa sions at constant sample
volume.
Pt-800 appm C Quenched from 1000- Precipitation of {100} C Westmacott and Perez
1760°C platelets (1979)
Sn 78K, 10~8n/cm2, White to grey trans- Fleeman and Dienes
~ 5 x 10-4 dpa formation accelerated (1955)
on post-irradiation an-
neal
U-585 and 1250ppm AI 670K, 520MWD/T, Production of fine dis- Hudson et al. (1974)
~500dpa, 3x 10 -4 persion of UAI2 par-
dpa/s ticles
U-I to 3.8w/oFe 65-400°C, up to 0.1~/o U6Fe precipitates ren- Bloch (1962)
burnup, ~ 103 dpa dered amorphous
U-9 to 13.5w/oMo 70-200°C, up to 0.09~o Stable ct-U and ~,' Bleiberg et al. (1956)
burnup, ~103dpa, U2Mo phases revert to
~3 x 10 4dpa/s metastable y phase
U-9w/oMo 50°C, ca. 103 dpa(?) Stable c t - U and ),' Konobeevsky et al.
U2Mo phases revert to (1956)
metastable ), phase
U-10w/oNb 70-180°C, up to 0.09~ Stable ct-U and y'U2Nb Bleiberg et al. (1956)
burnup, ~103dpa, phases revert to metas-
~3 x 10-4dpa/s table ~ phase.

~ 0 . 5 Tm a n d ~ 0.25 Tin, at which t e m p e r a t u r e the v a c a n c y b e c o m e s i m m o b i l e , s u b s t i t u t i o n a l


diffusion coefficients are the o r d e r o f 10-16cm2/s. Such a coefficient is sufficient to give
m i c r o s t r u c t u r a l changes which w o u l d be u n o b s e r v a b l y slow in the absence o f i r r a d i a t i o n .

17.2. T h e r m o d y n a m i c Effects

T h e free energy o f a solid is altered b y v a c a n c y c o n c e n t r a t i o n s in excess o f the e q u i l i b r i u m


value, regardless o f h o w the vacancies arose. Several m o d e l i n g c a l c u l a t i o n s have been m a d e
on the basis o f vacancies acting as a t h i r d c o m p o n e n t o f the solid solution.
C o n d i t i o n s u n d e r which the presence o f excess vacancies alone was sufficient to give
s u b s t a n t i a l c h a n g e s in p h a s e stability are highly restricted. T o date, such changes have been
f o u n d in b i n a r y alloys o n l y when one o f the following obtains:

(i) Excess defects are a s s u m e d to be present in o n l y one o f two m i c r o s t r u c t u r a l phases


(ii) P o i n t defects o r d e r in one o f the p u r e c o m p o n e n t s a n d u n d e r i r r a d i a t i o n , a n d the
d e f e c t - i n d u c e d p r e c i p i t a t e c o n t a i n s some 20~o defects
(iii) A clustering system is at a t e m p e r a t u r e a n d c o m p o s i t i o n n e a r the t o p o f a miscibility
gap.
N o n e o f the three c o n d i t i o n s has as yet been r e p o r t e d as o b t a i n i n g in e x p e r i m e n t a l l y
o b s e r v e d cases o f i r r a d i a t i o n - a l t e r e d p h a s e stability.
J . P . M . S 28,3-4 1
416 PROGRESS IN MATERIALS SCIENCE

17.3. Point Defect: Particle Association


Particles incoherent with the matrix will, except when both phases have the same atomic
volume, either consume or emit vacancies to reduce the strain energy associated with their
formation. Consumption of vacancies from a matrix supersaturated in vacancies gives a large
decrease in free energy and helps to stabilize the particle. Calculations based on the model
AI-Ge system predict that irradiation will increase greatly the nucleation rates of intra-
granular Ge particles, as has been found experimentally. Irradiation is also predicted to
induce the nucleation and growth of incoherent intragranular particles at matrix solute
concentrations well below the thermal solubility limit.
Point defects cannot co-precipitate with particles coherent with the matrix due to the nature
of the particle:matrix interface. Irradiation may, nonetheless, stabilize coherent particles
through irreversible vacancy: self interstitial recombination at the particle:matrix interface.
In model calculations performed for the A1-Zn system irradiation is predicted to reduce the
solubility of the coherent particles to well below the thermal value. These calculations are
in agreement with observations of irradiation-induced precipitation in the (thermally) single
phase region.

17.4. Solute Segregation


Irradiation-induced solute segregation to or away from point defect sinks is a major
concern in fast fission or fusion reactor materials.
Three different bases for solute segregation have been used in theoretical developments.
The simplest of these, the misfit model, is based on strain energy considerations. Atoms with
a smaller atomic volume than the matrix are predicted to preferentially become part of
dumbbell interstitials and in that configuration migrate to sinks. Undersized atoms would
thereby be segregated toward point defect sinks and oversized atoms would be segregated
away from sinks. This model has had substantial success in predicting the direction of solute
segregation.
A more general analysis of solute segregation must consider not only solute:interstitial
binding, but also solute:vacancy interactions. A quantitative theory for solute segregation
in dilute solutions has been developed which considers these interactions and allows the
complexes to decay as well as form. Equations are written for the time rate of change of
concentration for each defect or complex, and then solved numerically for the alloy of choice.
Quantitative calculations are based on accurate values of the various defect complex-binding
energies, which are seldom known more than approximately. Where binding energies are
known, theory and experiment are in agreement. Solute concentration enrichment factors of
100-fold at sinks are commonly predicted and observed.
Theoretical treatment of solute segregation is simpler in concentrated than in dilute
solutions. In the former case, each atom sees an "average" surrounding, and an analysis in
terms of macroscopic diffusion coefficients may be made.
In the case where self interstitials are chemically neutral, the slowest diffusing species is
segregated toward the sink. A slightly more complicated criterion must be used when self
interstitials contribute to the segregation process.

17.5. Spinodal Instabilities


One of the main irreversible processes in irradiated metals is point defect recombination,
which occurs preferentially at such traps as coherent particle: matrix interfaces. The flux of
PHASE STABILITY UNDER IRRADIATION 417

self interstitials to the particle is not balanced by any reverse interstitial flux. As such, any
component which preferentially takes the interstitial position will be continually "pumped"
into the particle. The result is a spinodal-type composition instability for certain signs of the
phenomenological diffusion coefficients. Diffusion coefficients relating a flux in one com-
ponent to a concentration gradient in another are most important in determining the existence
of a spinodal-type instability. Model calculations for the AI-Zn system predict the existence
of instability over substantial portions of what is, thermally, a single phase region in the phase
diagram.

17.6. Recoil Resolution


Energetic atoms may destabilize precipitate particles by knocking atoms off into the matrix.
Several theoretical studies have been made of this phenomenon, each subject to a certain set
of assumptions and approximations.
The rate of atomic loss by recoil increases more rapidly with particle size than does the
rate of re-precipitation from the matrix. As such, all analyses predict a final stable particle
size, which will be approached whether the initial particle size was larger or smaller. This size
is typically in the tens of nm range. Analyses assuming infinite particle spacing find the stable
size to depend on the diffusion coefficient. The diffusion coefficient does not enter the
equations for particle growth or shrinkage in analyses based on Wigner-type cells around the
particles.

17.7. Disorder Dissolution


Point defect recombination effectively cancels the displacement event in disordered alloys.
Such is not the case in ordered materials, where many recombination events will put atoms
on the "wrong" sub-lattice. The resulting disordered material is less stable thermodynamically
than the equilibrium ordered phase.
The disordering and reordering processes scale with particle size as does recoil resolution,
so that a particular stable particle size is again predicted.
Phase diagrams have been calculated for ordering systems under irradiation. At high
temperatures the thermally activated reordering process is so rapid that the phase diagram
is unchanged by irradiation. At lower temperatures, irradiation preferentially destabilizes the
ordered phases. Disordered phases dominate more and more of the phase diagram until, at
sufficiently low temperatures and sufficiently high displacement rates, the ordered phase
disappears entirely.

17.8. Coarsening Kinetics


Irradiation may influence particle coarsening (Ostwald ripening) through disordering,
recoil resolution, and point defect recombination.
In the case of disordering, a system of particles was found to approach a particular stable
size by one of two rate laws. This size was approached from above as r versus t and from
below by r 3 v e r s u s t.
In the case of recoil resolution a final stable particle size is again approached. Particles
much larger than this size decay as r 2 versus t, whereas those much smaller grow by r 3 v e r s u s
t. Particles near the stable size approach that size by a law exponential in time.
418 PROGRESS IN M A T E R I A L S SCIENCE

Point defect recombination at the particle :matrix interface affects coarsening kinetics for
only a relatively narrow size range of particles. Particle distributions centered at about 5 nm
diameters coarsen more rapidly under irradiation than they would thermally. (The same
irradiation-enhanced diffusion coefficient is assumed in both cases.) After developing beyond
this size range the particle size distribution is unchanged by the effects of point defect
recombination.

17.9. Al-based Alloys


Research in Al-based alloys has focused largely on the response of GP zones and transition
precipitates to irradiation. Irradiation of A1-Cu solid solutions tends to promote formation
of 0' and depress formation of 0". Irradiation usually gives huge increases in precipitate
nucleation rates in A1-Ge and A1-Si alloys. Co-precipitation of solute with supersaturated
vacancies is probably the cause of the increase, although solute segregation may play a role.
Electron irradiation elevates the solvus temperature in A1-Zn alloys by tens of degrees
depending on the displacement rate.

17.10. Cu-based Alloys


Strong segregation effects occur in Cu-based alloys for Ag, Be, and Fe solutes. Irradiation
generally accelerates GP zone formation and precipitation in Cu-Be alloys, and causes loss
of coherency in Cu-Co and Cu-Fe alloys. Irradiation gives a resistivity decrease in Cu-Ni
alloys, which is attributed to decomposition of the solid solution.

17.11. Fe-based Alloys


Solute:interstitial and solute:vacancy interactions have been found to either increase or
decrease the precipitation rate in dilute Fe-C and Fe-N alloys depending on irradiation
conditions. Preliminary studies on martensitic stainless steels have shown that M6C, MozC,
X, Laves, M23C6, and ~' phases are formed under irradiation.
Phase stability in irradiated austenitic stainless steels has been studied intensively for over
a decade. Many changes in stability are attributed to large amounts of solute segregation of
Cr, Mo, Ni, and Si to point defect sinks. Neutron irradiation of type 316 stainless steel gives
7' and G phases, which do not appear thermally, enhances the formation of the q, Laves,
and MC phases, and also alters the composition of the Laves phase. The a and Fe2P phases
may also be affected by irradiation. The temperatures of appearance and abundances of these
phases are very sensitive to minor changes in alloy composition.
Neutron, heavy ion, and electron irradiation induce copious NbC precipitation in
Nb-bearing stainless steels. Co-precipitation of supersaturated vacancies with the NbC is the
probable cause of enhanced NbC precipitate nucleation.
Irradiation of types 316, 316L, 321, and 347 stainless steels results in the fGrmation of two
kinds of magnetic phases. Some irradiations give blocky ferrite particles, and others produce
a very high concentration of fine superparamagnetic particles, thought to be Ni and Fe-rich.

17.12. Ni-based Alloys


Strong solute segregation has been observed for AI, Be, Cr, Cu, Mn, Mo, Si, and Ti solutes
in Ni-based alloys. Precipitation resulting from such solute segregation has given apparent
inverse coarsening in Ni-A1 alloys.
PHASE S T A B I L I T Y UNDER IRRADIATION 419

Precipitation occurs in irradiated Ni-Ge and Ni-Be alloys which are thermally single phase
and a dose-rate dependent threshold for precipitation is observed in Ni-Si alloys.
Irradiation has been observed to produce disorder in several Ni-based alloys. A particularly
interesting observation is the disordering of Ni3Si, and its replacement by NisSi2 which is not
stable thermally, but apparently is more irradiation-resistant than Ni3Si. Electron irradiation
renders equiatomic Ni-Ti alloys amorphous.
Irradiation of partly crystalline and partly glassy Ni-Nb alloys drives the material back
toward the glassy state. Irradiation of Inconel 706 and 718 gives a complex interplay between
7', Y", and r/phases.

17.13. Refractory Alloys


Irradiation induces strong segregation of AI, Mo, and V in Ti-based alloys. Segregation
of A1 in irradiated Ti-AI alloys gives redistribution of ~2 precipitates and formation of fl
phase. Relatively low-fluence neutron irradiation of the superconductor V3Si gives a 90~o
decrease in T,. and a factor of 10 3 increase in Jc.
Neutron irradiation of thermally single phase W-rich W-Re alloys gives the X, WRe3 phase
rather than the equiatomic a phase. The proposed mechanism for Z phase formation is solute
segregation, giving local increases in the Re concentration so that precipitation may occur,
and irradiation disordering, preventing the a phase from forming.
The effect of irradiation on co phase precipitation in Zr-Nb alloys is ambiguous, in some
cases enhancing co phase formation and in other cases having no effect.

17.14. Miscellaneous Alloys


Neutron irradiation vastly increases the rate of white to grey Sn transformation, probably
by the vacancy-rich cores of displacement cascades acting as nucleation sites for the less dense
grey Sn.
High-fluence neutron irradiation of U - M o and U - N b alloys causes the stable ~-U and 7'
phases to revert to the metastable high temperature 7 phase. The reversion apparently occurs
by irradiation disordering of the y' phase to give y phase.

ACKNOWLEDGMENTS
I am grateful to the many authors who allowed me to study their papers prior to
publication; their courtesy has contributed to the timeliness of this review. This review is
based on work supported by the National Science Foundation (Grants No. DMR-77-23402
and DMR-80-21244), and by the U.S. Department of Energy (Contracts No. DE-ACO2-
78ET-52027 and DE-ACO2-78ER-10107). The support of these agencies over the extended
period of preparation made this work possible. I am grateful to Dr. M. Mruzik for his
assistance in the early stages of the literature search, and to Dr. K. Urban for a careful reading
of the manuscript. I am especially grateful to the many colleagues who provided the electron
micrographs which are essential to illustrate the experimental results.

REFERENCES
AARONSON, H. I. and RUSSELL, K. C. 0983) Nucleation--Mostly homogeneous and in solids, in Proceedings
of the International Conference on Solid-Solid Phase Transformations, (eds., H. I. Aaronson, D. E. Laughlin,
R. F. Sekerka and C. M. Wayman) TMS-AIME, Warrendale, PA, pp. 371-397.
420 PROGRESS IN MATERIALS SCIENCE

ADDA, Y., BEYELER, M. and BREBEC, G. (1975) Radiation effects on solid state diffusion, Thin Solid Films, 25,
107-156.
AGARWAL, S. C., AYRAULT,G., POTTER, D. I., TAYLOR, A. and NOLFI, F. V., Jr. (1979) Microstructure of single
and dual-ion irradiated Fe--20Ni-15Cr and Ti-6A1-4V Alloys, J. Nucl. Mater. 85 & 86, 653-657.
AGARWAL,S. C., REFIN, L. E. and NOLFI, F. V., Jr. (1978) Irradiation-induced void swelling and solute segregation
in a V-ion-irradiated V-15wt%Cr alloy, J. Nucl. Matter. 78, 336-342.
AGARWAL, S. C. and TAYLOR, A. (1976a) Dose dependence of void swelling in vanadium irradiated with self
ions, in Radiation Effects and Tritium Technology for Fusion Reactors, Vol. I, (eds. J. S. Watson and F. W.
Wiffen), National Technical Information Service, Springfield, VA, pp. 1-150-I-159.
AGARWAL,S. C. and TAYLOR,A. (1976b) Radiation enhanced precipitation in a V-10wt%Ti Alloy, Topical Meeting
on the Technology of Controlled Nuclear Fusion, Richland, WA, Conf. 760-935--p. 3, pp. 949-955.
ANTHOrC¢, T. R. (1972) Solute segregation and stresses generated around growing voids in metals, in
Radiation-induced Voids in Metals, (eds. J. W. Corbett and L. C. Ianniello) U.S. Atomic Energy Commission,
Washington, DC, pp. 630-645.
APPLEBY, W. K., SANDUSKY,O. W. and WOLFF, U. E. (1973) Swelling resistance of a high nickel alloy, J. Nucl.
Mater. 43, 213-218.
APPLEBY, W. K. and WOLFF, U. E. (1973) Effects of second phase particles on irradiation swelling of austenitic
alloys, in Effects of Radiation on Substructure and Mechanical Properties of Metals and Alloys, A S T M STP
529, American Society for Testing and Materials, pp. 122-136.
ARNDT, R. A. and DAMASK, A. C. (1964) Kinetics of carbon precipitation in irradiated iron-I/I, Acta Met. 12,
341-345.
AYRAULT, G. (1981) Precipitation in single and dual-ion irradiated Ti-8Al-lV-1Mo, in Phase Stability During
Irradiation, (eds. J. R. Holland, L. K. Mansur and D. I. Potter) TMS-AIME, Warrendale, PA., 577-586.
BANERJEE,S. and URBAN,K. (1984) Kinetics of order~lisorder transformation in alloys under electron irradiation,
Phys. Stat. Sol. (a) 81, 145-162.
BANERJEE, S., URBAN, K. and WILKINS, M. (1983) The transition from the short-range to the long-range ordered
state in Ni4Mo under electron irradiation, in Proceedings of International Conference on Solid-Solid Phase
transformations, (eds. H. I. Aaronson, D. E. Laughlin, R. F. Sekerka and C. M. Wayman) TMS-AIME,
Warrendale, PA, pp. 311-315.
BANERJEE, S., URBAN, K. and WILKINS, M. (1984) Order~disorder transformation in Ni4Mo under electron
irradiation in a high voltage electron microscope, Acta Met. 32, 299-311.
BARBU, A. (1980) Precipitation induite et gonflement dans la solution solide sous saturee Ni-6~oat.Ge irradiee
aux electrons de l MeV, in Irradiation Behavior of Metallic Materials for Fast Reactor Core Components, (ed.
J. Poirier and J. M. Dupouy) CEA-DMECN, Gif-sur-Yvette, France, pp. 69-73.
BARBU, A. and ARDELL, A. J. (1975) Irradiation-induced precipitation in Ni-Si alloys, Scripta Met. 9, 1233-1237.
BARBU, A. and MARTIN, G. (1977) Radiation-induced precipitation in nickel silicon solid solutions: II.--Dose
rate effects, Scripta Met. ll, 771-775.
BARBU, A., MARTIN, G. and CHAMBEROD,A. (1980) Low flux radiation induced precipitation, J. Appl. Phys. 51,
6192-6196.
BARON, J. L., CADALBERT, R. and DELAPLACE, J. (1974) Mechanisme possible de formation d'une phase
ferromagnetique dans l'acier inoxydable Fe/Cr/Ni/Mo, 17/11/8/2 (Type 316L) irradi6 en pile, J. Nucl. Mater.
51, 266-268.
BARON, M. (1981) Kinetic models for irradiation-induced precipitate stability, in Phase Stability during Irradiation,
(eds. J. R. Holland, L. K. Mansur and D. I. Potter) TMS-AIME, Warrendale, PA, pp. 63-72.
BARON, M., CHANG, A. and BLE1BERG, M. L. (1977a) Theory of particle redistribution in an irradiation
environment, in Radiation Effects in Breeder Reactor Structural Materials, (eds. M. L. Bleiberg and J. W.
Bennett) TMS-AIME, New York, pp. 395404.
BARON, M., CHANG,A. and BLE1BERG,M. (1977b) Theory of particle redistribution in an irradiation environment,
Westinghouse Electric Corp. Report WARD-AD.3045-6, Madison, PA, 21 pp.
BARTELS, A., DWORSCHAK,F., MEURER, H-P., ABROMEIT,C. and WOLLENBERGER,H. (1979) Be segregation in
electron irradiated dilute Cu/Be alloys, J. Nucl. Matter. 83, 24-34.
BEELER, J. R., Jr. (1966) Displacement spikes in cubic metals. 1. or-iron, copper, and tungsten, Phys. Rev. 150,
470-487.
BELL, W. L. and LAURITZEN,T. (1982) Microstructural changes in neutron irradiated commercial alloys: A sequel,
in Effects of Radiation on Materials, Eleventh Conference, ASTM STP 782, (eds, H. R. Brager and J. S. Perrin)
Amer. Society for Testing and Materials, Philadelphia, pp. 139-151.
BELLER, M. (1973) Der Einfluss yon Leerstellen auf die Keimbildung inkoh/irenter Germaniumausscheidungen in
Aluminum-Germanium-Legierungen, Z. Metal/k, 64, 189-193.
BENTLEY,J. and WIFFEN, F. W. (1976) Swelling and microstructural changes in irradiated vanadium alloys, Nucl.
Tech. 30, 376--384.
BERTRAM, K., MINTER, F. J., HUDSON, J. A. and RUSSELL, K. C. (1978) Irradiation enhanced precipitation in
AI-Ge alloys, J. Nucl. Mater. 75, 42-51.
BEST, S. E. (1984) Thesis research in progress, M.I.T.
PHASE S T A B I L I T Y UNDER IRRADIATION 421

BHATTACHARYYA, S. K. and RUSSELL, K. C. (1972) Activation energies for the coarsening of compound
precipitates, Met. Trans. 3, 2195-2199.
BILSBY, C. F. (1975) A theoretical examination of the effect of irradiation-enhanced dissolution on diffusion
controlled coarsening kinetics, J. Nucl. Mater. 55, 125-133.
BissoN, A. and VOU1LLIN,M. (1970) Electron microscope study of needle shaped precipitates in irradiated stainless
steels, EURFNR-891, Euratom Fast Reactor Exchange Program, United States, 12 pp.
BLE1BERG, M. L., JONES, L. J. and LUSTMAN, B. 0956) Phase changes in pile-irradiated uranium-base alloys,
J. Appl. Phys. 27, 1270-1283.
BLOCH, J. 0962) Effect de l'irradiation par les neutrons sur les alliages uranium-fer a faible teneur en fer,
J. Nucl. Mater. 6, 203-212.
BLOOM, E. E. and STIEGLER,J. O. (1973) Effect of irradiation on the microstructure and creep-rupture properties
of type 316 stainless steel, in Effects of Radiation on Substructure and Mechanical Properties of Metals and Alloys,
ASTM STP 529, American Society for Testing and Materials, Philadelphia, pp. 360-382.
BOCQUET, J-L. and MARTIN, G. (1979) Irradiation-induced precipitation: A thermodynamical approach, J. Nucl.
Mater. 83, 186-199.
BOLTAX, A. (1956) Effects of radiation damage on precipitation hardening alloys with special reference to
copper-iron alloys, ASTM STP 208, pp. 183-190.
BOLTAX, A. (1965) Displacement spike effects in copper-rich copper-iron alloys, Nucl. Applic. l, 337-347.
BOOTHBY, R. M. and WILLIAMS,T. M. (1981) Irradiation-induced transformations and orientation relationships
in a 12Cr-13Ni steel, J. Nucl. Mater. 96, 64-70.
BRAGER, H. R. and GARNER, F. A. (1979) Dependence of void formation on phase stability in neutron irradiated
type 316 stainless steel, in Effects o f radiation on Materials, ASTM STP 683, American Society for Testing
and Materials, Philadelphia, pp. 207-232.
BRAGER, H. R. and GARNER, F. A. (1981a) Microchemical evolution of neutron-irradiated stainless steel, in Effects
of Radiation on Materials: Tenth Conference, ASTM STP 725, (eds, D. Kramer, H. R. Brager and J. S. Perrin)
American Society for Testing and Materials, Philadelphia, pp. 470-483.
BRAGER, H. R. and GARNER, F. A. (1981b) Radiation-induced evolution of the austenite matrix in silicon-modified
AISI 316 alloys, in Phase Stability During Irradiation, (eds, J. R. Holland, L. K. Mansur and D. I. Potter)
TMS-AIME, Warrendale, PA, pp. 219-235.
BRAGER, H. R. AND GARNER, F. A. (1981c) Comparison of the swelling and the microstructural/microchemical
evolution of AISI 316 irradiated in EBR-II and HFIR, J. Nucl. Mater. 103 & 104, 993-998.
BRAGER, H. R. and GARNER, F. A. (1982) Influence of neutron spectra on the radiation-induced evolution of
AISI 316, J. Nucl. Mater. 108 & 109, 347-358.
BRAGER, H. R. and GARNER, F. A. (1983a) Microstructural and microchemical comparisons of AISI 316 irradiated
in HFIR and EBR-II, J. Nucl. Mater. 117, 159-176.
BRAGER, H. R. and GARNER, F. A. (1983b) The microchemical evolution and swelling of AISI 316 irradiated
in HFIR and EBR-II, in Proceedings of International Conference on Dimensional Stability and Mechanical
Behavior o f Irradiated Metals and Alloys, Vol. II, British Nuclear Energy Society, Brighton, pp. 1-4.
BRAGER, H. R. and GARNER, F. A. (1983c) Radiation-induced evolution of Fe--Ni--Cr ternary alloys, in Damage
Analysis and Fundamental Studies, DOE/Er-O046/12, U.S. Dept. of Energy, Washington, DC, pp. 170-177.
BRAGER, H. R. and GARNER, F. A. (1983d) Stability of radiation induced ~,' phase in 316 stainless steel, J. Nucl.
Mater. 116, 267-271.
BRAGER, H. R. and GARNER, F. A. (1984) Microsegregation observed in Fe-35.5Ni-7.5Cr irradiated in EBR-II,
in Effects of Radiation on Materials, ASTM STP 870 (eds. J. Perrin and J. Koziol). American Society for Testing
and Materials, Philadelphia (in press).
BRAGER, H. R. and S~AALSUND, J. L. (1973) Defect development in neutron irradiated type 316 stainless steel,
J. Nucl. Mater. 46, 134-158.
BRAGER, H. R., STRAALSUND,J. L., HOLMES, J. J. and BATES, J. F. (1971) Irradiation-produced defects in austenitic
stainless steel, Met. Trans. 2, 1893-1904.
BRAILSEORD, A. D. (1980) Precipitate re-solution in low dose irradiations, J. Nucl. Mater. 91, 221-222.
BRA1LSFORO, A. D. and BULLOUGH, R. (1972) The rate theory of swelling due to void growth in irradiated metals,
J. Nucl. Mater. 44, 121-135.
BRASKI, D. N. (1981) The resistance of (Fe,Ni)3V long-range ordered alloys to neutron and ion irradiation,
J. Nucl. Matter. 103 & 104, 1199-1204.
BREMER, B. W. (1974) The effect of fast neutron irradiation upon the omega transformation process in zirconium-
niobium alloys. Sc.D Thesis, New York University School of Engineering and Science, 176 pp.
BRIMHALL, J. L., BAER, D. R. and JONES, R. H. (1981) Radiation induced segregation in candidate fusion reactor
alloys, J. Nucl. Mater. 103 & 104, 1379-1384.
BROWN, B. S., FREYHARDT, H. C. and BLEWITT, T. H. (eds) (1978) Radiation Effects on Superconductivity, J. Nucl.
Mater. 72, 300 pp.
BROWN, C. (1975) Voids and rod-shaped features in 316 stainless steel irradiated to low neutron doses in DFR,
in Consultant Symposium on the Physics of lrradiation Produced Voids, Report AERE-R7934, (ed. R. S. Nelson)
HMSO, London, pp. 83-89.
422 PROGRESS IN MATERIALS SCIENCE

BROWN, L. M., WOOLHOUSE, G. R. and VALDRE, U. (1968) Radiation-induced coherency loss in a Cu-Co alloy,
Phil. Mug. 17, 781-789.
BULLOUGH, R. and WILLIS, J. R. (1975) The stress-induced point defect-dislocation interaction and its relevance
to irradiation creep, Phil. Mug. 31, 855-861.
BUTLER, E. P. and ORCHARD, J. F. (1981) An experimental investigation of electron-irradiation-induced disordering
mechanisms, in Phase Stability during Irradiation, (eds, J. R. Holland, L. K. Mansur and D. I. Potter)
TMS-AIME, Warrendale, PA, pp. 315-327.
BYSTROV, L. N., IVANOV, L. I. and PLATOV. Yu.M. (1968) Disintegration of the super-saturated solid solution
Cu + 2.5 per cent Be on exposure to electron irradiation, Fiz. Metal. Metalloved, 25, 950-953.
CAHN, J. W. and MULLINS, W. W. (1962) Discussion of a paper by J. W. Kirkaldy, in Decomposition of Austenite
by Diffusional Processes, (eds, V. Zackay and H. I. Aaronson), Interscience, New York. pp. 123-129.
CARPENTER, R. W. and KENIK, E. A. (1977) Stability of chemical order in Ni4Mo alloy under fast electron
irradiation, in 35th Annual Proc. Electron Microscopy Society of America, (ed. G. W. Bailey) pp. 48-49.
CARPENTER, R. W. and Yoo, M. H. (1978) The Effect of Semicoherent precipitation on void swelling in A1-Cu
alloys, Met. Trans. 9A, 1739-1748.
CAUVIN, R. and MARTIN, G. (1979) Radiation induced homogeneous precipitation in undersaturated solid
solutions, J. Nucl. Mater. 83, 67-78.
CAUVIN, R. and MARTIN, G. (1981a) Solid solutions under irradiation. I. A model for radiation induced
metastability, Phys. Rev. B, 23, 3322-3332.
CAUVIN, R. and MARTIN, G. (1981b) Solid solutions under irradiation. II. Radiation-induced precipitation in
AI-Zn undersaturated solid solutions, Phys. Rev. B, 23, 3333-3348.
CAUVIN, R. and MARTIN, G. (1982) Solid solutions under irradiation. III. Further comments on the computed
solubility limit. Phys. Rev. B., 25, 3385-3388.
CAUVIN, R. and MARTIN, G. (1983a) Radiation induced solid solution destabilization, in Proceedings of
International Conference on Solid-Solid Phase Transformations, (eds, H. I. Aaronson, D. E. Laughlin, R. F.
Sekerka and C. M. Wayman). TMS-AIME, Warrendale, PA, pp. 287-291.
CAUVIN, R. and MARTIN, G. (1983b) Radiation induced solid solution instability, in Proceedings of International
Conference on Solid-Solid Phase Transformations, (eds, H. I. Aaronson, D. E. Laughlin, R. F. Sekerka and
C. M. Wayman) TMS-AIME, Warrendale, PA, pp. 281-286.
CAWTHORNE, C., FULTON, E. J., BRAMMAN,J. I., LINEKAR, G. A. B. and SHARPE, R. M. (1971) Electron microscope
observations of voids in cladding materials irradiated in DFR, in Voids Formed by Irradiation of Reactor
Materials, British Nuclear Energy Society European Conference, Reading, pp. 35-43.
CERASARA, S., FEDERIGHI,T. and P1ERAGOSTINI,F. (1964) Pre-precipitation rate in AI-10%Zn alloy neutron
irradiated at 78 K, Phil. Mug. 9, 623-633.
CHAMBEROD, A., BARRUEL, F. and PAULEV~, J. (1971) l~tude par anisotropie magnetique du debut de la mise en
ordre d'un alliage Fe-Ni 50-50 irradi6 aux neutrons, J. Phys. Chem. Solids, 32, 881-887.
CHANG, A. L. and BARON, M. (1979) Particle redistribution and phase stability in ion and neutron irradiated
gamma prime strengthened Fe--Cr-Ni based alloys, J. Nucl. Mater. 83, 214-222.
CHERNOCK, R. S. and RUSSELL, K. C. (1984) Phase stability in rapidly cooled Ni-Nb alloys under Ni z+ ion
irradiation, Acta Met. 32, 521-527.
CLAUDSON,T. T., BARKER,R. "~,V.and FISH, R. L. (1970) The effects of fast flux irradiation on the mechanical
properties and dimensional stability of stainless steel, Nucl. Appl. Tech. 9, 10-23.
CLINARD, F. W., Jr. and HoBas, L. W. (1984) Radiation effects in non-metals, in Physics of Radiation Effects
in Crystals, (eds. R. A. Johnson and A. N. Orlov) North-Holland, Amsterdam (in press).
COLEMAN, C. F., GILaERT, R. W., CARPENTER, G. J. C. and WEATHERLY, G. C. (1981) Precipitation in
Zr-2.5wt%Nb during neutron irradiation, in Phase Stability During Irradiation, (eds. J. R. Holland, L. K.
Mansur and D. I. Potter) TMS-AIME, Warrendale, PA, pp. 587-599.
COMPRELLI, F. A. and LEWIS, J. E. (1965) Microstructural evaluation of superheat cladding materials, General
Electric Co. Report, GEAP-4751, (Jan. 1965).
COOK, H. E. (1970) Brownian motion in spinodal decomposition, Acta Met. 18, 297-306.
CORBETT, J. W. and IANNIELLO, L. C. (Eds.) (1972) Radiation-lnduced Voids in Metals, U.S. Atomic Energy
Commission, Washington, D.C., 884 pp.
DAMASK, A. C., CHOW, J. G. Y., KELSCH, J. J. and WAGENBLAST, H. (1970) Precipitation induced by neutron
irradiation of Fe-C alloys, Phil. Mug. 22, 549-562.
DENBIGH, K. G. (1951) The Thermodynamics of the Steady State, Methuen, London, 103 pp.
DIENES, G. J. and VINEYARD,G. H. (1957) Radiation Effects in Solids, Chap. 2, pp. 6-55, Interscience, New
York.
ELLIOTT, R. P. (1965) Constitution of Binary Alloys, First Supplement, McGraw Hill, New York, 877 pp.
LACK, R. A., POTTER, D. I. and WIEDERSICH, H. (1979) Void formation and solute segregation in Ti-14.4at.~oAl,
J. Nucl. Mater. 80, 120-125.
FARRELL,K., BENTLEY~J. and BRASK1,O. N. (1977) Direct observation of radiation-induced coated cavities, Scripta
Met. 11, 243-248.
FARRELL, K., MAZ1ASZ, P. J., LEE, E. H. and MANSUR, L. K. (1983) Modification of radiation damage
microstructure by helium, Rad. Effects 78, 277-295.
PHASE S T A B I L I T Y U N D E R I R R A D I A T I O N 423

FISH, R. L., HOLMES, J. J. and LEGGETT, R. D. (1970) Burst strength of EBR-II irradiation fuel pin sections.
Nucl. Appl. Tech. 9, 528-535.
FISHER, S. B. and WILLIAMS, K. R. (1972) Irradiation enhanced precipitation in stainless steel, Phil. Mag. 25,
371-380.
FLEEMAN, J. and DIENES, G. J. (1955) Effect of reactor irradiation on the white-to-grey tin transformation, J.
Appl. Phys. 26, 652-654.
Fowler, R. H. (1966) Statistical Mechanics, 2d edn. Cambridge Univ. Press, England, 864 pp.
FROST, H. J. and RUSSELL,K. C. (1982a) Recoil resolution and particle stability under irradiation, J. Nucl. Mater.
103 & 104, 1427-1452.
FROST, H. J. and RUSSELL, K. C. (1982b) Particle stability with recoil resolution, Acta Met. 30, 953-960.
FROST, H. J. and RUSSELL,K. C. (1983) Precipitate stability under irradiation, in Phase Transformations and Solute
Redistribution in Alloys during Irradiation, (ed. F. V. Nolfi) Res Mechanica, Applied Science Publ., London
and New York, pp. 75-113.
GARNER, F. A. (1981a) The microchemical evolution of irradiated stainless steels, in Phase Stability during
Irradiation, (eds J. R. Holland, L. I. Mansur and D. I. Potter) TMS-AIME, Warrendale, PA, pp. 165-189.
GARNER, F. A. (1981b) Extrapolation of stress-affected swelling models into compressive and cyclic stress states,
in Damage Analysis and Fundamental Studies, DOE/ER-0046/5, U.S. Dept. of Energy, Washington, DC, pp.
198-218.
GARNER, F. A. (1983) Dependence of swelling on nickel and chromium content in Fe-Ni-Cr ternary alloys, in
Damage Analysis and Fundamental Studies, DOE/ER-0046/14, U.S. Dept. of Energy, Washington, DC, pp.
133-151.
GARNER, F. A. (1984) Recent insights on the swelling and creep of irradiated alloys, in Third Topical Meeting
on Fusion Reactor Materials, J. Nucl. Mater. 122 & 123, 459-471.
GARNER, F. A., GILBERT,E. R., GELLES,D. S. and FOSTER,J. P. (1982) Effect of temperature changes on swelling
and creep of AISI 316, in Effects of Radiation on Materials, Tenth Conference, A S T M STP 725, (eds. D. Kramer,
H. R. Brager and J. S. Perrin) American Society for Testing and Materials, Philadelphia, pp. 698-712.
GARNER, F. A. and PORTER, D. L. (1983) History dependence and consequences of the microchemical evolution
of AISI 316, in Effects of Radiation on Materials, Eleventh Conference, A S T M STP 782, (eds. H. R. Brager
and J. S. Perrin) American Society for Testing and Materials, Philadelphia, pp. 295-309.
GARNER, F. A. and WOLFER, W. G. (1984) Factors which determine the swelling behavior of austenitic stainless
steels, in Third Topical Meeting on Fusion Reactor Materials, d. Nucl. Mater. 122 & 123, 201-206.
GELLES, D. S. (1979) Solute segregation to point-defect sinks in neutron-irradiated nimonic PE-16, J. Nucl. Mater.
83, 200-207.
GELLES, D. S. (1981a) Microstructural examination of several commercial ferritic alloys irradiated to high fluence.
J. Nucl. Mater. 103 & 104, 975-980.
GELLES, D. (1981b) Gamma prime coarsening and redistribution in nimonic PE-16, in Effects of Radiation on
Materials, Tenth Conference, A S T M STP 725, (eds. D. Kramer, H. R. Brager and J. S. Perrin) American Society
for Testing and Materials, Philadelphia, pp. 562-582.
GELLES, D. S. (1983) Precipitation during irradiation, an experimental example, in Proceedings of International
Conference on Solid-Solid Phase Transformations, (eds. H. I. Aaronson, D. E. Laughlin, R. F. Sekerka and
C. M. Wayman) TMS-AIME, Warrendale, PA, pp. 293-298.
GELLES, D. S. and GARNER, F. A. (1979) An experimental method to determine the role of helium in neutron-
induced microstructural evolution, d. Nucl. Mater. 85 & 86, 689-694.
GLOBS, J. W. (1960) The Scient~c Papers of J. Willard Gibbs, Vol. l', Dover Publ., New York, 434 pp. (reissue
of 1906 Longmans, Green edition).
GITTUS, J. H. and MIODOWNIK, A. P. (1979) Predicting the effects of radiation upon the constitution of fusion
reactor materials, J. Nucl. Mater. 85 & 86, 621-625.
GITTUS, J. H. and WATKIN, J. S. (1977) A test of the hypothesis that when austenitic alloys are bombarded with
energetic particles, those having the greatest thermodynamic stability will also have the greatest dimensional
stability, J. Nucl. Mater. 64, 300-302.
GR1NCHUK, P. P. and KIRSANOV,V. V. (1974) Dynamic destruction of Guinier-Preston zones in the process of
irradiation, Fiz. Metal. Metalloved. 38, 756-765.
GRINCHUK, P. P. and LOBODA, YE. M. (1974) Influence of neutron bombardment on decomposition of a
supersaturated solid solution of beryllium in nickel, Fiz. Metal. Metalloved. 38, 329-336.
GUINAN, M. W., KINNEY,J. H., VAN KONYNENBURG,R. A. and DAMASK,A. C. (1981) Fusion neutron disordering
of Cu3Au, J. Nucl. Mater. 103 & 104, 1217-1220.
HAKEN, H. (1975) Cooperative phenomena in systems far from thermal equilibrium and in non-physical systems,
Rev. Mod. Phys. 47, 67-121.
HARR1ES, D. R., BOLLOUGH, R., CAWTHORNE, C., MOSEDALE, D., NELSON, R. S. and STANDRING, J. (1971)
Irradiation effects in fast reactor cladding and structural materials, in Fourth United Nations Conference on the
Peaceful Uses of Atomic Energy, Conf. 710901-294, pp. 1-14.
HAUBOLDH. G. and MARTINSON,n. (1978) Structure determination of self interstitials and investigation of vacancy
clustering in copper by diffuse x-ray scattering, J. Nucl. Mater. 69 & 70, 644-649.

J . P M S 28/3-4~J
424 PROGRESS IN M A T E R I A L S SCIENCE

HAUSER, O. and SCHENK, M. (1967) A thermodynamic treatment of radiation-induced phase transformations, and
its application to some perovskites, in Interactions o f Radiation with Solids, Proceedings of the Cairo Solid State
Conference o f Sept. 3-8 1966.. (ed. A. Bishay). Plenum Press, New York, pp. 429-435.
HEALD, P. T. and SPEIGHT, i . V. (1975) Point defect behavior in irradiated metals, Acta Met. 23, 1389-1399.
HERMAN, H. (1964) The effects of irradiation on the formation of Guinier-Preston zones, Acta Met. 12, 765-774.
HERNANOEZ, O. G. and POTTER, D. I. (1981) Stability of Zr-Nb omega phase during ion bombardment, in Phase
Stability During Irradiation, (eds. J. R. Holland, L. K. Mansur and D. I. Potter) TMS-AIME, Warrendale,
PA, Do. 601-612.
HERSCmTZ, R. and SEIDMAN,D. N. (1984a) An atomic resolution study of homogeneous radiation-induced
precipitation in a neutron-irradiated W-10at~oRe alloy, Acta Met. 32, 1141-1154.
HERSCHITZ, R. and SEIDMAN, D. (1984b) An atomic resolution study of radiation-induced precipitation and solute
segregation effects in a neutron-irradiated W-25at%Re alloy, Acta Met. 32, 1155-1171.
HIGGINS, P. R. B. and ROBERTS, A. C. (1966) Carbide precipitation and associated tensile behavior in a neutron
irradiated stainless steel, J. Iron & Steel Inst., May, 489-494.
HORAK, J. A. (1968) Pre-precipitation rate in AI-10%Zn alloy neutron irradiated at 78 K, Phil. Mag. 17, 643-646.
HORAK, J., BLEWITT, T. H. and FINE, M. E. (1968) Effects of neutron irradiation at 4.5 K on Guinier-Preston
zone formation in aluminum-zinc alloys, J. Appl. Phys. 39, 326-335.
HOWE, L. M. and RAINVILLE, M. H. (1977) A study of the irradiation behavior of ZrDAl, J. Nucl. Mater. 68,
215-234.
HUDSON, B., JACKSON, J. P. and CODD, I. (1974) The effect of neutron irradiation on the precipitate distribution
in adjusted uranium, J. Nucl. Mater. 52, 229-240.
HUDSON, J. A. (1978) Precipitation under irradiation, in Precipitation Processes in Solids, (eds. K. C. Russell and
H. I. Aaronson) TMS-AIME, New York, pp. 284-313.
HULL, n . and MOGFORD, [. L. (1961) Precipitation and irradiation hardening in iron, Phil Mag. 6, 535-546.
JANGHORBAN, K. and ARDELL,A. J. (1979) The early stages of irradiation-induced 7' precipitation in proton
irradiated nickel-silicon alloys, d. Nucl. Mater. 85 & 86, 719-723.
JOHNSON, R. A. and DAMASK, A. C. (1964) Point defect configurations in irradiated iron-carbon alloys, Acta Met.
12, 443-445.
JOHNSON, R. A. and LAM, N. Q. (1976) Solute segregation in metals under irradiation, Phys. Rev. B, 13, 4364-4375.
JOHNSON, R. A. and LAM, N. Q. (1977) Solute segregation to voids during irradiation, Phys. Rev. B, 15, 1794-1800.
JOHNSON, R. A. and LAM, N. Q. 0978) Solute segregation under irradiation, J. Nucl. Mater. 69 & 70, 424-433.
JOHNSTON, W. G., MORRIS, W. C. and TURKALO, A. M. (1977) Excess sub-surface swelling produced by Ni ion
bombardment, in Radiation Effects in Breeder Reactor Structural Materials, (eds M. L. Bleiberg and J. W.
Bennett) TMS-AIME, New York, pp. 421-430.
JONES, R. H. (1978) Thoria redistribution in a Ni/ThO 2 alloy irradiated with 5 MeV Ni z+, J. Nucl. Mater. 74,
163-166.
KARKIN, A. E., ARKHIPOV, V. E., GOSHCHITSKll, B. N., ROMANOV, E. P. and SEDOROV, S. K. (1976) Radiation
effects in the superconductor NbDSn, Phys. Stat. Sol. (A) 38, 433-438.
KATZ, L. E. and HERMAN, H. (1967) Deuteron-irradiation effects in A1-4Wt~oCu, Acta Met. 15, 416-417.
KATZ, L. E., HERMAN, H. and DAMASK, A. C. (1968) Precipitation in neutron irradiated Al-base Cu., Acta Met.
16, 939-945.
KAUFMAN, L. and NESOR, H. 0975) Relation of the thermochemistry and phase diagrams of condensed systems,
in Treatise on Solid State Chemistry. Vol. 5, (ed. N. B. Hannay) Plenum Press, New York, pp. 179-232.
KAUFMAN, L., WATKIN, J. S., GITTUS, J. H. and MIODOWNIK, A. P. (1977) The effect of irradiation on the phase
stability of the sigma phase, Calphad, 1, 281-290.
KEEFER, n . W. and PARD, A. G. (1972/1973) Effects of injected helium on void formation in proton-irradiated
Type 316 Stainless Steel, J. Nucl. Mater. 45, 55-59.
KEEFER, D. W., PARD, A. G. and KRAMER, n . (1972) Swelling as a function of displacement damage in
proton-irradiated type 316 stainless steel, in Radiation-Induced Voids in Metals, (eds J. W. Corbett and L. C.
Ianniello) U.S. Atomic Energy Commission, Oak Ridge, TN, pp. 511-521.
KEEFER, D. W., PARD, A. G., RHODES, C. G. and KRAMER, n . (1971) Proton irradiation effects in type 321 stainless
steel, J. Nucl. Mater. 39, 229-233.
KERNOHAN, R. H., BILLINGTON, n . S. and LEWIS, A. B. (1956) Effect of neutron irradiation on the
precipitation-hardening reaction in alloys containing Beryllium, J. Appl. Phys. 27, 40-42.
KINCHIN, G. H. and PEASE, R. S. (1955) The displacement of atoms in solids by radiation, Reports Progr. Phys.
18, 1-51.
KING, H. W. (1966) Quantitative size-factors for metallic solid solutions, J. Mater. Sci. 1, 79-90.
KING, R. T. and JOSTSONS, A. (1975) Irradiation damage in a 2.2 pct magnesium-aluminum alloy, Met. Trans.
6A, 863-868.
KINOSH1TA, C., HOBBS, L. W. and MITCHELL, T. E. (1981) Phase instability under electron irradiation in ~ Cu-Be
alloys, in Phase Stability during Irradiations, (eds J. R. Holland, L. K. Mansur and D. I. Potter) TMS-AIME,
Warrendale, PA, pp. 561-575.
KIRCHNER, L. G., SMIDT, F. A., JR., KULCINSKI, (3. L., SPRAGUE, J. A. and WESTMORELAND,J. E. (1976) Nickel
PHASE S T A B I L I T Y U N D E R I R R A D I A T I O N 425

ion damage in a precipitation hardened nickel-aluminum alloy, in Irradiation Effects on the Microstructure
and Properties of Metals, ASTM STP 611, American Society for Testing and Materials, Philadelphia, pp.
370-384.
KLAUMONZER,S., SCWOMACHER,B., R~NTZSCH, S., VO~L, G., S6LDNER, L. and BIEGER,H. (1982) Severe radiation
damage by heavy ions in glassy Pdg0Si20, Acta Met. 30, 1493-1502.
KNOLL, R. W., WILKES, P. and KULONSKI, G. L. (1981) Solute redistribution in heavy-ion bombarded copper
base alloys, in Phase Stability during Irradiation, (eds. J. R. Holland, L. K. Mansur and D. I. Potter)
TMS-AIME, Warrendale, PA, pp. 123-137.
KocH, R., WArn, R. P. and WOLLENBERGER,H. (1981) TEM-investigation of the microstructural evolution in
simulation-irradiated Cu-Be alloys, J. Nucl. Mater. 103 & 104, 1211-1216.
KONOBEEVSKY,S. T., PRAVDYUK,N. F. and KUTAITSEV,V. I. (1956) Effect of irradiation on structure and properties
of fissionable materials, Proc. of International Conference on the Peaceful Uses of Atomic Energy 7, 433-440.
KRISHNAN, K. and ABROMEIT, C. (1984) Calculation of radiation induced instability in concentrated alloys,
J. Phys. F. 14, 1103-1116.
KULCINSKI, G. L. (1979) Fusion reactors: Their challenge to material scientists, Contemp. Phys. 20, 417-447.
KURDJUMOV, G. V., KRITSr~YA, V. I., IL'1NA, V. A., KRULIKOVSKAYA,M. P., PARSmNA, 1. YA. and CHIRKO,
L. I. (1978) Change in crystal structure of austenite and martensite of aluminum steels under ~,-irradiation.
Soy. Phys. Dokl. 23, 853-854. (English translation.)
LAIDLER, J. J., GARNER, F. A. and THOMASL. E. (1976) Simulation experiments in the high voltage electron
microscope, in Radiation Damage in Metals, (eds. N. L. Peterson and S. D. Harkness.) American Society for
Metals, Metals Park, OH, Chap. 7, pp. 194-226.
LAM, N. Q., JANGHORBAN,K. and ARDELL, A. J. (1981) On the modelling of irradiation-induced homogeneous
precipitation in proton-bombarded Ni-Si solid solutions, J. Nucl. Mater. 101, 314-325.
LANDAUER, R. (1978) Stability in the dissipative steady state, Physics Today, November, 23-29.
LEE, E. H., MAZlASZ, P. J. and ROWCLIFVE,A. F. (1981) The structure and composition of phases occurring in
austenitic stainless steels in thermal and irradiation environments, in Phase Stability during Irradiation, (eds
J. R. Holland, D. I. Potter and L. K. Mansur) TMS-AIME, Warrendale, PA, pp. 191-218.
v. LENSA, W., BAR~LS, A., DWORSCnAK, F. and WOLLENBERGER,H. (1977) Radiation-induced Be segregation
in the alloy Cu/182 at.ppm Be, J. Nucl. Mater. 71, 78-81.
LIFSmTZ, I. M. and SLYOZOV,V. V. (1961) Kinetics of precipitation from supersaturated solid solutions, J. Phys.
Chem. Solids 19, 35-50.
LIou, K. Y. and WILKES, P. (1979) The radiation disorder model of phase stability, J. Nucl. Mater. 87, 317-330.
LIou, K. Y., WILKES, P., KULCINSm, G. L. and BELLEN,J. H. (1979) Void swelling and phase instability in heavy
ion irradiated Mo-Zr alloy, J. Nucl. Mater. 85 & 86, 735-738.
LITTLE, E. A. and STO~R, L. P. (1982) Microstructural stability of fast reactor irradiated 10-12% Cr
ferritic-martensitic stainless steels, in Effects of Radiation on Materials: Eleventh Conference, ASTM STP 782,
(eds H. R. Brager and J. S. Perrin) American Society for Testing and Materials, Philadelphia, pp. 207-233.
Ltu, H. C., KINOSH1TA,C. and MITCHELL, T. E. (1981) Defect aggregation and disordering in Ni3AI and NiAI
by electron irradiation, in Phase Stability during Irradiation, (eds J. R. Holland, L. K. Mansur and D. I. Potter)
TMS-AIME, Warrendale, PA, pp. 343-355.
LIu, K. S., KAWANO,O. and MURAKAM1,Y. (1972) Structural changes in age-hardenable aluminum alloys induced
by low temperature neutron irradiation, Rad. Effects, 15, 37-49.
LOTHE, J. (1966) Simplified considerations of the Onsager symmetry in the general diffusion equation of nucleation
theory, J. Chem. Phys. 45, 2678-2680.
LOTT, R. G., KULONSKI, G. L., WILKES, P. and SMITH, H. V., Jr. (1979) Effect of nickel and nitrogen on void
formation in ion bombarded vanadium, J. Nucl. Mater. 85 & 86, 751-755.
MANNING, J. R. (1968) Diffusion Kinetics for Atoms in Crystals, J. Van Nostrand, Princeton. 257 pp.
MANSUR, L. K. (1978) Void swelling in metals and alloys under irradiation: An assessment of the theory, Nucl.
Tech. 40, 5-34.
MARTH, P., AARONSON,H. I., LORIMER,G., BARTEL,T. L. and RUSSELL,K. C. (1976) Application of heterogeneous
nucleation theory to precipitate nucleation at G. P. zones, Met. Trans. 7A, 1519-1528.
MARTIN, G. (1975) Instabilit6 des solides cristallins sous irradiation, Phil. Mag. 32, 615-627.
MARTIN, G. (1980) Contribution of dissipative processes to radiation induced solid solution instability, Phys. Rev.
B 21, 2122-2130.
MARTIN, G., CAUVIN, R., BOCQUET,J. L. and BARBU, A. (1981) Dose rate effects on solid solution stability, in
Phase Stability during Irradiation, (eds J. R. Holland, L. K. Mansur and D. I. Potter) TMS-AIME, Warrendale,
PA, pp. 43-62.
MARWICK,A. D. (1978) Segregation in irradiated alloys: The inverse Kirkendall effect and the effect of constitution
on void swelling, J. Phys. F: Metal Phys. 8, 1849-1861.
MARWlCK, A. D., KENNEDY,W. A. D., MAZEY, D. J. and HUDSON, J. A. (1978) Segregation of nickel to voids
in an irradiated high-nickel alloy, Harwell Report AERE9-9103, AERE Harwell, England, 12 pp.
MARWXCK,A. D., PILLER, R. C. and SIVELL, P. M. (1979) Mechanisms of radiation-induced segregation in dilute
nickel alloys, J. Nucl. Mater. 83, 35~1.
426 PROGRESS IN M A T E R I A L S SCIENCE

MAYDET, S. I. and RUSSELL, K. C. (1977) Precipitate stability under irradiation: Point defect effects, J. Nucl. Mater.
64, 101-114.
MAZEY, D. J., HARRIES, D. R. and HUDSON, J. A. (1979) The effect of silicon and titanium on void swelling
and phase stability in 12Cr/15Ni/Fe alloys irradiated with 46 MeV nickel ions, in Irradiation Behavior of Metallic
Materials for Fast Reactor Core Components, (Eds. J. Poirier and J. M. Dupouy) CEA-DMECN, Gif-sur-Yvette,
France, pp. 61-67.
MAZEY, D. J., HARRIES, D. P. and HUDSON, J. A. (1980) The effects of silicon and titanium on void swelling
and phase stability in 12Cr-15Ni austenitic alloys irradiated with 46MeV nickel ions, J. Nucl. Mater. 89,
155-181.
MAZEY, D. J., HUDSON, J. A. and NELSON, R. S. (1971) The dose dependence of void swelling in AISI 316 stainless
steel under 20 MeV C 2+ irradiation at 525°C, J. Nucl. Mater. 41, 257-273.
MAZIASZ, P. J. (1979a) Precipitation response of austenitic stainless steel to simulated fusion environment, in The
Metal Science of Stainless Steels, (eds. E. W. Collings and H. W. King) TMS-AIME, Warrendale, PA, pp.
160-180.
MAZIASZ, P. J. (1979b) The precipitation response of 20~o cold worked type 316 stainless steel to simulated fusion
irradiation, J. Nucl. Mater. 85 & 86, 713-717.
MAZIASZ, P. J. (1982) The effects of increased helium content on void formation and solute segregation in neutron
irradiated type 316 stainless steel, J. Nucl. Mater. 108 & 109, 359-384.
MAZIASZ, P. J., HORAK, J. A. and Cox, B. L. (1981) The influence of both helium and neutron irradiation on
precipitation in 20~o CW austenitic stainless steel, in Phase Stability during Irradiation, (eds. J. R. Holland,
L. K. Mansur and D. I. Potter) TMS-AIME, Warrendale, PA, pp. 271-292.
MCELROY, W. N. and FARRAR, H., IV (1972) Helium production in stainless steel and its constituents as related
to LMFBR development, in Radiation Induced Voids in Metals, (eds. J. W. Corbett and L. C. Ianniello) U.S.
Atomic Energy Commission, Washington, D.C., pp. 187-228.
MEIRE-HmMER, R., KOPFER, H. (1982) Influence of neutron radiation induced defects on the superconducting
properties of VDSi, J. Nucl. Mater. 108 & 109, 593-602.
MERRICK, H. F. (1978) Precipitation in nickel-base alloys, in Precipitation Processes in Solids, (eds. K. C. Russell
and H. I. Aaronson) TMS-AIME, Warrendale, PA, pp. 161-190.
MIURA, K., FIORE, N. F. and ALLEN, C. W. (1975) The influence of Mg in ~/-irradiation-induced changes in A1
alloys, J. Nucl. Mater. 55, 53-63.
MOZER, B., KEATING, D. T. and Moss, S. C. (1968) Neutron measurement of clustering in the alloy CuNi, Phys.
Rev. 175, 868-876.
MRUZIK, i . R. and RUSSELL, K. C. (1978) The effect of irradiation on the nucleation of incoherent precipitates,
J. Nucl. Mater. 78, 343-353.
MUKAI, T. and MITCHELL, T. E. (1982) Radiation-induced homogeneous precipitation in Ni-la/oBe alloys, J. Nucl.
Mater. 105, 149-158.
MURRAY, G. T. and TAYLOR, W. E. (1954) Effect of neutron irradiation on a supersaturated solid solution of
beryllium in copper, Acta Met. 2, 52-62.
NELSON, R. S., Ed. (1975) The Physics of Irradiation Produeed Voids, HMSO, London. 350 pp.
NELSON, R. S., HUDSON, J. A. and MAZEY, D. J. (1972) The stability of precipitates in an irradiation environment,
J. Nucl. Mater. 44, 318-330.
NEMOTO, M., OGUCH1, T. and SUTO, H. (1971) Effects of electron irradiation on precipitation phenomena in a
high voltage electron microscope, J. Japan Inst. Metals, 35, 886-891.
NOWICK, A. S. and WEISBERG, L. R. (1958) A simple treatment of ordering kinetics, Acta Met. 6, 260-265.
NUTTALL, K. and FAULKNER, D. (1977) The effect of irradiation on the stability of precipitates in Zr-2.5Wt~oNb
alloys, J. Nucl. Mater. 67, 131-139.
OHNUKI, S., TAKAHASH1,H. and TAKEYAMA,T. (1981) Void swelling and segregation of solute in ion-irradiated
ferritic steels, J. Nucl. Mater. 103 & 104, 1121-1126.
OHNUKI, S., TAKEYAMA, T. and TAKAHASHI, H. (1982) Radiation-induced segregation at internal defect sinks in
electron irradiated FCC alloys, in Point Defects and Defect Interactions in Metals, (eds. J-I. Takamura, M.
Doyama, and M. Kiritani) Univ. of Tokyo Press, pp. 954-957.
OKAMOTO, P. R. and PENN, L. E. (1979) Radiation induced segregation in binary and ternary alloys, J. Nucl.
Mater. 83, 2-23.
OKAMOTO, P. R., TAYLOR, A. and WmDERSlCH, H. (1975) Effect of Be doping on void swelling in nickel, in Proc.
International Conference on Fundamental Aspects of Radiation damage in Metals, USAERDA Conf. 751006-P-2,
pp. 1188-1195.
OKAMOTO,P. R. and WIEDERSICH, H. (1974) Segregation of alloying elements to free surfaces during irradiation,
J. Nucl. Mater. 53, 336-345.
OLANDER, D. R. (1976) Fundamental Aspects of Nuclear Reactor Fuel Elements, Technical Information Center,
ERDA, Washington, Chap. 17, pp. 373415.
OSTROVSKY, Z. E., SERVYAEV,G. A., GRINCHUK, P. P., VOTINOV, S. N. and PROKrtOROK, V. I. (1971) The effect
of neutron irradiation on second phase precipitation processes, Trans. Iron & Steel Inst. Japan, 11, 289-293.
OZAWA, E. and KIMORA, H. (1970) Excess vacancies and the nucleation of precipitates in aluminum-silicon alloys,
Acta Met. 18, 995-1004.
PHASE S T A B I L I T Y U N D E R I R R A D I A T I O N 427

PARKER, C. A, and RUSSELL, K. C. (1982) Calculation of cavity nucleation under irradiation with continuous
helium generation, in Effects of Radiation on Materials: Eleventh Conference, ASTM STP 782, (Eds. H. R. Brager
and J. S. Perrin) American Society for Testing and Materials, Philadelphia, pp. 1042-1053.
PEARSON, W. B. (1967) Lattice Spacings and Structures of Metals and Alloys, Pergamon Press, New York. Vol
l, 1044 pp; Vol. 2, 1446 pp.
PETERSON,D. T. (1982) Swelling in neutron irradiated titanium alloys, in Effects of Radiation on Materials: Eleventh
Conference A S T M STP 782, (eds H. R. Brager and J. S. Perrin) American Society for Testing and Materials,
Philadelphia, pp. 260-274.
PrlSTER, J. C. (1975) Ion implantation, in Radiation Damage Processes in materials, (ed. C. H. S. Dupuy) Noordhoff
International Publ., Leyden, pp. 467-475.
PIERAGOSTINI,F., CERASARA,S. and FEDERIGHI,T. (1966) Clustering in A1-4%Cu alloy after neutron irradiation
at -195°C, Acta Met. 14, 450-452.
I~ERCY, G. R. (1962) The detection of displacement spikes by measuring the effect of particle size on the dissolution
of precipitate particles by fast neutron irradiation, J. Phys. Chem. Solids 23, 463-477.
POERSCHKE, R. and WOLLENBERGER,H. (1975) Radiation enhanced diffusion in Cu-Ni alloys below 200 K, Rad.
Effects 24, 217-221.
POERSCHKE, R. and WOLLENBERGER,H. (1976) Kinetics of interstitialcy diffusion in electron-irradiated Cu-Ni
alloys, J. Phys. F: Metal Phys. 6, 27-41.
PORTER, D. L, and WOOD, E. L. (1979) In-reactor precipitation and ferritic stabilization in neutron-irradiated
stainless steel, J. Nucl. Mater. 83, 90-97.
PORTER, L. F. and DIENES, G. J. (1959) Effect of neutron irradiation on the martensite transformation in
iron-nickel alloys, Trans. Met. Soc. A I M E 215, 854-863.
POTTER, D. I. (1978) Radiation-induced precipitate redistribution in thin foils, Rad. Effects 35, 115-117.
POTTER, D. I. (1981) Ion-bombardment-induced instabilities in ordered precipitates, in Phase Stability during
Irradiation, (eds. J. R. Holland, L. K. Mansur and D. I. Potter) TMS-AIME, Warrendale, PA, pp. 521-546.
POTTER, D. I. and HERNANDEZ,O. G. (1981) Ion-bombardment-induced disordering of y' Ni3Si, Acta Met. 29,
187-196.
POTTER, D. I. and HOFF, H. A. (1976) Irradiation effects on precipitation in ~/?' Ni-Al alloys, Acta Met. 24,
1155-1164.
POTTER, D. I. and McCORMICK, A. W. (1979) Irradiation enhanced coarsening in Ni-12.8at.%A1, Acta Met. 27,
933-94 I.
POTTER, D. I., OKAMOTO,P. R., WIEDERSICH,H., WALLACE,J. R. and MCCORMXCK,A. W. (1979) Heterogeneous
precipitation at internal and external surfaces during irradiation of Ni-12.Tat.%Si, Acta Met. 27, 1175-1185.
POTTER, D. I., REHN, L. E., OKAMOTO,P. R. and WIEDERSICH,H. (1977) Void swelling and segregation in dilute
nickel alloys, in Radiation Effects in Breeder Reactor Structural Materials, (eds. M. L. Bleiberg and J. W.
Bennett) TMS-AIME, New York, pp. 377-386.
POTTER, D. I. and RYDING, D. G. (1977) Precipitate coarsening, redistribution and renucleation during irradiation
of Ni-6.35w/oA1, J. Nuc/. Mater. 71, 14-24.
POTTER, D. I. and WIEDERSICH,H. (1979) Mechanisms and kinetics of precipitate restructuring during irradiation,
J. Nucl. Mater. 83, 208-213.
PRAMANIK,D. and SEIDMAN,D. N. (1983) Atomic resolution observations of nonlinear depleted zones in tungsten
irradiated with metallic diatomic molecular ions, J. Appl. Phys. $A, 6352-6367.
PRIGOGINE, I. (1967) Introduction to Thermodynamics of Irreversible Processes, 3rd edn., Interscience, New York.
147 pp.
RAU, R. C. and LADD, R. L. (1969) Cylindrical damage shells in irradiated vanadium, J. Appl. Phys. 40, 2899-2904.
RECHTIN, M. D., VANDERSANDE, J. B. and BALDO, P. M. (1978) Ion-implantation damage in amorphous and
crystalline Nb40Nir0, Scripta Met. 12, 639-649.
REHN, L. E., AGARWAL,S. C. and NOLFI, F. V., Jr. (1979) Radiation-induced solute segregation in a V-15wt~Cr
alloy, J. Nucl. Mater. 85 & 86, 763-767.
REHN, L. E., OKAMOTO,P. R., POTTER, D. I. and WIEDERSlCH,H. (1978) Effect of solute misfit and temperature
on radiation-induced segregation in binary Ni alloys, J. Nucl. Matter. 74, 242-251.
REHN, L. E., WAGNER,W. and WIEDERSICH,H. (1981) Radiation-induced segregation in concentrated Cu-Ni alloys,
Scripta Met. 15, 683-687.
REYNOLDS, M. B., Low, J. R., JR. and SULLIVAN,L. O. (1955) Study of the radiation stability of austenitic type
347 stainless steels, J. Metals, 7, 555-559.
RICHARDS, J. T. (1955) Effect of irradiation on beryllium copper, Acta met. 3, 211-212.
Ro, H. and MITCHELL, T. E. (1978) Effects of electron irradiation on precipitation in Ni-Al alloys, Met. Trans.
9a, 1749-1760.
ROBROCK, K-H. and OKAMOTO,P. R. (1980) Radiation-induced precipitation in a single-phase Ni-6at.%Si alloy,
in Irradiation Behavior o f Metallic Materials for Fast Reactor Core Components, (eds. J. Poirier and J. M.
Dupouy) CEA-DMECN, Gif-sur-Yvette, France, pp. 57-60.
ROSENBAUM, H. S. and TURNBULL, D. (1959) Metallographic investigation of precipitation of silicon from
aluminum, Acta Met. 7, 664-674.
428 PROGRESS IN MATERIALS SCIENCE

RUSBRIDGE, K. L. (1981) A Study o f Electron and Ion Irradiation of Aluminum-Germanium Alloys, Ph.D. Thesis,
Univ. of Salford, England, 166 pp.
RUSBRIDGE, K. L. (1983) Amorphisation of Ge precipitates in an A1-Ge alloy by electron bombardment, Rad.
Effects 60, 277-291.
RUSSELL, K. C. (1971) Nucleation of voids in irradiated metals, Acta Met. 19, 753-758.
RUSSELL, K. C. (1980) Nucleation in solids: The induction and steady state effects, Adv. Coll. Int. Sci. 13, 205-318.
RUSSELL, K. C. and AARONSON, H. I. (1975) Sequences of precipitate nucleation, J. Mater. Sci. 10, 1991-1999.
RUSSELL, K. C. and POWELL, R. W. (1973) Dislocation loop nucleation in irradiated metals, Acta Met. 21, 187-193.
SAIEDFAR, M. S. and RUSSELL, K. C. (1979) Oxide dispersoid stability in irradiated alloys, J. Nucl. Mater. 85
& 86, 931-934.
SANTHANAM, A. T., TAYLOR, A., KESTEL, B. J. and STEVES, C. (1975) Swelling studies of vanadium and a
vanadium-lwt~otitanium alloy using ion simulation techniques, J. Vac. Sci. Tech. 12, 528-531.
SCHULSON, E. M. (1979) The ordering and disordering of solid solutions under irradiation. J. Nucl. Mater. 83,
239-264.
SCH/3LE, W. (1964) Note on the paper of Cerasara, S., Federighi, T. and Pieragostini, F., entitled "Pre-precipitation
rate in Al-10~oZn alloy neutron irradiated at 78 K", Phil. Mag. 10, 913-915.
SCH~LE, W. (1969) Pre-precipitation rate in an aluminum-10~o Zinc alloy irradiated at 78K, Phil. Mag. 19,
1085-1088.
SCHLILE, W., SPINDLER, P. and LANG, E. (1975) Entmischung in Kupfer-Nickel-Legierungen durch thermische
Behandlung und durch Neutronbestrahlung, Z. Metallkunde 66, 50--56.
SETHI, V. K. and ORAMOTO, P. R. (1981) Radiation-induced segregation in complex alloys, in Phase Stability
during Irradiation, (eds. J. R. Holland, L. K. Mansur and D. I. Potter) TMS-AIME, Warrendale, PA, pp.
109-121.
SHARPE, R. M. and WHAPHAM, A. D. (1974) Effects of irradiation on ~/' precipitates in nimonic PE-16 alloy,
in Conference on Irradiation Behavior in Fuel Cladding and Core Components, KTG/BNES, Karlsruhe, Germany,
pp. 103-107.
SHEPHERD, J. P. (1969) Irradiation-induced precipitation in a niobium-stabilized stainless steel, Metal Sci. J. 3,
229-234.
SHEPHERD, J. P. (1976) Excess vacancy concentrations and the nucleation of niobium carbide precipitates in
austenitic stainless steel, Metal Sci., May 1976, 174-179.
SHIRAISHI, K. (1970) Precipitation hardening of aluminum 2.2wt~olithium alloy after neutron irradiation, Trans.
Japan. Inst. Metals 11, 381-384.
SHRIVER, I . L. and WUTTIG, M. (1972) Irradiation enhanced decomposition of a nickel-carbon solid solution,
Acta Met. 20, 1-4.
SIKKA, V. K. and MOTEFF, J. (1974) Identification of ~t-Mn crystal structure in neutron-irradiated W-Re alloy,
Met. Trans. 5, 1514-1517.
SILVESTRE,G., SILVENT,A., REGNARD,C. and SAINFORT,G. (1975) Alliages de nickel-fer et de nickel-silicum ne
gonflant pas sous irradiation aux neutrons rapides, J. Nucl. Mater. 57, 125-135.
SIZMANN, R. (1978) The effect of radiation upon diffusion in metals, J. Nucl. Mater. 69 & 70, 386-412.
SKLAD,P. S. and MITCHELL,Z. E. (1974) Radiation-enhanced precipitation in A1-4~oCu by high voltage electron
microscopy, Scripta Met. 8, Ill3-1118.
SKLAD, P. S. and MITCHELL,Z. E. (1975) Effects of electron irradiation on precipitation in A1-3.Sw/oCu, Acta
Met. 23, 1287-1302.
SPRAGUE, J. A., SMIDT, F. A., Jr. and REED, J. R. 0979) The microstructures of some refractory metals and
alloys following neutron irradiation at 650°C, 3. Nucl. Mater. 85 & 86, 739-743.
STANLEY,J. T. (1964) The effect of irradiation on precipitation of nitrides in iron, in Diffusion in Body Centered
Cubic Metals, American Society for Metals, Metals Park, OH, pp. 349-356.
STANLEY,J. T. (1979) Magnetic properties of irradiated austenitic stainless steel, J. Nucl. Mater. 85 & 86, 787-791.
STANLEY, J. Z. and GARR, K. R. (1975) Ferrite formation in neutron irradiated type 316 stainless steel, Met.
Trans. 6A, 531-535.
STEIGLER, J. O. (1975) Proceedings of the Workshop on Correlation of Neutron and Charged Particle Damage, Conf.
760673, Oak Ridge National Laboratory, Oak ridge, TN, 376 pp.
TAKAHASHI, H., OHNUKI, S. and TAKEYAMA,T. (1981) Radiation-induced segregation at internal sinks in electron
irradiated binary alloys, 3. Nucl. Mater. 103 & llM, 1415-1420.
TAKEYAMA, T., OHNUKI, S., TAKAHASHI, H., SATO, Y., and MOCmZUKI, S. (1982a) Radiation-induced segregation
in ion-irradiated vanadium-carbon alloys, Bulletin o f the Faculty of Engineering, Hokkaido University, Vol. 110,
pp. 209-218.
TAKEYAMA, T. and TAKAHASHI, H. (1974) Effect of electron irradiation on the precipitation phenomena of a Fe-C
alloy, J. Japan. Inst. Metals 38, 138-143.
TAKEYAMA, T. and TAKAHASHI, H. (1975a) Effect of electron irradiation on precipitation of carbon and nitrogen
in alpha iron, in Proceedings of lnternational Conference on Fundamental Aspects o f Radiation Damage in Metals,
USAERDA Conf 751006, Gatlinburg, TN, Vol. II, pp. 1100-1106.
TAKEYAMA, T. and TAKAHASHI, H. (1975b) The effect of electron irradiation on the precipitation of Fe-N alloy,
J. Phys. Soc. Japan, 38, 1783.
PHASE S T A B I L I T Y U N D E R I R R A D I A T I O N 429

TAKEYAMA,T., TAKAHASHI,H., and OHmml, S. (1982b) Effect of substitutional and interstitial elements on void
formation in neutron-irradiated vanadium alloys, J. Nucl. mater. 106 & 109, 465-475.
THOMAS, G., MORI, H., FUJITA, H. and SINCLAIR, R. (1982) Electron irradiation-induced crystalline amorphous
transitions in Ni-Ti alloys, Scripta Met. 16, 589-592.
THOMAS, L. E. (1981) The stability of y' and ~," in inconel 706 under neutron irradiation, in Phase stability during
Irradiation, (eds. J. R. Holland, L. K. Mansur and D. I. Potter) TMS-AIME, Warrendale, PA, pp. 237-255.
THOMAS, L. E. (1982) Analytical electron microscopy of neutron-irradiated reactor alloys, in Proceedings, 40th
Annual Meeting, Electron Microscopy Society of America, (ed. G. W. Bailey,) Claitors Publ., Baton Rouge, LA,
pp. 594-597.
THOMPSON, i . W. (1969) Defects and Radiation Damage in Metals, Cambridge Univ. Press. London, 384p.
TJHIA, E., WILKES,P. and KULCINSKI,G. L. (1980) Irradiation-induced precipitation of AI-Si alloys in the HVEM,
Rad. Effects, 51, 49-56.
TSUBAKIHARA, H., KATAH, S., NISHIMOTO, M. and OKADA, T. (1975) Neutron irradiation damage of the
superconducting properties in vanadium and Nb-Ti alloys, in Fundamental Aspects of Radiation Damage in
Metals, Conf. 751006-P1, (eds M. T. Robinson and F. W. Young Jr.) National Technical Information Service,
Springfield, VA, pp. 1155--1161.
TUCKER,C. W., Jr. and WEBB, M. B. (1959) Electron irradiation of aluminum-copper alloys, A cta Met. 7, 187-190.
URBAN, K. and MARTIN,G. (1982) Precipitate coarsening by point defect recombination in alloys under irradiation,
Acta Met. 30, 1209-1218.
URBAN, K. and MARTIN, G. (1983) Precipitate coarsening in alloys under irradiation, in Proceedings of an
International Conference on Solid-Solid Phase Transformations, (eds. H. I. Aaronson, D. E. Laughlin, R. F.
Sekerka and C. M. Wayman) TMS-AIME, Warrendale, PA, pp. 317-322.
URBAN, K. and YOSHIDA,N. (1981) The threshold energy for atom displacement in irradiated copper studied
by high-voltage electron microscopy, Phil. Mag. A 44, 1193-1212.
URBANIC, V. F., LESURF, J. E. and JOHNSON, A. B., Jr. (1975) Effect of aging and irradiation on the corrosion
of Zr-2.Swt%Nb, Corrosion, NACE, 31, No. 1, pp. 15-20.
VAIDYA,W. V. (1979) Modification of the precipitate interface under irradiation and its effect on the stability
of precipitates, J. Nucl. Mater. 83, 223-230.
VOGL, G. and WEISS, B. (1965) Der Einflus von Neutronenbestrahlung auf die Ausschiedungskinetik in eider
iibers/ittigten AI-Cu-Legierung, Acta Met. 13, 578-582.
WAGENBLAST,H. and DAMASK, A. C. (1962) Kinetics of carbon precipitation in irradiated iron, J. Phys. Chem.
Solids, 23, 221-227.
WAGENBLAST, H., FUJITA, F'. E. and DAMASK,A. C. (1964) Kinetics of carbon precipitation in irradiated iron.
IV. Electron microscope observations, Acta Met. 12, 347-353.
WAGNER, A. and SEIDMAN,D. N. (1979) Direct observation of solute segregation to voids in a fast-neutron
irradiated Mo-lat.%Ti alloy, 3. Nucl. Mater. 83, 48-56.
WAGNER, C. (1961) Theorie der Alterung von Niederschlagen durch Umlosen, Z. Elektrochem. 65, 581-591.
WAGNER, W., POERSCHKE,R. and AXMANN, H. (1980) Neutron scattering studies of an electron irradiated
62Ni-41.4at.%65Cu alloy, Phys. Rev. B, 21, 3087-3099.
WAGNER, W., POERSCHKE, R. and WOLLENBERGER,H. (1982) Short range clustering and long range periodic
decomposition of an electron-irradiated Ni-Cu alloy, J. Phys. F.: Met. Phys. 12, 405-424.
WAHl, R. P., and WOLLENBERGER,H. (1983) Microstructural evolution in a Cu-l.35at.%Be alloy under electron
irradiation in a high voltage microscope, J. Nucl. Mater. 113, 207-210.
WANG, Z., AYRAULT, G. and W1EDERSICH,H. (1982) Segregation in irradiated titanium alloys, J. Nucl. Mater,
10~ & 109, 331-338.
WATKIN, J. S., GITTUS, J. H. and STANDRING,J. (1977) The influence of alloy constitution on the swelling of
austenitic stainless steels and nickel based alloys, in Radiation effects in Breeder Reactor Structural Materials,
(eds. M. L. Bleiberg and J. W. Bennett) TMS-AIME, New York, pp. 467-477.
WEAVER, L., HUDSON, B. and NUTTING, J. (1977) The enhanced growth of precipitates in A1-4%Cu alloys under
irradiation, Report AERE-R8850, AERE, Harwell, England, 26 pp.
WEBER, H. W. (1982) Neutron irradiation effects on alloy superconductors, d. Nucl. Mater. 108 & 109, 572-584.
WEBER, W. J., KULC1NSKI, G. L,, LOTT, R. G., WILKES, P. and SMITH, H. V., Jr. (1975) Ion simulation studies
of void formation in high purity vanadium, in Proceedings of the International Conference on Radiation Effects
and Tritium Technology, Gatlinburg, TN, 1-130-I-149.
WEISS, B. and STICKLER,R. (1972) Phase instabilities during high temperature exposure of 316 austenitic stainless
steel, Met. Trans. 3, 851-866.
WESTMACOTT,K. H. and PEREZ, M. I. (1979) The co-precipitation of vacancies and carbon atoms in quenched
platinum, J. Nucl. Mater. 83, 231-237.
WIEDERSICH,H., OKAMOTO,P. R. and LAM, N. Q. (1977) Solute segregation during irradiation, in Radiation Effects
in Breeder Reactor Structural Materials, (eds. M. L. Bleiberg and J. W. Bennett) TMS-AIME, New York, pp.
801-819.
WIEDERSICH,H., OKAMOTO,P. R. and LAM, N. Q. (1979) A theory of radiation-induced segregation in concentrated
alloys, J. Nucl. Matter. 83, 98-108.
430 PROGRESS IN MATERIALS SCIENCE

WIGNER, E. P. (1946) Theoretical physics in the metallurgical laboratory of Chicago. J. Appl. Phys, 17, 857-863.
WILKES, P. (1979) Phase stability under irradiation--A review of theory and experiment, J. Nucl. Mater. 83,
166-175.
WILKES, P., LIOU, K. Y. and LOTT, R. G. (1976) Comments on radiation induced phase instability, Rad. Effects,
29, 249-251.
WILLIAMS, R. K., WIFFEN,F. W., BENTLEY,J. and STEIGLER,J. O. (1983) Precipitation in fast neutron-irradiated
tungsten-rhenium alloys, Met. Trans. 14A, 655-666.
WILLIAMS, T. M. (1982) Precipitation in irradiated and unirradiated austenitic stainless steels, in Effects on
Radiation on Materials; Eleventh Conference, ASTM STP 782, (eds. H. R. Brager and J. S. Perrin) American
Society for Testing and Materials, Philadelphia, pp. 166-185.
WILLIAMS,T. M. and ARKELL, D. R. (1979) Void-swelling in 20% cold worked austenitic stainless steel irradiated
with 22 MeV C 2+ or 46.5 MeV Ni 6÷ ions, J. Nucl. Mater. 80, 79-87.
WILLIAMS, T. M., ARKELL, D. R. and EYRE, B. (1977) The void swelling behavior of solution treated FV 548
stainless steel irradiated with 22 MeV C 2÷ and 46.5 MeV Ni 6+ ions and the influence of heat treatment, J. Nucl.
Mater. 68, 69-81.
WILLIAMS,T. M. and EYRE, B. L. (1976) Void swelling in solution treated FV 548 steel irradiated in a high voltage
electron microscope, J. Nucl. Mater. 59, 18-28.
WILLIAMS, T. M., TICHTMARSH, J. M., and ARKELL, D. R. (1982) Void-swelling and precipitation in
neutron-irradiated niobium stabilised austenitic stainless steel, J. Nucl. Mater. 107, 222-244.
WOHOFSKY,O., and WA1DELICH,W. (1968) Kleinwinkelstreuung an strahleninduzierten Ausscheidungen in Pb-Ni
Legierungen, Z. Angew Phys. 25, 372-379.
WOLFER, W. G. and ASHKIN,M. (1975) Stress-induced diffusion of point defects to spherical sinks, J. Appl. Phys.
46, 547-557.
WOOLHOUSE, G. R. and IPOHORSKI, M. (1971) On the interaction between radiation damage and coherent
precipitates, Proc. R. Soc. Lond. A, 324, 415-431.
YAMANE, T., TAKAHASHI,J. and YAMASHITA,N. (1979) Effect of electron irradiation on precipitation in Fe-N
alloy, Rad. Effects 40, 95-96.
YAMAUCH1, H., SANCHEZ, J. M., de FONTAINE, n . and KIKUCHI, R. (1979) A thermodynamical approach to
irradiation-induced precipitation in undersaturated solid solutions, in Irradiation Behavior of Metallic Materials
For Fast Reactor Core Components, (eds. J. Poirier and J. M. Dupouy) CEA-DMECN, Gif-sur-Yvette, France,
pp. 81-87.
YANG, W. J. S. (1982) Grain boundary segregation in solution-treated nimonic PEI6 during neutron irradiation,
J. Nucl. Mater. 108 & 109, 339-346.
YANG, W. J. S., BRAGER,n . R. and GARNER, F. A. (1981) Radiation-induced phase development in AISI 316,
in Phase Stability during Irradiation, (eds. J. R. Holland, L. K. Mansur and D. I. Potter) TMS-AIME,
Warrendale, PA pp. 257-269.
YANG, W. J. S., and GARNER, F- A. (1983) Relationship between phase development and swelling of AISI 316
during temperature changes, in Effects of Radiation on Materials, Eleventh Conference, ASTM STP 782, (eds.
H. R. Brager and J. S. Perrin) American Society for Testing and Materials, Philadelphia, pp. 186-206.
YOSmDA, H. (1969) Ageing characteristics of neutron-irradiated copper-9.5at.%beryllium alloy, Phil. Mag. 19,
987-991.
YOSI-nDA, H. and SAGANE, T. (1972) Zone formation and precipitation in Cu-Be alloys during aging under
irradiation, J. Nucl. Sci. Tech. 9, 1-6.
YOSrtIDA, H., YAMAMOTO,S., MURAKAMI,Y. and KODAKA,H. (1971) The effects of neutron irradiation on aging
and precipitation phenomena in copper-beryllium alloys with or without additional elements, Trans. Japan.
Inst. Metals, 12, 229-237.

NOTES ADDED IN PROOF

R e c o i l Resolution
Recoil r e s o l u t i o n t r e a t m e n t s d e s c r i b e d in S e c t i o n 7 are alike i n r e p r e s e n t i n g the r e s o l u t i o n
p r o c e s s as a s o u r c e t e r m in the d i f f u s i o n e q u a t i o n , w h i c h is t h e n solved s u b j e c t to v a r i o u s
assumptions and approximations.
I n spite o f differing a s s u m p t i o n s a n d a p p r o x i m a t i o n s , all the s o u r c e t e r m m o d e l s m a k e the
s a m e p r e d i c t i o n : i r r a d i a t i o n m a y c a u s e small particles to g r o w at the e x p e n s e o f l a r g e r o n e s
a n d give a stable, s t e a d y - s t a t e d i s t r i b u t i o n o f s m a l l e r particles.
M a r t i n [ A - l ] v e r y r e c e n t l y t o o k a n e n t i r e l y different a p p r o a c h b y a d d i n g the recoil
PHASE STABILITY UNDER IRRADIATION 431

resolution displacement jumps to thermally activated jumps to obtain an irradiation-altered


diffusion equation. The resulting flux equation was:
{(f / ) B/' 63C t33C~)

where
M = thermal atomic mobility
N,. = number of atoms per unit volume
f" = sec__ond derivative of free energy with respect to composition
/9 B - K~ 2 = ballistic diffusion coefficient
K = atomic displacement rate
f2= mean square displacement distance
x = gradient energy.
Equation A-(l) is simply the well known diffusion equation of Cahn [A-2] w i t h f " replaced
by f" + DB/M. Thus, as was found by Maydet and Russell in a study of irradiation-enhanced
precipitation (Section 4.1), the effect of irradiation is to contribute a kinetic term to the free
energy to produce an irradiation-altered potential, which functions in some ways as a free
energy.
A Lyupanov functional analysis showed that in the regular solution model the effect of
irradiation is to cause the irradiated alloy to have the configuration at temperature T that
it would have at a temperature T' outside irradiation:

where/9 is the interdiffusion coefficient in the absence of ballistic effects.


The solubility of all particles is thus increased over that in the absence of irradiation. The
onset of irradiation would then cause partial dissolution of a system of particles to establish
the irradiation-enhanced solubility, but would in no event cause small particles to grow at the
expense of the larger ones, as predicted by source term analyses.
Equation A-(l) is subject to some serious approximations./)B was assumed independent
of composition, and the equation only describes diffusion over distances large compared to
the recoil distance, which may be the order of 100 A. The latter limitation is particularly
severe as Eq. A-(l) would not describe fine scale precipitation or spinodal decomposition, nor
would it describe events occurring at or near a sharp particle:matrix interface. Yet, it is not
at all apparent how a ballistic diffusion model, even relieved from these limitations, would
give inverse coarsening.
Source term treatments are also approximations, as noted earlier. Such recoil resolution
calculations also take the solute concentration at the particle:matrix interface as the thermal
equilibrium value even though, as shown in Section 4, vacancy: self-interstitial recombination
at the particle: matrix interface may alter particle solubility. In addition, none of the existing
theories of recoil resolution account for irradiation-induced solute segregation or for point
defect generation within the particles. Further theoretical work is needed.

Ordering and Disordering


Martin [A-l] and Bellon and Martin [A-3] applied the ballistic diffusion/Lyupanov
functional formulation used for recoil resolution to develop a theory for the order-disorder
reaction in irradiated alloys. They used the Bragg-Williams approximation to determine alloy
432 PROGRESS IN MATERIALS SCIENCE

stability and found qualitative agreement between the theory and the observed temperature
regions of stability of LRO and SRO in HVEM irradiated Ni4Mo described in Section 14.
Absence of quantitative agreement was attributed to the inadequacy of the simple
Bragg-Williams solid solution model.

Aluminum Based Alloys


Rusbridge [A-4] recently studied the effect of recoil resolution in precipitate stability during
irradiation of an Al-l.9w/oGe alloy. Two-phase samples of Ge precipitates in an A1 matrix
were irradiated at -100°C with 200keV A1÷ ions to doses between 6 and 122 dpa.
Transmission electron microscopy observations showed that individual precipitates were
replaced by families of precipitates as a result of low-temperature irradiation and warming
to room temperature. The microstructure at room temperature following low-temperature
irradiation was thought to be determined by the amount of cascade mixing of solute atoms
and solvent atoms during irradiation and the degree of re-nucleation in the solute-enriched
region surrounding the original preciptate site during warming to room temperature. The
results agreed with the predictions of the theory of Nelson et al., discussed in Section 7.

Copper Based Alloys


Mori et al. [A-5] irradiated a series of alloys in the 2MV HVEM at 160K and
1 x 1024e-/m2.s (about 10 -2 dpa/s). Presumably, displacement levels of tens of dpa were
reached. Of the alloys, Cu3 Ti2, CuTi, CuZr, and Cut0Zr7 became amorphous whereas Cu4 Ti
and CuZn remained crystalline.

Iron Based Alloys, Ferritic


Brimhall, et al. [A-6] irradiated FeA1 and Fe3A1 with 2.5 MeV Ni + ions at a displacement
rate of 3 x 10 -3 dpa/s. Irradiation was at or near room temperature at a displacement level
typically from 1-10 dpa, and occasionally to 45 dpa. Neither alloy became amorphous.
Berkowitz et al. [A-7] bombarded FeTsB25 with 4 MeV Ar ions. Irradiation was at room
temperature at a displacement rate of 7 x 10 -4 dpa/s near the surface. The starting material
was rapidly solidified ribbons which had been annealed to produce mostly Fe3 B with small
amounts of Fe2B and ~-Fe. All phases were crystalline. The Fe3 B phase became amorphous
after. 15 dpa, and diffraction spots corresponding to the Fe2 B phase disappeared between 17
and 50 dpa, due to either amorphization or ion beam mixing. The ct-Fe particle size decreased
with increasing dose and the particles disappeared after 50 dpa, apparently by the sort of ion
beam mixing described by Rusbridge [A-4] in her studies of irradiated AI-Ge alloys.
Light water reactor pressure vessels, typically made of low alloy ferritic steels, operate at
temperatures in the vicinity of 300°C and are usually exposed to a fast neutron fluence
of < 1020n/cm 2 (E > 1 MeV) over the lifetime of the reactor. Such a fluence gives < 0.05 dpa
and would have little or no effect on most other alloys. Yet these fast neutron fluences give
very severe embrittlement of the steel, which makes the pressure vessel potentially susceptible
to catastrophic service failure [A-8]. Copper and phosphorus, which appear in the steel as
tramp elements, are known to enhance greatly the embrittling effect of irradiation [A-8].
The embrittling effect has been attributed to strengthening due to such irradiation-induced
defect aggregates as voids, dislocation loops, various kinds of precipitates, and point-
defect: solute complexes.
PHASE STABILITY UNDER IRRADIATION 433

Grant and Fortner [A-9] were the first to suggest that fine copper precipitates were the cause
of the embrittlement, but had no microstructural evidence to support their suggestion.
Subsequently, Brenner et al. [A-10] made a field-ion microscope study of a neutron irradiated
Fe-0.34w/oCu alloy. Some 8 x 1017/cm 3 defects of 6/~ mean diameter were found; these were
tentatively identified as Cu-stabilized voids.
Later, Odette and Sheeks [A-11] and Odette [A-12] proposed a mechanism by which
Cu-rich precipitates could be formed at displacement cascades and grow by irradiation-
enhanced diffusion. Frisius et al. [A-13] interpreted their small angle neutron scattering results
on irradiated steels in terms of fine Cu-rich precipitates.
Finally, Miller and Brenner [A-14] performed a detailed field-ion microscope study of a
neutron irradiated A302B pressure vessel steel. They found ,,-1015/cm3 Cu-rich and P-rich
zones in the form of thin discs or spheres. Voids and molybdenum carbides also formed, but
annealing cycles which removed the embrittlement did not remove these aggregates.
Accordingly, fine Cu-rich and P-rich zones or precipitates were established as causing
irradiation-induced pressure vessel embrittlement.

Nickel Based Alloys


Lam [A- 15] recently noted that coarsening kinetics of the 7' phase in HVEM or heavy ion
irradiated Ni-AI alloys may be strongly affected by spatial variations in defect production
rates and concentrations. Similar effects due to non-uniform damage were found in irradiated
Ni-Si alloys (see Section 14).
Brimhall et al. [A-6] irradiated NiA1, N%A1, NiMo, NiTi and NiTi2 with 2.5 MeV Ni ÷ ions
at a displacement rate of 3 x 10 -3 dpa/s. Irradiation was at or near room temperature, usually
to displacement levels in the 1-10 dpa range, although levels of 45 dpa were achieved in some
cases. Of the alloys, NiTi, NiTi2, and NiMo became amorphous, while NiAI and Ni3AI
remained crystalline.
Mori et al. [A-5] irradiated several Ni-based alloys in the 2 MV HVEM at a temperature
of 160 K and a fluence of 1 × 1024e/mZ.s (about 10-2 dpa/s). Presumably, displacement levels
of tens of dpa were reached. NiTi became amorphous while NiA1 and Ni3A1 remained
crystalline, as in the investigations of Brimhail et al. [A-6].

Refractory Alloys
Brimhall et al. [A-6] irradiated a series of refractory alloys with 2.5 MeV Ni + ions at a
displacement rate of 3 x 10-3dpa/s. Irradiation was at or near room temperature to
displacement levels usually in the 1-10 dpa range, although levels of 45 dpa were achieved
on some alloys. All the refractory alloys, TiFe, MoNi, Zr3A1, and Zr2A1 became amorphous.
Brimhall et al. attributed the crystalline-amorphous transition to a critical defect con-
centration giving such a high internal energy level that the amorphous state was preferred.
Compounds with low solubility ranges were predicted to be especially sensitive to de-
stabilization through changes in composition or defect concentration.
Mori et al. [A-5] recently irradiated several refractory alloys in the 2 MV HVEM at 160 K
and a flux of 1 x 1024e/m 2. s (about 10 -2 dpa/s). Presumably, displacement levels of tens of
dpa were reached. Of these alloys, TiCo2, TiFe2, TiMn2, MoNi, NbTNi 6, Zr2Ni, and Zr2AI
became amorphous, while CoTi, Cr2Ti, TiFe, and Zr3A1 remained crystalline. Attempts to
rationalize the behavior of these and the Cu-based and Ni-based alloys irradiated in this study
434 PROGRESS IN MATERIALS SCIENCE

on the basis of solubility ranges and relative positions of the components in the periodic table
were only partly successful.
As described in Section 15, ZraAI has been amorphized by Ar + ion irradiation at 300 K.
Electrons are apparently less effective than Ar ÷ and Ni ÷ ions in amorphizing.

Miscellaneous Alloys
Brimhall et al. [A-6] irradiated ReTa and U6Fe with 2.5 MeV Ni ÷ ions at a displacement
rate of 3 x 10 -3 dpa/s. Irradiation was at or near room temperature to displacement levels
typically in the 1-10 dpa range. Both compounds became amorphous as a result of the
irradiation.

References for Notes Added in Proof


A-1. MARTIN, G. (1984) Phase stability under irradiation: Ballistic effects, Phys. Rev. B, 30, 1424-1435.
A-2. CArlN, J. W. (1961) On spinodal decomposition, Acta Met. 9, 795-801.
A-3. BELLON,P. and MARTIN, G. (1984) A Lyupanov function approach to phase stability under irradiation. II.
The order-disorder transition in Ni4Mo (submitted to Phys. Rev. B).
A-4. RUSBmDOE,K. L. (1983) Dissolution of precipitates in Al~oe during 200 keV A1 ÷ ion irradiation, J. Nucl.
Mater. 119, 41-50.
A-5. MORI, H., FUJITA, H., TENDO, M. and FUJITA, M. (1984) Amorphous transition in intermetallic compounds
induced by electron irradiation, Scripta Met. 18, 783-788.
A-6. BmMHALL,J. L., KISSINGER, H. E. and CHARLOT, L. A. (1982) Radiation induced amorphous transition in
intermetallic compounds, in Metastable Materials Formation by Ion Implantation (eds S. T. Picraux and
W. J. Choyke) Elsevier, New York, pp. 235-241.
A-7. BERKOWITZ,A. E., JOHNSTON,W. B., MOGRO-CAMPERO,A. and WALTER,J. L. (1982) Structure and property
changes during ion bombardment of crystalline Fe75B25, in Metastable Materials Formation by Ion
Implantation (eds S. T. Picraux and W. J. Choyke) Elsevier, New York, pp. 195-202.
k-8. STEELE, L. E. (1975) Neutron Irradiation of Reactor Pressure Vessel Steels, International Atomic Energy
Agency, Vienna, 235 pp.
A-9. GRANT, S. P. and FORTNER, E. (August 1972) Effects of neutron fluence on steel weld metal for reactor vessels,
Metals Eng. Quarterly 12, 17-24.
A-10. BRENNER,S. S., WAGNER, R. and SPITZNAGEL,J. A. (1978) Field ion microscope detection of ultra-fine defects
in neutron irradiated Fe-0.34pct Cu alloy, Met. Trans. 9A, 1761-1764.
A-11. OOETTE,G. R. and SHEEKS, C. K. (1981) A model for displacement cascade induced microvoid and precipitate
formation in dilute iron-copper alloys, in Phase Stability during Irradiation (eds J. R. Holland, L. K. Mansur
and D. I. Potter) TMS-AIME, Warrendale, PA, pp. 415-436.
A-12. ODETTE, G. R. (1983) On the dominant mechanism of irradiation embrittlement of pressure vessel steels,
Scripta Met. 17, 1183-1188.
A-13. FmsIos, F., KAMPMANN, R., BEAVEN, P. A. and WAGNER, R. (1983) Influence of copper on the defect
microstructure and radiation strengthening of iron, in Proceedings of Conference on Dimensional Stability and
Mechanical Behavior of Irradiated Metals and Alloys, British Nuclear Energy Soc., Brighton, UK, pp.
171-174.
A-14. MILLER, M. K. and BRENNER, S. S. (1984) FIM/Atom probe study of irradiated pressure vessel steels, Res.
Mechanica 10, 161-168.
A-15. LAM, N. Q. (1984), research in progress.

Vous aimerez peut-être aussi