You are on page 1of 27

1.

07 Radiation Damage Using Ion Beams


G. S. Was
University of Michigan, Ann Arbor, MI, USA
R. S. Averback
University of Illinois at Urbana-Champagne, Urbana, IL, USA

ß 2012 Elsevier Ltd. All rights reserved.

1.07.1 Introduction 195


1.07.2 Motivation for Using Ion Beams to Study Radiation Damage 196
1.07.3 Review of Aspects of Radiation Damage Relevant to Ion Irradiation 197
1.07.3.1 Defect Production 197
1.07.3.2 Primary and Weighted Recoil Spectra 199
1.07.3.3 Damage Morphology 200
1.07.3.4 Damage Rate Effects 202
1.07.4 Contributions of Ion Irradiation to an Understanding of Radiation Effects 204
1.07.4.1 Electron Irradiations 204
1.07.4.1.1 Displacement threshold surfaces 204
1.07.4.1.2 Point defect properties 205
1.07.4.2 Ion Irradiations 206
1.07.4.2.1 The damage function 206
1.07.4.2.2 Freely migrating defects 207
1.07.4.2.3 Alloy stability under ion irradiation 207
1.07.4.2.4 Mechanical properties 209
1.07.4.2.5 Multiple ion beams 209
1.07.4.2.6 Swift ions 209
1.07.4.3 Comparison with Neutrons 211
1.07.5 Advantages and Disadvantages of Irradiations using Various Particle Types 215
1.07.5.1 Electrons 216
1.07.5.2 Heavy Ions 217
1.07.5.3 Light Ions 219
1.07.6 Practical Considerations for Radiation Damage Using Ion Beams 219
References 221

NWC Normal water chemistry


Abbreviations
PKA Primary knock-on atom
AES Auger electron spectroscopy
RCS Recoil collision sequence
APT Atom probe tomography
RIS Radiation induced segregation
bcc Body-centered cubic
SCC Stress corrosion cracking
BWR Boiling water reactor
STEM/EDS Scanning transmission electron
dpa Displacements per atom
microscopy/energy dispersive
fcc Face-centered cubic
spectrometry
FMD Freely migrating defect
TEM Transmission electron microscopy
FP Frenkel pair
IASCC Irradiation assisted stress corrosion
cracking
IGSCC Intergranular stress corrosion 1.07.1 Introduction
cracking
LWR Light water reactor Radiation effects research has been conducted using
MD Molecular dynamics a variety of energetic particles: neutrons, electrons,
NRT Norgett–Robinson–Torrens protons, He ions, and heavy ions. Energetic ions

195
196 Radiation Damage Using Ion Beams

can be used to understand the effects of neutron of conditions, which is precisely what is required for
irradiation on reactor components, and interest in investigations of the basic damage processes. Simula-
this application of ion irradiation has grown in recent tion by ions allows easy variation of the irradiation
years for several reasons including the avoidance of parameters such as dose, dose rate, and temperature
high residual radioactivity and the decline of neutron over a wide range of values.
sources for materials irradiation. The damage state and One of the prime attractions of ion irradiation is
microstructure resulting from ion irradiation, and thus the rapid accumulation of end of life doses in short
the degree to which ion irradiation emulates neutron periods of time. Typical neutron irradiation experi-
irradiation, depend upon the particle type and the ments in thermal test reactors may accumulate dam-
damage rate. This chapter will begin with a summary age at a rate of 3–5 dpa year1. In fast reactors, the
of the motivation for using ion irradiation for radiation rates can be higher, on the order of 20 dpa year1. For
damage studies, followed by a brief review of radia- low dose components such as structural components
tion damage relevant to charged particles. The contri- in boiling water reactor (BWR) cores that typically
bution of ion irradiation to our understanding of have an end-of-life damage of 10 dpa, these rates are
radiation damage will be presented next, followed by acceptable. However, even the higher dose rate of a
an account of the advantages and disadvantages of fast reactor would require 4–5 years to reach the peak
the various ion types for conducting radiation damage dose of 80 dpa in the core baffle in a pressurized
studies, and wrapping up with a consideration of prac- water reactor (PWR). For advanced, fast reactor con-
tical issues in ion irradiation experiments. cepts in which core components are expected to
receive 200 dpa, the time for irradiation in a test
reactor becomes impractical.
1.07.2 Motivation for Using Ion In addition to the time spent ‘in-core,’ there is an
Beams to Study Radiation Damage investment in capsule design and preparation as well
as disassembly and allowing for radioactive decay, add-
In the 1960s and 1970s, heavy ion irradiation was ing additional years to an irradiation program. Analysis
developed for the study of radiation damage pro- of microchemical and microstructural changes by
cesses in materials. As ion irradiation can be atom probe tomography (APT), Auger electron spec-
conducted at a well-defined energy, dose rate, and troscopy (AES) or microstructural changes by energy
temperature, it results in very well-controlled experi- dispersive spectroscopy via scanning transmission
ments that are difficult to match in reactors. As such, electron microscopy (STEM-EDS) and mechanical
interest grew in the use of ion irradiation for the property or stress corrosion cracking (SCC) evalua-
purpose of simulating neutron damage in support of tion can take several additional years because of the
the breeder reactor program.1–3 Ion irradiation and precautions, special facilities, and instrumentation
simultaneous He injection were also used to simulate required for handling radioactive samples. The result
the effects of 14 MeV neutron damage in conjunction is that a single cycle from irradiation through micro-
with the fusion reactor engineering program. The analysis and mechanical property/SCC testing may
application of ion irradiation (defined here as irradi- require over a decade. Such a long cycle length does
ation by any charged particle, including electrons) not permit for iteration of irradiation or material
to the study of neutron irradiation damage caught conditions that is critical in any experimental research
the interest of the light water reactor community to program. The long cycle time required for design and
address issues such as swelling, creep, and irradiation irradiation also reduces flexibility in altering irradia-
assisted stress corrosion cracking of core structural tion programs as new data become available. The
materials.4–6 Ion irradiation was also being used to requirement of special facilities, special sample
understand the irradiated microstructure of reactor handling, and long irradiation time make the cost
pressure vessel steels, Zircaloy fuel cladding, and for neutron irradiation experiments very high.
materials for advanced reactor concepts. In contrast to neutron irradiation, ion (heavy,
There is significant incentive to use ion irradiation light, or electrons) irradiation enjoys considerable
to study neutron damage as this technique has the advantages in both cycle length and cost. Ion irradia-
potential for yielding answers on basic processes tions of any type rarely require more than several
in addition to the potential for enormous savings in tens of hours to reach damage levels in the 1–100 dpa
time and money. Neutron irradiation experiments range. Ion irradiation produces little or no residual
are not amenable to studies involving a wide range radioactivity, allowing handling of samples without
Radiation Damage Using Ion Beams 197

the need for special precautions. These features application, ion irradiation provides a low-cost and
translate into significantly reduced cycle length and rapid means of elucidating mechanisms and screening
cost. The challenge then is to verify the equivalency materials for the most important variables.
between neutron and ion irradiation in terms of A final challenge is the volume of material that can
the changes to the microstructure and properties be irradiated with each type of radiation. Neutrons
of the material. The key question that needs to be have mean free paths on the order of centimeters
answered is how do results from neutron and charged in structural materials. One MeV electrons penetrate
particle irradiation experiments compare? How, for about 500 mm, 1 MeV protons penetrate about 10 mm,
example, is one to compare the results of a component and 1 MeV Ni ions have a range of less than 1 mm.
irradiated in-core at 288  C to a fluence of 1  1021 n Thus, the volume of material that can be irradiated
cm2 (E > 1 MeV) over a period of one year, with an ion with ions from standard laboratory-sized sources
irradiation experiment using 3 MeV protons at 400  C (TEMs, accelerators), is limited.
to 1 dpa (displacements per atom) at a dose rate of
105 dpa s1 (1 day), or 5 MeV Ni2þ at 500  C to
10 dpa at a dose rate of 5  103 dpa s1 (1 h)? 1.07.3 Review of Aspects of Radiation
The first question to resolve is the measure of radia- Damage Relevant to Ion Irradiation
tion effect. In the Irradiation assisted stress corrosion
cracking (IASCC) problem in LWRs, concern has cen- 1.07.3.1 Defect Production
tered on two effects of irradiation: radiation-induced The parameter commonly used to correlate the dam-
segregation of major alloying elements or impurities to age produced by different irradiation environments is
grain boundaries, which may cause embrittlement or the total number of displacements per atom (dpa).
enhance the intergranular stress corrosion cracking Kinchin and Pease7 were the first to attempt to deter-
(IGSCC) process, and hardening of the matrix that mine the number of displacements occurring during
results in localized deformation and embrittlement. irradiation and a modified version of their model
The appropriate measure of the radiation effect in the known as the Norgett–Robinson–Torrens (NRT)
former case would then be the alloy concentration at model8 is generally accepted as the international
the grain boundary or the amount of impurity standard for quantifying the number of atomic dis-
segregated to the grain boundary. This quantity is placements in irradiated materials.9 According to the
measurable by analytical techniques such as AES, APT, NRT model, the number of Frenkel pairs (FPs),
or STEM-EDS. For the latter case, the measure of the nNRT(T ), generated by a primary knock-on atom
radiation effect would be the nature, size, density, and (PKA) of energy T is given by
distribution of dislocation loops, black dots, and the
kED ðT Þ
total dislocation network, and how they impact nNRT ðT Þ ¼ ½1
the deformation of the alloy. Hence, specific and mea- 2Ed
surable effects of irradiation can be determined for where ED(T ) is the damage energy (energy of the
both neutron and ion irradiation experiments. PKA less the energy lost to electron excitation), Ed is
The next concern is determining how ion irradia- the displacement energy, that is, the energy needed to
tion translates into the environment describing neu- displace the struck atom from its lattice position, and k
tron irradiation. That is, what are the irradiation is a factor less than 1 (usually taken as 0.8). Integration
conditions required for ion irradiation to yield the of the NRT damage function over recoil spectrum and
same measure of radiation effect as that for neutron time gives the atom concentration of displacements
irradiation? This is the key question, for in a postirra- known as the NRT displacements per atom (dpa):
diation test program, it is only the final state of ðð
the material that determines equivalence, not the dpa ¼ fðEÞvNRT ðT ÞsðE; T ÞdT dE ½2
path taken. Therefore, if ion irradiation experiments
could be devised that yielded the same measures where f(E) is the neutron flux and s(E,T ) is the proba-
of irradiation effects as observed in neutron irradiation bility that a particle of energy E will impart a recoil
experiments, the data obtained in postirradiation energy T to a struck atom. The displacement damage is
experiments will be equivalent. In such a case, ion accepted as a measure of the amount of change to the
irradiation experiments can provide a direct substitute solid due to irradiation and is a much better measure
for neutron irradiation. While neutron irradiation will of an irradiation effect than is the particle fluence.
always be required to qualify materials for reactor As shown in Figure 1, seemingly different effects of
198 Radiation Damage Using Ion Beams

300 300

250 LASREF, 40 ⬚C 250 LASREF, 40 ⬚C


Yield stress change (MPa)

Yield stress change (MPa)


RTNS-II, 90 ⬚C RTNS-II, 90 ⬚C
OWR, 90 ⬚C OWR, 90 ⬚C
200 200

150 150

100 100

50 50

0 0
1017 1018 1019 1020 10−3 10−2
Neutron fluence, E > 0.1 MeV DPA

Figure 1 Comparison of yield stress change in 316 stainless steel irradiated in three facilities with very different neutron
energy flux spectra. While there is little correlation in terms of neutron fluence, the yield stress changes correlate well
against displacements per atom (dpa). Reprinted, with permission, from ASTM, copyright ASTM International, 100 Barr
Harbor Drive, West Conshohocken, PA 19428.

10−14
1014
Neutron flux/lethergy (n cm−2 s−1)

7.5 MeV tantalum


proton flux (ions cm−2 s−1)

1012 10−15
5 MeV nickel
Calculated dpa/(incident particle) (cm2)

1010
10−16 20 MeV carbon
Protons
108 ITER be first wall
HFIR target
FFTF mid-core 10−17
106 PWR 1/4-T RPV

1.3 MeV hydrogen


104 −9
10 10−7 10−5 10−3 10−1 10 10−18
Particle energy (MeV)

Figure 2 Energy spectrum for neutrons from a variety


10−19
of reactor types and a monoenergetic proton beam.
Reproduced from Stoller, R. E.; Greenwood, L. R. 14 MeV neutrons
J. Nucl. Mater. 1999, 271–272, 57–62.
−20 1 MeV neutrons
10
irradiation on low temperature yield strength for the
same fluence level (Figure 1(a)) and disappear when
10−21
dpa is used as the measure of damage (Figure 1(b)). 0 2 4 6 8 10 12
A fundamental difference between ion and neu- Distance into solid (m)
tron irradiation effects is the particle energy spectrum Figure 3 Displacement–damage effectiveness for various
that arises because of the difference in the way the energetic particles in nickel. Reproduced from Kulcinski,
particles are produced. Ions are produced in accel- G. L.; Brimhall, J. L.; Kissinger, H. E. In Proceedings of
erators and emerge in monoenergetic beams with Radiation-Induced Voids in Metals; Corbett, J. W.,
Ianiello, L. C., Eds.; USAEC Technical Information Center:
very narrow energy widths. However, the neutron Oak Ridge, TN, 1972; p 453, CONF-710601.
energy spectrum in a reactor extends over several
orders of magnitude in energy, thus presenting a
much more complicated source term for radiation Another major difference in the characteristics of
damage. Figure 2 shows the considerable difference ions and neutrons is their depth of penetration. As
in neutron and ion energy spectra and also between shown in Figure 3, ions lose energy quickly because
neutron spectra in different reactors and at different of high electronic energy loss, giving rise to a spa-
locations within the reactor vessel. tially nonuniform energy deposition profile caused
Radiation Damage Using Ion Beams 199

by the varying importance of electronic and nuclear recoil atom to other target atoms. The fraction of
energy loss during the slowing down process. Their recoils between the displacement energy Ed, and T is
penetration distances range between 0.1 and 100 mm ð
1 T
for ion energies that can practically be achieved by PðE; T Þ ¼ sðE; T 0 ÞdT 0 ½6
laboratory-scale accelerators or implanters. By virtue N Ed
of their electrical neutrality, neutrons can penetrate where N is the total number of primary recoils and
very large distances and produce spatially flat dam- s(E,T ) is the differential cross-section for a particle
age profiles over many millimeters of material. of energy E to create a recoil of energy T. The recoil
Further, the cross-section for ion–atom reaction is fraction is shown in Figure 4, which reveals only a
much greater than for neutron–nuclear reaction giving small difference between ions of very different masses.
rise to a higher damage rate per unit of particle fluence. Figure 5 shows the difference in the types
The damage rate in dpa per unit of fluence is propor- of damage that are produced by different types of
tional to the integral of the energy transfer cross-section
and the number of displacements per PKA, nNRT(T):
1.0
ð gE He
Rd
¼ sðE; T ÞnNRT ðT ÞdT ½3
Nf Ed 0.8 H
Kr
Fraction of recoils
where Rd is the number if displacements per unit vol-
Ar
ume per unit time, N is the atom number density, and f 0.6
is the particle flux (neutron or ion). In the case of Ne
neutron–nuclear interaction described by the hard-
0.4 Fraction of recoils
sphere model, eqn [3] becomes with energy above
  Ed and below T
Rd gE
¼ ss ½4 0.2
Nf 4Ed 1 MeV ions ® Cu

where g ¼ 4mM/(m þ M)2, M is the target atom mass, m 0


101 102 103 104
is the neutron mass, E is the neutron energy, and ss is T (eV)
the elastic scattering cross-section. For the case of ion–
Figure 4 Integral primary recoil spectra for 1 MeV
atom interaction described by Rutherford scattering,
particles in copper. Curves plotted are the integral fractions
eqn [3] becomes of primary recoils between the threshold energy and recoil
  energy, T from eqn [6]. Reproduced from Averback, R. S.
Rd pZ2 Z2 e4 M1 gE J. Nucl. Mater. 1994, 216, 49.
¼ 1 2 ln ; ½5
NI 4EEd M2 Ed
1 MeV electrons E 106
where e is the unit charge, M1 is the mass of the ion, and T = 60 eV
M2 is the mass of the target atom. As shown in Figure 3, e = 50−100% 105
for comparable energies, 1.3 MeV protons cause over Tn
100 times more damage per unit of fluence at the 104
1 MeV protons Ti
sample surface than 1 MeV neutrons, and the factor T = 200 eV
e = 25% 103
for 20 MeV C ions is over 1000. Of course, the damage Tp
depth is orders of magnitude smaller than that for 102
Te
neutron irradiation. 1 MeV heavy ions
T = 5 keV Ed 101
e = 4% E
1.07.3.2 Primary and Weighted
Recoil Spectra 1 MeV neutrons
T = 35 keV
A description of irradiation damage must also con- e = 2%
sider the distribution of recoils in energy and space.
Figure 5 Difference in damage morphology, displacement
The primary recoil spectrum describes the relative efficiency, and average recoil energy for 1 MeV particles of
number of collisions in which the amount of energy different types incident on nickel. Reproduced from
between T and T þ dT is transferred from the primary Was, G. S.; Allen, T. R. Mater. Char. 1994, 32, 239.
200 Radiation Damage Using Ion Beams

particles. Light ions such as electrons and protons 1.0


will produce damage as isolated FPs or in small Copper
clusters while heavy ions and neutrons produce dam-
0.8 Protons
age in large clusters. For 1 MeV particle irradiation of
copper, half the recoils for protons are produced with
energies less than 60 eV while the same number 0.6 Ne
Kr

W (T)
for Kr occurs at about 150 eV. Recoils are weighted
toward lower energies because of the screened
Coulomb potential that controls the interactions of 0.4
charged particles. For an unscreened Coulomb inter-
Neutrons
action, the probability of creating a recoil of energy 0.2
T varies as 1/T2. However, neutrons interact as
hard spheres and the probability of creating a recoil
of energy T is independent of recoil energy. 0
101 102 103 104 105 106 107
In fact, a more important parameter describing the T (eV)
distribution of damage over the energy range is a
combination of the fraction of defects of a particular Figure 6 Weighted recoil spectra for 1 MeV particles in
copper. Curves representing protons and neutrons are
energy and the damage energy. This is the weighted calculated using eqns [9] and [10], respectively. W(T ) for other
average recoil spectrum, W(E,T ), which weights the particles were calculated using Lindhard cross-sections
primary recoil spectrum by the number of defects or and include electronic excitation. Reproduced from
the damage energy produced in each recoil: Averback, R. S. J. Nucl. Mater. 1994, 216, 49.
ðT
1
W ðE; T Þ ¼ sðE; T 0 ÞED ðT 0 ÞdT 0 ½7 While heavy ions come closer to reproducing the
ED ðEÞ Ed
energy distribution of recoils of neutrons than do
ð T^ light ions, neither is accurate in the tails of the distri-
ED ðEÞ ¼ sðE; T 0 ÞED ðT 0 ÞdT 0 ½8 bution. This does not mean that ions are poor simu-
Ed lations of radiation damage, but it does mean that
^ is the maximum recoil energy given by damage is produced differently and this difference
where T
^ will need to be considered when designing an irradi-
T ¼ gEi ¼ 4EiM1M2/(M1 þ M2)2. Ignoring electron
ation program that is intended to produce microche-
excitations and allowing ED(T ) ¼ T, then the
mical and microstructural changes that match those
weighted average recoil spectra for Coulomb and
from neutron irradiation.
hard sphere collisions are
There is, of course, more to the description of
lnT  lnEd radiation damage than just the number of dpa.
WCoul ðE; T Þ ¼ ½9
^  lnEd
lnT There is the issue of the spatial distribution of damage
production, which can influence the microchemistry
T 2  Ed2 and microstructure, particularly at temperatures
WHS ðE; T Þ ¼ ½10
Ed2 where diffusion processes are important for micro-
structural development. In fact, the ‘ballistically’
Equations [9] and [10] are graphed in Figure 6 for
determined value of dpa calculated using such a
1 MeV particle irradiations of copper. The character-
displacement model is not the appropriate unit to
istic energy, T1/2 is that recoil energy below which
be used for dose comparisons between particle
half of the recoils are produced. The Coulomb
types. The reason is the difference in the primary
forces extend to infinity and slowly increase as the
damage state among different particle types.
particle approaches the target; hence the slow
increase with energy. In a hard sphere interaction,
the particles and target do not interact until their
1.07.3.3 Damage Morphology
separation reaches the hard sphere radius at which
point the repulsive force goes to infinity. A screened The actual number of defects that survive the dis-
Coulomb is most appropriate for heavy ion irradia- placement cascade and their spatial distribution
tion. Note the large difference in W(E,T ) between in solids will determine the effect on the irradiated
the various types of irradiations at E ¼ 1 MeV. microstructure. Figure 7 summarizes the effect of
Radiation Damage Using Ion Beams 201

Total dpa

Particle type
and energy

Loss to Freely migrating


displacement defects
cascades

Mutual Loss to sinks Loss at grain


recombination in matrix boundaries
outside of cascade
Void swelling
Defect diffusion loop structure Boundary
matrix chemistry structure
and micro
chemistry

Radiation-induced
segregation
Figure 7 History of point defects after creation in the displacement cascade.

damage morphology from the viewpoint of the grain Displacement cascade efficiency
boundary and how the defect flow affects radiation- (x)
induced grain boundary segregation. Of the total
defects produced by the energetic particle, a fraction Intracascade thermal
recombination (z )
appears as isolated, or freely migrating defects, and the
balance is part of the cascade. The fraction of the Surviving defect
‘ballistically’ produced FPs that survive the cascade fraction (QDF) (x – z )
quench and are available for long-range migration is
an extremely important quantity and is called the Clustered point defect Isolated point defect
fraction (CDF) (d i,v) fraction (IDF) (g i,v)
migration efficiency, e. These ‘freely migrating’ or ‘avail-
able migrating’ defects10 are the only defects that will
affect the amount of grain boundary segregation, Mobile Immobile
which is one measure of radiation effects. The migra- clusters clusters
tion efficiency can be very small, approaching a few
percent at high temperatures. The migration effi- Evaporating
defects
ciency, e, comprises three components:
Available
gi,v: the isolated point defect fraction, defects (li,v)
di,v: clustered fraction including mobile defect
clusters such as di-interstitials, and Figure 8 Interdependence of isolated point defects,
z: fraction initially in isolated or clustered form mobile defect clusters, and thermally evaporating defect
clusters that contribute to the fraction of surviving defects
after the cascade quench that is annihilated during
that are ‘available’ for radiation effects. Reproduced from
subsequent short-term (>1011 s) intracascade Zinkle, S. J.; Singh, B. N. J. Nucl. Mater. 1993, 199, 173.
thermal diffusion.
They are related as follows:
Due to significant recombination in the cascade,
e ¼ di þ g i þ z i ¼ d v þ g v þ z v ½11
only a fraction (30%) is free to migrate from the
Figure 8 shows the history of defects born as vacan- displacement zone. These defects can recombine out-
cies and interstitials as described by the NRT model. side of the cascade region, be absorbed at sinks in the
202 Radiation Damage Using Ion Beams

matrix (voids, loops), or be absorbed at the grain Table 1 Efficiency for producing freely migrating
boundaries, providing for the possibility of radiation- defects, g, in nickel by different kinds of irradiations (Ed ¼ 40
eV, riv ¼ 0.7 nm) using Lindhard’s analytical differential
induced segregation.
collision cross-section
The fraction of defects that will be annihilated
after the cascade quench by recombination events Irradiation  (%)
among defect clusters and point defects within the 1 MeV Hþ 24.0
same cascade (intracascade recombination), z, is 2 MeV Hþ 19.2
about 0.07, for a migration efficiency of 0.3 (see 2 MeV Liþ 16.9
below for additional detail).10 The clustered fraction, 1.8 MeV Neþ 8.7
d includes large, sessile clusters and small defect 300 keV Niþ 2.3
3 MeV Niþ 3.8
clusters that may be mobile at a given irradiation 3.5 MeV Krþ 3.0
temperature and will be different for vacancies and 2 keV Oþ 9.8
interstitials. For a 5 keV cascade, di is about 0.06 and
dv is closer to 0.18.10 Some of these defects may be Source: Naundorf, V. J. Nucl. Mater. 1991, 182, 254.
able to ‘evaporate’ or escape the cluster and become
‘available’ defects (Figure 8). radius so that the nearby FPs neither recombine nor
This leaves g, the isolated point defect fraction cluster. The model follows each generation of the
that are available to migrate to sinks, to form clus- collision and calculates the fraction of all defects
ters, to interact with existing clusters, and to partic- produced that remain free. Results of calculation
ipate in the defect flow to grain boundaries that using the Naundorf model are shown in Table 1 for
gives rise to radiation-induced segregation. Owing several ions of varying mass and energy. Values of Z
to their potential to so strongly influence the irra- range between 24% for proton irradiation to 3% for
diated microstructure, defects in this category, along heavy ion (krypton) irradiation. Recent results,13
with defects freed from clusters, make up the freely however, have shown that the low values of FMD
migrating defect (FMD) fraction. Recall that electrons efficiency for heavy ion or neutron irradiation cannot
and light ions produce a large fraction of their be explained by defect annihilation within the parent
defects as isolated FPs, thus increasing the likeli- cascade (intracascade annihilation). In fact, cascade
hood of their remaining as isolated rather than clus- damage generates vacancy and interstitial clusters
tered defects. Despite the equivalence in energy that act as annihilation sites for FMD, reducing the
among the four particle types described in Figure 5, efficiency of FMD production. Thus, the cascade
the average energy transferred and the defect pro- remnants result in an increase in the sink strength
duction efficiencies vary by more than an order of for point defects and along with recombination in the
magnitude. This is explained by the differences in original cascade, account for the low FMD efficiency
the cascade morphology among the different parti- measured by experiment.
cle types. Neutrons and heavy ions produce dense
cascades that result in substantial recombination
1.07.3.4 Damage Rate Effects
during the cooling or quenching phase. However,
electrons are just capable of producing a few widely As differences in dose rates can confound direct
spaced FPs that have a low probability of recombi- comparison between neutron and ion irradiations, it
nation. Protons produce small widely spaced cas- is important to assess their impact. A simple method
cades and many isolated FPs due to the Coulomb for examining the tradeoff between dose and temper-
interaction and therefore, fall between the extremes ature in comparing irradiation effects from different
in displacement efficiency defined by electrons and particle types is found in the invariance requirements.
neutrons. For a given change in dose rate, we would like to know
The value of g has been estimated to range from what change in dose (at the same temperature) is
0.01 to 0.10 depending on PKA energy and irradia- required to cause the same number of defects to be
tion temperature, with higher temperatures resulting absorbed at sinks. Alternatively, for a given change
in the lower values. Naundorf12 estimated the freely in dose rate, we would like to know what change in
migrating defect fraction using an analytical treat- temperature (at the same dose) is required to cause
ment based on two factors: (1) energy transfer to the same number of defects to be absorbed at
atoms is only sufficient to create a single FP, and sinks. The number of defects per unit volume, NR,
(2) the FP lies outside a recombination (interaction) that have recombined up to time t, is given by Mansur14
Radiation Damage Using Ion Beams 203

ðt invariant at a fixed dose, the following relationship


NR ¼ Riv Ci Cv dt ½12 between ‘dose rate and temperature’ must hold:
   
0 kT12 K2
Evm þ2Evf ln K1
where Riv is the vacancy–interstitial recombination T2  T1 ¼     ½16
coefficient and Ci and Cv are interstitial and vacancy 1  EvmkT 1
þ2Evf ln K2
K1
concentrations, respectively. Similarly, the number of
where Evf is the vacancy formation energy. In the steady-
defects per unit volume that are lost to sinks of type j,
state recombination dominant regime, for NS to be invariant
NSj, up to time t, is
at a fixed temperature, the following relationship
ðt between ‘dose (F) and dose rate’ must hold:
NSj ¼ kSj Cj dt ½13  1=2
F2 K2
0 ¼ ½17
F1 K1
where kSj is the strength of sink j and Cj is the sink
concentration. The ratio of vacancy loss to interstitial Finally, in the steady-state recombination dominant regime,
loss is for NS to be invariant at a fixed dose rate, the following
relationship between ‘dose and temperature’ must hold:
RS ¼
NSv
½14    
2kT12
NSi Evm ln F 2
F1
T2  T1 ¼     ½18
where j ¼ v or i. The quantity NS is important in 1  kT 1
ln F2
Evm F1
describing the microstructural development involving
total point defect flux to sinks (e.g., RIS), while RS is the Figure 9 shows plots of the relationship between the
relevant quantity for the growth of defect aggregates ratio of dose rates and the temperature difference
such as voids that require partitioning of point defects required to maintain the same point defect absorption
to allow growth. In the steady-state recombination dominant at sinks (a), and the swelling invariance (b).
regime, for NS to be invariant at a fixed dose, the follow- The invariance requirements can be used to
ing relationship between ‘dose rate (Ki) and temperature prescribe an ion irradiation temperature–dose rate
(Ti)’ must hold: combination that simulates neutron radiation. We
 2   take the example of irradiation of stainless steel
kT1 K2
Evm ln K1 under typical BWR core irradiation conditions of
T2  T1 ¼     ½15
1  kT 1
ln K2 4.5  108 dpa s1 at 288  C. If we were to conduct
Evm K1
a proton irradiation with a characteristic dose rate of
where Evm is the vacancy migration energy. In the 7.0  106 dpa s1, then using eqn [15] with a vacancy
steady-state recombination dominant regime, for RS to be formation energy of 1.9 eV and a vacancy migration

700 50

600
40
Eνm = 0.5
DTemperature (⬚C)
DTemperature (⬚C)

500 Em
ν = 0.5
30
400 1.0
1.5
300 20

200 1.0
10
1.5
100

0 0
1 10 100 1000 1 10 100 1000
(a) Ratio of dose rates (b) Ratio of dose rates

Figure 9 Temperature shift from the reference 200  C required at constant dose in order to maintain (a) the same point
defect absorption at sinks, and (b) swelling invariance, as a function of dose rate, normalized to initial dose rate. Results are
shown for three different vacancy migration energies and a vacancy formation energy of 1.5 eV. Adapted from Mansur, L. K.
J. Nucl. Mater. 1993, 206, 306–323; Was, G. S. Radiation Materials Science: Metals and Alloys; Springer: Berlin, 2007.
204 Radiation Damage Using Ion Beams

energy of 1.3 eV, the experiment will be invariant in the good agreement that is possible, while on the other,
NS with the BWR core irradiation (e.g., RIS) at a the extreme caution that is necessary in extrapolating
proton irradiation temperature of 400  C. Similarly, results of ion irradiations to long-term predictions
using eqn [16], a proton irradiation temperature of of materials evolution in a nuclear environment.
300  C will result in an invariant RS (e.g., swelling
or loop growth). For a Ni2þ ion irradiation at a dose
1.07.4.1 Electron Irradiations
rate of 103 dpa s1, the respective temperatures are
675  C (NS invariant) and 340  C (RS invariant). In The unique feature of electron irradiations in compari-
other words, the temperature ‘shift’ due to the higher son to ions and neutrons is that they create defects in
dose rate is dependent on the microstructure feature very low-energy recoil events. As a consequence, nearly
of interest. Also, with increasing difference in dose all FPs are produced in isolation. This has been of
rate, the DT between neutron and ion irradiation foremost importance in developing our understanding
increases substantially. The nominal irradiation tem- of radiation damage, as it made studies of defect crea-
peratures selected for proton irradiation, 360  C and tion mechanisms as well as the fundamental properties
for Ni2þ irradiation, 500  C represent compromises of FPs possible. Recall that the properties of vacancies
between the extremes for invariant NS and RS. andvacancy clusters, for example, formation and migra-
tion energies, stacking fault energies, etc., could be
determined from quenching studies. It is not possible,
1.07.4 Contributions of Ion however, to quench in interstitials in metals. Very little
Irradiation to an Understanding of was therefore known about this intrinsic defect prior
Radiation Effects to about 1955 when irradiation experiments became
widely employed. In this section, we highlight some of
Ion irradiations have been critical to the development the key findings derived from these past studies.
of both our fundamental and applied understanding
of radiation effects. As discussed in Sections 1.07.2 1.07.4.1.1 Displacement threshold surfaces
and 1.07.3, it is the flexibility of such irradiations and The creation of a stable FP requires that a lattice
our firm understanding of atomic collisions in solids atom receives an energy greater than Tm, which is
that afford them their utility. Principally, ion irradia- the minimum displacement energy. This value has
tions have enabled focused studies on the isolated been determined experimentally in many materials
effects of primary recoil spectrum, defect displace- by measuring the change in some physical property,
ment rate, and temperature. In addition, they have such as electrical resistivity or length change, as a
provided access to the fundamental properties of function of maximum recoil energy of a target atom.
point defects, defect creation, and defect reactions. Such experiments are practical only for electron
In this section, we highlight a few key experiments irradiations for which recoil energies can be kept
that illustrate the broad range of problems that can low, but with the irradiation particles still penetrating
be addressed using ion irradiations. We concentrate deeply into, or through, the specimen. Typical values
our discussion on past ion irradiations studies that are shown in Table 2.
have provided key information required by modelers As a crystal is not homogeneous, the threshold
in their attempts to predict materials behavior in energy depends on the crystallographic direction in
existing and future nuclear reactor environments, which the knock-on atom recoils. The anisotropy of
and particularly information that is not readily the threshold energy surface has been mapped out in
available from neutron irradiations. In addition, we various crystals by measuring the production rate of
include a few comparative studies between ion defects as a function of both the electron energy, near
and neutron irradiations to illustrate, on one hand, threshold, and the orientation of single crystalline

Table 2 Minimum displacement energies in pure metals, semiconductors, and stainless steel (SS)

Materials Al Cgraph Cu Fe Ge Mo Ni W Si SS

Tm (eV) 16 25 19 17 15 33 23 41 13 18

Source: Lucasson, P. In Fundamental Aspects of Radiation Damage in Metals; Robibnson, M. T., Young, F. W., Jr., Eds.; ERDA Report
CONF-751006; 1975; p 42; Andersen, H. H. Appl. Phys. 1979, 18, 131.
Radiation Damage Using Ion Beams 205

specimens with respect to the electron beam direc- The primary knock-on atom in an RCS recoils in
tion.15,16 The total cross-section for FP production the direction of its nearest neighbor, h110i in fcc
rate is given by the expression crystals, and replaces it, with the neighbor recoiling
2ðp p=2
also in the h110i and replacing its neighbor. A vacancy
ð
dsðy2 ; E1 Þ df2 is left at the primary recoil site, and an interstitial is
sd ðY1 ; F1 ; E1 Þ ¼ created at the end of the sequence. Replacement
dy2 2p ½19
0 0 sequences are the most efficient way to separate
nðY2 ; F2 ; T Þdy2 the interstitial far enough from its vacancy, 2–3
interatomic spacings, for the FP to be stable. While
where the subscripts 1 and 2 refer to incoming elec- the lengths of these sequences are still debated, it is
tron and recoiling ion, respectively, and Y1, F1, Y2, clear that the mechanism results in both defect pro-
F2 are the polar and azimuthal angles of the electron duction and atomic mixing. For neutron irradiations,
beam relative to the crystal axis; y2, f2, are these higher energy recoils are numerous, and the average
same angles relative to the beam direction; n is the displacement energy, Ed, becomes more relevant for
anisotropic damage function. Near threshold, n ¼ 1 calculations of defect production (see eqn [1]). This
for T > Tm, and 0 for T < Tm. By measuring the value, which can be obtained by averaging over the
production rate for many sample orientations and threshold displacement energy surface, is usually dif-
energies, the damage function can be obtained using ficult to determine experimentally. A rough estimate,
eqn [19], although various approximations are required however, can be obtained from, Td  1.4Tm in fcc
in the deconvolution. The results are illustrated metals and 1.6Tm in bcc metals.19
in Figure 10 for Cu.17 It is noteworthy that the
minimum threshold energy is located in the vicinity 1.07.4.1.2 Point defect properties
of close-packed directions. This is also true for bcc As FPs are produced in isolation during electron
metals. The anisotropy reflects the basic mechanism irradiation, the properties of single point defects
of defect production, viz., replacement collision and their interactions with impurities and sinks
sequences (RCSs), which had been identified by can be systematically investigated. An example is
molecular dynamics simulations as early as 1960.18 shown in Figure 11(a), where the results of low-
temperature isochronal annealing of Cu are shown
[111] following 1.4 MeV electron irradiation at 6 K.20
Recovery is observed to occur in ‘stages.’ These stud-
55 56 ies have revealed that interstitial atoms become
406 150 mobile at very low temperatures, always below
43 60 61 100 K, in so-called Stage I, while vacancies become
47
a 683 81 215 253 mobile at higher temperatures, Stage III. The various
18
43 43 31 24 26
substages IA–IE seen in Figure 11(a) arise from the
2 interaction between interstitial–vacancy pairs, which
b 208 103 11 4 0
22 are produced in close proximity. Stage IE refers to
28 45 25 20 22
6 9 69 3
the free migration of interstitials in the lattice, away
3 4
from its own vacancy, and annihilation at distant
30 26 24 23 20 20
27 vacancies; these interstitials are freely migrating as
11 10 3 4 6 3 4
discussed earlier. For comparison, Stage I annealing
23 25 25 29 23 21 20
23 of Cu following neutron irradiation is shown in
4 4 5 12 9 4 4 3
Figure 11(b).21 Notice that the close pair substages
23 23 25 29 23 21 20 22 are suppressed during neutron irradiation, illustrat-
18
2 2 2 4 10 3 3 2 3
ing the dramatic difference in the defect production
[100] [110] process for these types of irradiation. Similarly,
(a)
annealing studies on electron-irradiated Al doped
Figure 10 Displacement energy threshold surface for Cu. with Mg or Ga impurities are shown in Figure 12.22
The general anisotropy is typical of all fcc metals, although For these, it is observed that Stage I recovery is
specific values vary. bcc metals show similar behavior of
minima along close-packed directions. Reproduced from
suppressed as interstitials trap at impurities and do
King, W. E.; Merkle, K. L.; Meshii, M. Phys. Rev. B 1981, not recombine. The recovery at higher temperature,
23, 6319. in Stage II, reveals distinct subannealing stages.
206 Radiation Damage Using Ion Beams

IB ID 40

ID
7 IC Cu
30

dDr (% K–1)
(% K–1)

5
dDr

dT
dT

20
4 IE

°
°

IA IC

1
− Dr
1
− Dr

3
10 IB
2
IA
1

0 0
10 20 30 40 50 60 70 80 10 20 30 40 50 60 70
(a) T (K) (b) T (K)

Figure 11 Low-temperature isochronal annealing of Cu following (a) electron (reproduced from Corbett, J. W.; Smith, R. B.;
Walker, R. M. Phys. Rev. 1959, 114, 1452) or (b) fast neutron irradiation (reproduced from Burger, G.; Isebeck, K.;
Volkl, J.; Schilling, W.; Wenzl, H. Z. Angew. Phys. 1967, 22, 452).

threshold energies using low energy protons, to tens


100
Dr [nW.cm]
of keV using MeV self-ions. In addition, defect pro-
90 Al-0.06° at.%Mg °
Al-0.085 at.%Ga
2.5 duction rates can be varied over many orders of
magnitude, reaching values over 0.1 dpa s1. More-
80 2.7
Al-(99.995%pure) 3.3
70
over, by using more than one ion beam, the primary
Dr/Dr0 (%)

60
recoil spectrum can be tailored to closely match that
50
produced by an arbitrary fission neutron spectrum.
40
30
1.07.4.2.1 The damage function
20
Calculations of defect production, eqn [2], require
10
knowledge of the damage function, n(T ). While it is
0
10 30 100 300 not possible to measure this function directly, as no
T (K)
irradiation creates monoenergetic recoils except near
Figure 12 Recovery of electrical resistivity in Al, the surface, it can be obtained by measuring defect
Al–0.06 at.% Ga, and Al–0.085 at.% Ga following 1 MeV production for a wide range of ion irradiations and
electron irradiation. Reproduced from Garr, K. R.; Sosin,
subsequently deconvoluting eqn [3]. Low-energy
A. Phys. Rev. 1969, 162, 669.
light ions, for example, weight the recoil spectrum
near the threshold energy, 25–100 eV, while more
These annealing stages are generally attributed to
energetic heavy ions weight it at high energies.
either the interstitial dissociating from the impurity,
Results are shown for Cu in Figure 13. Here, electri-
or the interstitial–impurity complex migrating to a
cal resistivity measurements are employed to monitor
vacancy or a defect sink. Migrating interstitial–solute
the absolute number of FPs produced per unit dose
complexes lead to segregation. A compilation of the
of irradiation. Included in this figure are the damage
properties of point defects for many metals, and their
efficiency function, x(T1/2), deduced from the experi-
interactions with impurities can be found in Ehrhart.23
ments and x(T ) calculated using molecular dynamics
This information has played a crucial role in develop-
computer simulation. The damage efficiency function
ing an understanding of radiation damage in more
is defined as
complex engineering alloys and under more complex
irradiation conditions. nðT Þ ¼ xðT ÞnNRT ðT Þ; ½20
where nNRT(T ) is the NRT damage function defined
1.07.4.2 Ion Irradiations
by eqn [1]. The good agreement between experiment
Ion irradiations are the most flexible method for and simulations illustrates that the damage function
irradiating materials. As discussed in Section 1.07.2, in Cu is now well understood. This is now true for
the primary recoil spectrum can be shifted from near many other pure metals as well.24 In alloys and ceramic
Radiation Damage Using Ion Beams 207

Cu
1 H 1.0 1 MeV H
Experiment
Calculation

Relative efficiency
0.8 He 0.8
Li
0.6
x

CN 0.6
O Cu
Ne Kr 2 MeV He
0.4 Ar Fe
Ag Bi 0.4 2 MeV Li
FF
0.2
FN 0.2
MD simulation 3 MeV Ni
0 3.25 MeV Kr
10 2
10 3
10 4 5
10 0
T, T1/2 (eV) 102 103 104 105 106 107
T1/2 (eV)
Figure 13 Damage function efficiency factor of Cu
(see eqn [20]) showing the decrease in efficiency versus Figure 14 Relative efficiencies for producing freely
cascade energy. The experimental data (solid squares) migrating defects plotted as a function of the characteristic
represent efficiencies for different ion irradiations plotted recoil energy, T1/2. Reproduced from Rehn, L. E.;
versus the characteristic cascade energy for the irradiation, Okamoto, P. R.; Averback, R. S. Phys. Rev. 1984,
T1/2 (see text). The open triangles represent the efficiency B30, 3073.
versus cascade energy, T, obtained by molecular dynamics
(MD) simulation. The open circles represent the calculated
efficiencies for the different irradiations using the MD defects, extracting quantitative information about
efficiency function and eqn [2]. Reproduced from Averback, freely migrating defects from such experiments is
R. S.; de la Rubia, T. D. In Solid State Physics; Ehrenreich, difficult. These measurements, unlike the damage
H., Spaepen, F., Eds.; Academic Press: New York, 1998; pp function, require very high doses, and several dpa;
281–402.
the buildup of the sink structure must be adequately
taken into account. It is also difficult to estimate,
materials, however, the damage function remains for example, how many interstitials are required to
poorly known. transport one Si atom to the surface. We mention in
passing that experiments performed using ordering
1.07.4.2.2 Freely migrating defects kinetics in order–disorder alloys have provided a
The damage function refers to the number of FPs more direct measure of the number of freely migrat-
created within the first several picoseconds of the ing defects (vacancies in this case), as these experi-
primary recoil event. At longer times, defects migrate ments require doses less than 107 dpa so that no
from their nascent sites and interact with other damage build-up can occur.25 These experiments
defects and microstructural features. As noted earlier, show similar effects of primary recoil spectrum on
many radiation effects, such as radiation-enhanced the fraction of freely migrating defects, although the
diffusion, segregation, and void swelling, depend more fractions of such defects were found to be somewhat
strongly on the number of defects that escape their higher in these experiments, 5–10%. These frac-
nascent cascades and migrate freely in the lattice before tions are in good agreement with radiation-enhanced
annihilating, trapping, or forming defect clusters. The diffusion experiments using self-ions on Ni, when the
same general approach used to determine the damage effect of sink strength is taken into account.26
function has been employed to determine the relative
fraction of freely migrating defects, that is, e/nNRT, as 1.07.4.2.3 Alloy stability under ion irradiation
illustrated by Figure 14. Here, the relative number of Irradiation of materials with energetic particles drives
Si atoms segregating to the surface during irradiation, them from equilibrium, and in alloys, this becomes
per dpa, is plotted versus a characteristic energy of manifest in a number of ways. One of them concerns
the recoil spectrum, T1/2. It is seen that the fraction nonequilibrium segregation. The creation of large
decreases rapidly with increasing recoil energy. Simi- supersaturations of point defects leads to persistent
lar experiments were performed using radiation- defect fluxes to sinks. In many cases, these point defect
enhanced diffusion, as described in Section 1.07.2. fluxes couple with solutes, resulting in either the
While ion irradiation has proved extremely useful enrichment or depletion of solutes at these sinks.
in illustrating the spectral effects on freely migrating This effect was first discovered by using in situ electron
208 Radiation Damage Using Ion Beams

irradiations in a high voltage electron microscope,27 form in the sample. At the location of peak damage,
and it has been systematically investigated subse- the concentration of interstitials is the highest, and
quently using ion irradiations,28 as the surface sink hence these defects flow outward from this region.
provides a convenient location to measure composi- These interstitials form interstitial–solute complexes
tion changes. Unlike neutron irradiation, moreover, with Si, resulting in a Si flux out of this area as well,
the damage created by ions is generally inhomoge- depleting the region of Si. As a consequence, a region
neous, reaching a peak level at some depth in the depleted of Ni3Si precipitates is observed at the
sample. As a consequence, point defect fluxes ema- peak damage depth. Note too that the surface sink
nate from these regions. An example of this effect is for interstitials leads to enrichment of Si, resulting in
shown in Figure 15 where a Ni–12.7 at.% Si alloy a surface layer of Ni3Si. The region just below the
was irradiated with protons. As the alloy is supersat- surface accordingly becomes depleted of Si, leaving a
urated with Si prior to irradiation, Ni3Si precipitates zone depleted of Ni3Si precipitates.
While irradiation induced segregation can lead to
nonequilibrium segregation and precipitation in sin-
gle phase alloys, irradiation can also lead to dissolu-
tion of precipitates in nominally two-phase alloys.
An interesting example of this behavior concerns
Ni–12 at.% Al alloys irradiated with 300 keV Ni
ions.29 These alloys were first annealed at high tem-
peratures to develop a two-phase structure of Ni3Al
(g0 ) and Ni–10.5 at.% Al (g). The initial precipitate
size, depending on the annealing time was 2.5 or
Ni3Si surface film Peak 4.6 nm. As shown in Figure 16, the precipitates dis-
Ni plating damage
Bombarded surface region order during irradiation at room temperature, owing
to atomic mixing in cascades. The rate of disordering
Figure 15 Behavior of silicon in a Ni–12.7 Si alloy
following irradiation with protons. Note the region depleted depends on the size of the precipitates, being slowest
of Ni3Si precipitates at the peak damage location and just for homogeneous Ni3Al sample and fastest in the
below the surface. Courtesy of P. R. Okamoto. alloy with the smallest precipitates. The authors

1.0
1.0 Ni3AI Ni3AI
Degree of LRO S/S0

Degree of LRO S/S0

NiAI (r = 4.6 nm) NiAI (r = 4.6 nm)


0.8 NiAI (r = 2.5 nm)
0.6
0.5
0.4
550 ⬚C
0.2
450 ⬚C
0.0 0.0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.0 2.0 4.0 6.0
(a) Irradiation dose F (dpa) (b) Irradiation dose F (dpa)

0.8
0 dpa 2 dpa 5 dpa
f(r )

0.4

0.0
0.0 4.0 8.0 0.0 4.0 8.0 0.0 4.0 8.0
(c) Radius (nm)
Figure 16 (a) Disordering rate Ni3Al precipitates in two-phase Ni–12 at.% Al alloys and homogeneous Ni3Al during 300 keV
Ni bombardment at room temperature; (b) same as (a) but irradiation at 550  C; (c) size distribution of Ni3Al precipitates
after irradiation to two doses. After 5 dpa, a steady state size is obtained. Reproduced from Schmitz, G.; Ewert, J. C.;
Harbsmeier, F.; Uhrmacher, M.; Haider, F. Phys. Rev. B 2001, 63, 224113.
Radiation Damage Using Ion Beams 209

suggest that the reason for this dependence on pre- ions with lower energies, E  1–4 MeV, have also
cipitate size is that atomic mixing reduces the been used in deformation studies; for these, however,
concentration of Al in the precipitates, which specimen must be very thin, 200 nm, and effects of
thereby accelerates the disordering. When the same the surface must be taken into account.34,35
irradiation is performed at higher temperatures, and
radiation-enhanced diffusion takes place, the system 1.07.4.2.5 Multiple ion beams
does not completely disorder, but rather remains One of the difficulties in using ion beams to simulate
partially ordered, owing to a competition between neutron irradiation damage is the potential for miss-
disordering in the displacement cascades and reor- ing certain synergistic behaviors in the damage evo-
dering by radiation-enhanced diffusion. Noteworthy, lution. For example, neutron irradiation leads to
however, is the size of the precipitate, as shown in transmutation products and the generation of He
Figure 16(c), where it is observed that the precipi- and fission gases in addition to displacement damage.
tates initially shrink in size, but then reach a steady Generation of gas is particularly relevant to 14 MeV
state radius. Therefore, unlike in thermal aging, pre- neutron irradiation for which large amounts of
cipitates in irradiated alloys can reach a stable steady He and H are produced. Ion beams, however, offer
state size that is a function of irradiation intensity and the opportunity of using two or even three beams
temperature. Similar behavior has been observed in simultaneously and thus to tailor test irradiations to
two-phase immiscible alloys in which case a steady meet expected reactor conditions; see, for example,
state size of precipitates is formed.30 This so-called Serruys et al.36 This is often not possible in existing
‘patterning’ phenomenon has been explained on the test reactor facilities, and the building of new test
basis of a competition between disordering by atomic facilities for fusion machines has been formidably
mixing in energetic collision events and reordering expensive. The application of multiple ion beams is
during thermally activated diffusion. For patterning, illustrated in Figure 17 in a study of void swelling in
however, it is required that the atomic relocation vanadium. Here, the synergistic effects of simulta-
distances during collisional mixing are significantly neously implanting 350 keV H and 1 MeV He, while
larger than the nearest neighbor distance. An inter- irradiating with 12 MeV Ni ions are shown. Without
esting consequence of this requirement in regard to the He beam, swelling is negligible, even with the
the present discussion of using ion irradiation to implantation of H, but with it, the H greatly enhances
simulate neutron damage is that electron and proton the swelling. H implantation, on the other hand, is
irradiations, which do not produce energetic cascades seen to reduce the density of cavities.
or long relocation distances, should not induce com-
positional patterning, but heavy ions and fast neutron 1.07.4.2.6 Swift ions
irradiation, which do produce cascades, will cause An important contribution to the damage in nuclear
patterning. Further details can be found in Enrique31 fuels derives from fission fragments. There are two
and Enrique et al.32 groups of fission products: one group with atomic
number near 42 (Mo) and energy 100 MeV and
1.07.4.2.4 Mechanical properties the other with atomic number near 56 (Ba) and
Measurements of mechanical properties on irra- energy 70 MeV. The maximum electronic stopping
diated materials usually require bulk samples and powers of these energetic particles, 18 keV nm1
therefore neutron irradiation. Ion beams, however, for the heavier and 22 keV nm1 for the lighter, are
can be employed for some measurements, such as far greater than their respective nuclear stopping
plastic deformation. Typically, these experiments powers. Similar to ion irradiation studies described
employ high energy protons, E >  2 MeV, or He above, where the primary recoil spectrum can be
ions, E > 7 MeV, as these particles can penetrate systematically varied, the masses and energies of
through thin foils, such as Fe or steel, that are greater ions can be varied to examine effects of electronic
than 15 mm in thickness. Moreover, displacement stopping power. An example is shown in Figure 18
rates 105 dpa s1 are obtainable without excessive where the electronic stopping power is plotted as a
beam heating.33 Deformation experiments have also function of energy (per nucleon) for different ion
been performed using GeV heavy ions, as these pen- irradiations of UO2. The two boxes in the figure
etrate targets several microns in thickness. The dis- indicate stopping powers associated with the fission
placement rates, however, are low as most of the beam fragments and the heavy particle recoils of a emit-
energy is lost through electronic excitations. Heavy ters. One of the questions addressed by such studies
210 Radiation Damage Using Ion Beams

20

Cavity density (1020 m−3)


20

Swelling (%)
15
10
20 10
20
5
15
0
0
20 10 10
20
He

He
10
(ap

10 5

(ap
pm

pm
0
)d

0 0 0

)d
20
pa

20 10

pa
10 0
–1

0 pa–1

–1
p p m ) d p a–1 (b) H (a p p m ) d
(a) H(a
Figure 17 Cavity volume fraction (a) and cavity density (b) in pure vanadium irradiated with 12 MeV Ni3þ ions to 30 dpa
at 873 K with and without simultaneous irradiation of He and H. Reproduced from Sekimura, N.; Iwai, T.; Arai, Y.; et al.
J. Nucl. Mater. 2000, 283–287, 224–228.

70
dE/dx (keV nm–1)

Energy (MeVamu–1)

238U
20 1.4
1.2
60 Light FP Heavy FP 1.0 208Pb
0.8
10
0.6
50 0.4
197Au
235
U Fission 0.2
0 0.0
−8 −6 −4 −2 0 2 4 6 8
dE/dx (keV nm–1)

40 FPs range (μm)


129Xe

116Sn
30
GANIL GSI 106Cd
HMI TASCC 100Mo
127I
20 dE/dxFP

70Zn
Recoils
10
Zn70Zn
Efission
0
10−5 10−4 10−3 10−2 10−1 1 10 102
Energy (MeVamu–1)
Figure 18 Plot of dE/dx as a function of the energy for a series of ions. The circle indicates the conditions for 72 MeV
ions of 127I. The two large squares show dE/dx representative of fission products and for the heavy recoil atoms of
a-decaying actinides. The inset shows the energy loss and the remaining energy of typical light and heavy fission
products along their range of 7 mm length. Reproduced from Matzke, Hj.; Lucuta, P. G.; Wiss, T. Nucl. Instrum. Meth. B
2000, 166–167, 920.
Radiation Damage Using Ion Beams 211

has been the formation of fission fragment tracks. 24 8


Tracks have not yet been observed in the bulk of CP-316 SS
Protons at 360 ⬚C to 1.0 dpa
UO2 due to fission; however, by using ion irradiation, Neutrons at 275 ⬚C to
7
the stopping powers could be increased. The dashed 1.1⫻1021 n cm–2 (~1.5 dpa)
20
line at 29 keV nm1 in Figure 18 represents the 6

Measured Cr or Ni (wt%)
threshold stopping power for track formation.37

Measured Si (wt%)
5
This value is 30% greater than the maximum for
16 Cr
fission fragments, thus helping to explain why fission
4
fragment tracks are not seen in the bulk. Such tracks
are observed, however, close to the surface. They are Ni
3
explained by fission products passing near or parallel 12
to the surface and creating shock waves which inter- 2
act with the surface.38 These studies have also been Si
useful in gaining important data for understanding 8 1
fission gas evolution in nuclear fuels. For example,
72 MeV iodine ions (see Figure 18), approximate 0
−12 −8 −4 0 4 8 12
very closely the stopping power of fission fragments. Distance from grain boundary (nm)
Such studies have shown that 72 MeV I irradiations
Figure 19 Comparison of grain boundary segregation of
cause Kr atoms preimplanted into UO2 to nucleate Cr, Ni, and Si in commercial purity 16 stainless steel
into bubbles, and preformed bubbles to undergo res- following irradiation with either protons or neutrons to
olution. A radiation-enhanced diffusion coefficient similar doses. From Was, G. S.; Busby, J. T.; Allen, T.; et al.
for the Kr was estimated from these studies to be J. Nucl. Mater. 2002, 300, 198–216.
D  1.2  1030 cm5  F_ , where F_ is the fission rate
per cubic centimeter, and found independent of tem- 3 MeV proton irradiation at 360  C to similar doses.
perature below 500  C (see Matzke et al.37 for details). Figure 19 compares the RIS behavior of Cr, Ni, and
The importance of such studies as these is that Si in a 316 stainless steel alloy following irradiation to
the basic processes in complex nuclear fuels can be approximately 1 dpa. Neutron irradiation results are
elucidated by studies that carefully control singly in open symbols and proton irradiation results are in
the irradiation conditions and materials parameters solid symbols. This dose range was chosen as an
in the fuel, such as fission gas concentration, damage, etc. extreme test of proton irradiation to capture the
‘W’-shaped chromium depletion profile caused by
irradiation of a microstructure, which contained
1.07.4.3 Comparison with Neutrons
grain boundaries that were enriched with chromium
Proton irradiation has undergone considerable prior to irradiation. Note that the two profiles track
refinement as a radiation damage tool. Numerous each other extremely well, both in magnitude and
experiments have been conducted and compared spatial extent. Good agreement is obtained for all
to equivalent neutron irradiation experiments in three elements.
order to determine whether proton irradiations cap- Figure 20 shows a comparison of the dislocation
ture the effects of neutron irradiation on microstruc- microstructure as measured by the dislocation loop
ture, microchemistry, and hardening. In some cases, size distribution (Figure 20(a)) and the size and
benchmarking exercises were conducted on the same number density of dislocation loops (Figure 20(b))
native alloy heat as neutron irradiation in order to for 304 SS and 316 SS. The main features of the loop
eliminate heat-to-heat variations that may obscure size distributions are similar for the two irradiations,
comparison of the effects of the two types of irradiat- viz. a sharply peaked distribution in the case of 304
ing particles. The following examples cover a number SS and a flatter distribution with a tail in the case of
of irradiation effects on several alloys in an effort to 316 SS. The agreement in loop size is good for the
demonstrate the capability of proton irradiation 304 SS alloy, while loops are smaller for the proton-
to capture the critical effects of neutron irradiation. irradiated 316 alloy. The loop density is about a
Figures 19–23 show direct comparisons of the factor of 3 less for the proton-irradiated case than
same irradiation feature on the same alloy heats the neutron-irradiated case, which is expected as the
(commercial purity (CP) 304 and 316 stainless steels) proton irradiation temperature was optimized to
following either neutron irradiation at 275  C or track RIS (higher temperature) rather than the
212 Radiation Damage Using Ion Beams

Fraction of loop population (%)


50
Fraction of loop population (%)
30 Protons at 360 ⬚C (1.0 dpa)
Protons at 360 ⬚C 40
(1.0 dpa) Neutrons at 275 ⬚C (1.1 dpa)
Neutrons at 275 ⬚C 30
20 (0.7 dpa)
20
10
CP 304 SS 10 CP 316 SS

0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
(a) Loop diameter (nm) Loop diameter (nm)

12 1024

10
Loop diameter (nm)

Loop density (m−3)


8 1023

4 1022
304 316 304 316
2 Protons at 360 ⬚C Protons at 360 ⬚C
Neutrons at 275 ⬚C Neutrons at 275 ⬚C
0 1021
0 1 2 3 4 5 6 0 1 2 3 4 5 6
(b) Dose (dpa) Dose (dpa)

Figure 20 Comparison of (a) loop size distributions and (b) loop diameter and loop number density for commercial
purity 304 and 316 stainless steels irradiated with neutrons or protons to similar doses. From Was, G. S.; Busby, J. T.;
Allen, T.; et al. J. Nucl. Mater. 2002, 300, 198–216.

1500 1500
CP 304 SS CP 316 SS
Yield strength (MPa)

Yield strength (MPa)

1000 1000

500 500

0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
(a) Dose (dpa) (b) Dose (dpa)

Protons at 360 ⬚C (hardness)


Neutrons at 275 ⬚C (hardness)
Neutrons at 275 ⬚C (shear punch)

Figure 21 Comparison of hardening in commercial purity 304 (a) and 316 (b) stainless steel irradiated with neutrons or
protons to similar doses. From Was, G. S.; Busby, J. T.; Allen, T.; et al. J. Nucl. Mater. 2002, 300, 198–216.

dislocation loop microstructure. That the loop sizes much smaller than those from neutron irradiation.
and densities are even close is somewhat remarkable The surviving fraction of interstitial loops, however,
considering that loop density is driven by in-cascade is greater for proton irradiation, partially compensat-
clustering, and cascades from proton irradiation are ing the greater loop formation rate under neutron
Radiation Damage Using Ion Beams 213

Fast neutron fluence (E > 1 MeV) ⫻1025 n m−2 1.4


0 0.5 1 1.5 2 2.5 3 3.5 4
Neutrons
1.2
100 With He
CP 304 SS
Protons at 360 ⬚C Without He
Neutrons at 275 ⬚C 1
NWC
80
Measured IG percentage

0.8

s/s0
60 0.6

0.4
40

0.2
20
0
0 0.5 1 1.5 2 2.5
Dose (dpa)
0
0 1 2 3 4 5 6 Figure 24 Comparison of relaxation in residual stresses
Dose (dpa) between neutron- and proton-irradiated stainless steel after
removing the effect of thermally-induced relaxation. From
Figure 22 Comparison of the extent of intergranular Sencer, B. H.; Was, G. S.; Yuya, H.; Isobe, Y.; Sagasaka, M.;
stress corrosion cracking in commercial purity 304 stainless Garner, F. A. J. Nucl. Mater. 2005, 336, 314–322.
steel following similar stress corrosion cracking tests of
either neutron- or proton-irradiated samples from the same
heat. From Was, G. S.; Busby, J. T.; Allen, T.; et al. J. Nucl.
Mater. 2002, 300, 198–216. are again similar, with proton irradiation resulting in
slightly lower hardness. Figure 22 shows the IASCC
susceptibility of CP 304 SS as measured by the %IG on
the fracture surface following constant load testing
80 0.14
(neutron-irradiated samples) and constant extension
Ni+ irradiation
70 675 ⬚C 0.13 rate testing (proton-irradiated samples) in BWR nor-
Swelling (%) neutron and Ni ion

60
140 dpa
0.12
mal water chemistry (NWC). Despite the significantly
Swelling (%) protons

different testing mode, the results are in excellent


50 Proton irradiation 0.11 agreement in that both proton and neutron irradiation
400 ⬚C
40 3.0 dpa 0.10 result in the onset of IGSCC, at about 1 dpa.40
Figure 23 shows the swelling behavior in austen-
30 Neutron irradiation 0.09 itic stainless steels as a function of nickel content
510 ⬚C
20 2.6 ⫻ 1026 n m−2 0.08 for proton, Ni ion, and neutron irradiation. While
E > 0.1 MeV these experiments were conducted on different sets
10 0.07
of alloys, and under highly disparate irradiation con-
0 0.06 ditions, they all show the same dependence of nickel
0 20 40 60 80 100
Bulk nickel concentration (at.%)
on swelling. In the two commercial purity alloys,
no voids were formed in either neutron or proton-
Figure 23 Effect of bulk nickel concentration on swelling
irradiated samples.
resulting from irradiation with different particles: neutrons,
nickel ions, and protons. Reproduced from Allen, T. R.; As a last example of stainless steel alloys,
Cole, J. I.; Gan, J.; Was, G. S.; Dropek, R.; Al Kenik, E. Figure 24 shows the relaxation of residual stress by
J. Nucl. Mater. 2005, 341, 90–100. neutron and proton irradiation. Here again, results
are from different alloys and different types of
tests, but both show the same dependence of stress
irradiation and resulting in loop densities that are relaxation on dose.
within a factor of 3.39 The next examples are from reactor pressure vessel
Figure 21 shows a comparison of irradiation hard- steel and Zircaloy. Figure 25 shows an experiment on
ening between the two types of irradiation. The results model reactor pressure vessel alloys in which the
214 Radiation Damage Using Ion Beams

500 300
VA, Fe, neutron Tin = 300 ⬚C (all)
VA, Fe, proton −7
Proton: 3–7 ⫻ 10 dpa s–1
−10
400
VA, Fe, electron Neutron: 3 ⫻ 10 dpa s–1 280
VD, Fe–0.9Cu–1.0 Mn, neutron −9
Electron: 7 ⫻ 10 dpa s–1
VD, Fe–0.9Cu–1.0 Mn, proton
Change in yield strength (MPa)

VD, Fe–0.9Cu–1.0 Mn, electron


260

Vickers hardness (Hv)


VH, Fe–0.9Cu, neutron
300
VH, Fe–0.9Cu, proton

240
200
220

100
200
p - Zircaloy 4, 350 ⬚C
n - Zircaloy 2, 350–400 ⬚C
0
180 p - Zircaloy 4, 310 ⬚C

−100 160
10−5 10−4 10−3 10−2 10−1 0 1 2 3 4 5 6 7 8
Dose (dpa) Dose (dpa)
Figure 25 Irradiation hardening in model reactor pressure Figure 26 Hardening of Zircaloy-4 irradiated with 3 MeV
vessel steels following neutron, proton, and electron protons at 310 and 350  C and comparison to neutron-
irradiation at about 300  C. From Was, G. S.; Hash, M.; irradiated Zircaloy-2. From Zu, X. T.; Sun, K.; Atzmon, M.;
Odette, G. R. Philos. Mag. 2005, 85(4–7), 703–722. et al. Philos. Mag. 2005, 85(4–7), 649–659.

same model alloy heats were irradiated with neu- 2.5 MeV electrons, 3.0 MeV protons, and fission neu-
trons, electrons, or protons at 300  C to doses span- trons.41 An attempt was made to keep all irradiation
ning two orders of magnitude. The alloys include a variables constant during the experiments, sample
high-purity Fe heat (VA) that hardens very little purity, defect production rate, and temperature; only
under irradiation, an Fe–0.9Cu (VH) heat that hard- the primary recoil spectrum was varied. The results
ens rapidly initially, followed by a slower hardening for nucleation rates of voids and void swelling are
rate above 0.1 mpda, and a Fe–0.9Ce–1.0Mn alloy shown in Figure 27(a) and 27(b), respectively.
(VD) in which the hardening rate is greatest over Clearly observed is that void swelling and void
the dose range studied. Despite the very different nucleation are significantly enhanced for neutron
compositions and hardening rates, the results of the irradiation in comparison to proton or electron
three types of irradiation agree well. irradiation. This result is notably in strong contrast
Figure 26 shows hardening for Zircaloy-2 and to the efficiencies obtained for defect production and
Zircaloy-4 irradiated with either neutrons or protons. radiation-induced segregation (or FMDs) for these
Although the irradiations were not conducted three types of irradiation. The reduced efficiency of
on the same heats of material, or using similar the production of FMDs was attributed to defect anni-
irradiation parameters, there is good agreement in hilation within the cascade core; these results for void
the magnitude and dose dependence of hardening. swelling, however, indicate that the defect clustering
Proton irradiation also induced amorphization of a process is also critical to microstructural evolution in
Zr(Fe,Cr)2 precipitate after irradiation to 5 dpa irradiated alloys. Singh and coworkers41,42 argue that
at 310  C, similar to that observed in reactor. These the clustering of interstitials in cascades, and their
examples represent a comprehensive collection of collapse into dislocation loops, result in interstitial
comparison data between proton and neutron irradia- migration by one-dimensional glide of loops, the
tion and taken together serve as a good example for the so-called production bias model.43 As a consequence,
capability of charged particles to emulate the effect of interstitials and vacancies become efficiently sepa-
neutron irradiation on the alloy microstructure. rated. Swelling therefore is more severe for irradia-
As a final example, to emphasize the care that tions that produce energetic cascade, for example,
must be exercised in extrapolating the results of one neutrons, than for those that do not, electrons. Proton
type of irradiation to make predictions for another, irradiation is intermediate; that is, small cascades are
we discuss a comparison of void swelling in Cu due to produced.
Radiation Damage Using Ion Beams 215

100
Copper
523 K
10−1
1022
Copper
10−2 523 K
Swelling (%)

1021

Void density (m−3)


Fission neutrons
10−3
20 3 MeV protons
10

10−4
Fission neutrons 1019
3 MeV protons 2.5 MeV electrons
2.5 MeV electrons
10−5 −4 1018 −4
10 10−3 10−2 10−1 100 10 10−3 10−2 10−1 100
Dose (NRT dpa) Dose (NRT dpa)

Figure 27 Void swelling as a function of dose in oxygen-free high conductivity (OFHC)-copper during irradiations with
electrons, protons, and fission neutron. Reproduced from Singh, B. H.; Eldrup, M.; Horsewell, A.; Ehrhart, P.; Dworschak,
F. Philos. Mag. A 2000, 80, 2629.

1.07.5 Advantages and 60


Disadvantages of Irradiations using
ton = 50 ms
Various Particle Types
50 toff = 2000 ms

Each particle type has its advantages and disadvan-


tages for use in the study of radiation effects or for 40
emulating neutron irradiation damage. Common dis-
K0/K0,avg

advantages of charged particle beams are the lack of


30
transmutation reactions and the need to use a raster-
scanned beam. With the exception of some minor
transmutation reactions that can occur with light 20
ion irradiation, charged particles do not reproduce
the types of transmutation reactions that occur in
10
reactor core materials due to the interaction with
K0,avg
neutrons. The most important of these is the produc-
tion of He by transmutation, particularly in alloys 0
0 1 2 3 4 5 6 7
that contain elements such as Ni or B. But a second
Time (ms)
consideration is that of a raster-scanned beam in
which any volume element of the target is exposed Figure 28 The effect of a raster-scanned beam on the
instantaneous production rate of point defects with the
to the beam for only a fraction of the raster-scan
same time averaged rate as a continuous source. From
cycle. For a typical beam scanner and beam para- Was, G. S.; Allen, T. R. In Radiation Effects in Solids, NATO
meters, the fraction of time that any particular vol- Science Series II: Mathematics, Physics and Chemistry;
ume element in the solid is being bombarded is Sickafus, K. E., Kotomin, E. A., Uberuaga, B. P., Eds.;
0.025. Thus, the instantaneous dose rate during Springer: Berlin, 2007; Vol. 235, pp 65–98.
the beam-on portion of the cycle is 40 times that of
the average, Figure 28. The result is that the defect rate in raster-scanned systems will be less, and must
production rate is very high and defects can anneal be accounted for.
out in the remaining 0.975 portion of the cycle One objective of ion irradiation is to emulate
before the beam again passes through the volume the effect of neutrons, and a second is to understand
element. As such, the effective defect production basic physical radiation damage processes, for which
216 Radiation Damage Using Ion Beams

neutron irradiation is often less well suited. While a hot filament or a field emission gun as an electron
ion irradiation can be conducted with great control source. An advantage is that the same instrument
over temperature, dose rate, and total dose, such used for irradiation damage can be used to image
control is a challenge to reactor irradiations. For the damage. Another advantage is that the high dose
example, instrumented tubes with active temperature rate requires very short irradiation time, but will
control are expensive to design, build, and operate. also require a large temperature shift as explained
Even so, frequent power changes can be difficult in the Section 1.07.3.
to handle as the flux–temperature relationship will There are several disadvantages to electron irra-
change and this can result in artifacts in the irradiated diation using a TEM. First, energies are generally
microstructure.44 On the other hand, temperatures in limited to 1 MeV. This energy is sufficient to produce
cheaper irradiation vehicles that use passive gas gaps an isolated FP in transition metals, but not cascades.
and gamma heating (such as ‘rabbit’ tubes) are known The high dose rate requires high temperatures that
with even less certainty. While neutron dosimetry must be closely monitored and controlled, which is
is used in some experiments, doses and dose rates difficult to do precisely in a typical TEM sample
are often determined by neutronic models of the stage. Another drawback is that as irradiations are
core locations and are not verifiable. As such, ion often conducted on thin foils, defects are created in
irradiations enjoy the advantage of better control close proximity to the surface and their behavior may
and verification of irradiation conditions as compared be affected by the presence of the surface. Perhaps
to neutron irradiation. Table 3 provides a list for the most serious drawback is the Gaussian shape to
each of three particle types: electrons, heavy ions, the electron beam that can give rise to strong dose
and light ions (protons), and they are discussed in rate gradients across the irradiated region. Figure 29
detail in the following sections. shows the composition profile of copper around a
grain boundary in Ni–39%Cu following electron
irradiation. Note that while there is local depletion
1.07.5.1 Electrons
at the grain boundary (as expected), the region adja-
Electron irradiation is easily conducted in a high- cent to the minimum is strongly enriched in copper
voltage transmission electron microscope using either because of the strong defect flux out of the irradiated

Table 3 Advantages and disadvantages of irradiations with various particle types

Advantages Disadvantages

Electrons
Relatively ‘simple’ source – TEM Energy limited to 1 MeV
Uses standard TEM sample No cascades
High dose rate – short irradiation times Very high beam current (high dpa rate) leading to large temperature
shifts relative to neutrons
Poor control of sample temperature
Strong ‘Gaussian’ shape (nonuniform intensity profile) to beam
No transmutation
Heavy ions
High dose rate – short irradiation times Very limited depth of penetration
High Tavg Strongly peaked damage profile
Cascade production Very high beam current (high dpa rate) leading to large temperature
shifts relative to neutrons
Potential for composition changes at high dose via implanted ion
No transmutation
Light ions
Accelerated dose rate – moderate irradiation times Minor sample activation
Smaller, widely separated cascade
Modest DT required No transmutation
Good depth of penetration
Flat damage profile over tens of microns

Source: Was, G. S.; Allen, T. R. In Radiation Effects in Solids, NATO Science Series II: Mathematics, Physics and Chemistry;
Sickafus, K. E., Kotomin, E. A., Uberuaga, B. P., Eds.; Springer: Berlin, 2007; Vol. 235, pp 65–98.
Radiation Damage Using Ion Beams 217

8
9
7
8
6 494 ⬚C
D+

Si concentration (at.%)
7
5
Cu concentration (at.%)

e-beam 6
4 diameter
8
6 400 ⬚C
e−
7
5
6
4
5
6
−4 −3 −2 −1 0 1 2 3 4
400 ⬚C
Distance from grain boundary (μm)
5
Figure 31 Comparison of (a) deuteron and (b) electron
4 irradiation showing the greater amount of segregation and
−6 −4 −2 0 2 4 6 the narrower profile for the deuteron irradiation. From
Distance from grain boundary (μm) Wakai, E. Trans. J. Nucl. Mater. 1992, 33(10), 884.

Figure 29 Enrichment of copper surrounding a local


depletion at the grain boundary. The enrichment is caused zone defined by the horizontal line below the spec-
by the high defect flux away from the irradiated region trum. This outward-directed defect flux causes a
defined by the horizontal line. From Ezawa, T.; Wakai,
E. Ultramicroscopy 1991, 39, 187.
reversal in the direction of segregation from that
caused by a defect flux to the sink. Another often
observed artifact in electron irradiation is very
broad grain boundary enrichment and depletion
60
profiles. Figure 30 shows that the enrichment profile
for Ni and the depletion profiles for Fe and Cr in
stainless steel have widths on the order of 75–100 nm,
50 which is much greater than the 5–10 nm widths
Solute concentration (wt%)

Iron observed following neutron irradiation under similar


Chromium conditions and model simulations of radiation-
Nickel induced segregation. A similar effect was observed
40
by Wakai45 using electron and Dþ irradiation of the
same alloy in which the segregation profile was much
30 higher and narrower around the grain boundary in
the deuteron-irradiated sample as compared to the
electron irradiation (Figure 31).
20

1.07.5.2 Heavy Ions


10 Heavy ions enjoy the benefit of high dose rates
0 100 200 300 400 500 600
Distance (nm)
resulting in the accumulation of high doses in short
times. Also, because they are typically produced in
Figure 30 Broad grain boundary enrichment and the energy range of a few MeV, they are very efficient
depletion profiles in Fe–20Cr–25Ni–0.75Nb–0.5Si following
irradiation with electrons at 420  C to 7.2 dpa. From
at producing dense cascades, similar to those pro-
Ashworth, J. A.; Norris, D. I. R.; Jones, I. P. J. Nucl. Mater. duced by neutrons. The disadvantage is that as with
1992, 189, 289. electrons, the high dose rates require large
218 Radiation Damage Using Ion Beams

15
dpa versus depth for

14 MeV nickel ions (dpa per 1016 ions cm−2)


various ions
incident on nickel C

5 MeV carbon ions (dpa per 1016 ions cm−2)


12 1.2

(dpa per 1016 ions cm−2)


8.1 MeV aluminum ions
Ni
5 1.0
9
Al
4 0.8

6 3 0.6

2 0.4
3
1 0.2

Ni
Al
C
0 0 0
0 0.5 1 1.5 2 2.5 3
Depth (μm)

Figure 32 Damage profiles for C, Al, and Ni irradiation of a nickel target at energies selected to result in the same
penetration depth. From Whitley, J. B. Ph.D. Thesis, University of Wisconsin-Madison, Madison, WI, 1978.

250 1.5 1.5


3.2
5 MeV Ni2+ on Ni
Displacement rate (10−3 dpa s–1)

2.8
200
2.4
Observed 1
1
Swelling (%)

2 150

1.6
dpa 100
1.2
0.5 0.5
0.8
50
0.4

0 0 0 0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
(a) Depth (μm) (b) Depth (μm)
þ
Figure 33 (a) Subsurface swelling resulting from 5 MeV Ni ion irradiation of Fe–15Cr–35Ni at 625  C and (b) displacement
rate and ion deposition rate calculated for 5 MeV Ni2þ on nickel. Adapted from Garner, F. A. J. Nucl. Mater. 1983, 117,
177–197; Lee, E. H.; Mansur, L. K.; Yoo, M. H. J. Nucl. Mater. 1979, 85&86, 577–581.

temperature shifts so that irradiations must be con- peaks sharply at only 2 mm below the surface. As a
ducted at temperatures of 500  C in order to create result, regions at a very well-defined depth from the
similar effects as neutron irradiation at 300  C. surface must be isolated and sampled in order to
Clearly, there is not much margin for studying neu- avoid dose or dose rate variation effects from sample
tron irradiations at higher reactor temperature to sample. Small errors (500 nm) made in locating the
as higher ion irradiation temperatures will cause volume to be characterized can result in a dose that
annealing. Another drawback is the short penetration varies by a factor of 2 from the target value.
depth and the continuously varying dose rate over A problem that is rather unique to nickel ion irra-
the penetration depth. Figure 32 shows the damage diation of stainless steel or nickel-base alloys is that in
profile for several heavy ions incident on nickel. addition to the damage they create, each bombarding
Note that the damage rate varies continuously and Ni ion constitutes an interstitial. Figure 33(a) shows
Radiation Damage Using Ion Beams 219

that 5 MeV Ni2þ irradiation of a Fe–15Cr–35Ni alloy Figure 34 shows schematics of 3.2 MeV proton and
resulted in high swelling in the immediate subsurface 5 MeV Ni2þ damage profiles in stainless steel. Super-
region compared to that near the damage peak. As imposed on the depth scale is a grain structure with a
shown in Figure 33(b), the Ni2þ ions come to rest at a grain size of 10 mm. Note that with this grain size,
position just beyond the peak damage range. So even there are numerous grain boundaries and a significant
though the peak damage rate is about 3 that at the irradiated volume over which the proton damage rate
surface, swelling at that location is suppressed by is flat. The dose rate for proton irradiations is 2–3
about a factor of 5 compared to that at the surface.46 orders of magnitude lower than that for electrons or
The reason is that the bombarding Ni2þ ions consti- ions, thus requiring only a modest temperature shift,
tute interstitials and the surplus of interstitials near but as it is still 102–103 times higher than neutron
the damage peak results in a reduction of the void irradiation, modest doses can be achieved in reason-
growth rate.47,48 In the dose rate–temperature regime ably short irradiation time.
where recombination is the dominant point defect The disadvantages are that because of the
loss mechanism, interstitials injected by Ni2þ ion small mass of the proton compared to heavy ions,
bombardment may never recombine as there is no the recoil energy is smaller and the resulting damage
corresponding vacancy production. morphology is characterized by smaller, more widely
spaced cascades than with ions or neutrons. Also,
1.07.5.3 Light Ions as only a few MeV are required to surmount the Cou-
lomb barrier for light ions, there is also a minor amount
In many ways, proton irradiation overcomes the
of sample activation that increases with proton energy.
drawbacks of electron and neutron irradiation.
The penetration depth of protons at a few MeV can
exceed 40 mm and the damage profile is relatively flat
1.07.6 Practical Considerations for
such that the dose rate varies by less than a factor
Radiation Damage Using Ion Beams
of 2 over several tens of micrometers. Further, the
depth of penetration is sufficient to assess such prop-
In the process of setting up an ion irradiation experi-
erties as irradiation hardening through microhardness
ment, a number of parameters that involve beam
measurements, and stress corrosion cracking through
crack initiation tests such as the slow strain rate test.
105

Ion
10−15 H
2+ 1000 He
5 MeV Ni
10−16 Ni
Calculated range (μm)

10−17
dpa/(ion cm−2)

3.2 MeV protons


10
10−18

10−19

10−20 0.1

1 MeV neutrons
10−21 Calculated by SRIM 2000
Stainless steel (Fe–20Cr–10Ni)
10−22 0.001
0 10 20 30 40 0.01 0.1 1 10 100
Depth (μm) Energy (MeV)
Figure 34 Damage profiles for 1 MeV neutrons, 3.2 MeV Figure 35 Range of hydrogen, helium, and nickel ions in
protons, and 5 MeV Ni2þ ions in stainless steel. From Was, stainless steel as a function of ion energy. From Was, G. S.;
G. S.; Allen, T. R. In Radiation Effects in Solids, NATO Allen, T. R. In Radiation Effects in Solids, NATO Science
Science Series II: Mathematics, Physics and Chemistry; Series II: Mathematics, Physics and Chemistry; Sickafus,
Sickafus, K. E., Kotomin, E. A., Uberuaga, B. P., Eds.; K. E., Kotomin, E. A., Uberuaga, B. P., Eds.; Springer: Berlin,
Springer: Berlin, 2007; Vol. 235, pp 65–98. 2007; Vol. 235, pp 65–98.
220 Radiation Damage Using Ion Beams

characteristics (energy, current/dose) and beam- behavior during proton irradiation vary with energy,
target interaction must be considered. ASTM E 521 dose rate, the time to reach 1 dpa, deposited energy,
provides standard practice for neutron radiation and the maximum permissible beam current (which
damage simulation by charged-particle irradiation49 will determine the dose rate and total dose), given a
and ASTM E 693 provides standard practice for temperature limitation of 360  C. With increasing
characterizing neutron exposures in iron and low energy, the dose rate at the surface decreases because
alloy steels in units of dpa.9 One of the most important of the drop in the elastic scattering cross-section
considerations is the depth of penetration. Figure 35 (Figure 36(a)). Consequently, the time to reach a
shows the range versus particle energy for protons, target dose level, and hence the length of an irradia-
helium ions, and nickel ions in stainless steel as tion, increases rapidly (Figure 36(b)). Energy depo-
calculated by SRIM.50 The difference in penetration sition scales linearly with the beam energy, raising
depth between light and heavy ions is over an order the burden of removing the added heat in order
of magnitude in this energy range. Figure 36 shows to control the temperature of the irradiated region
how several other parameters describing the target (Figure 36(c)). The need to remove the heat due
to higher energies will limit the beam current at
a specific target temperature (Figure 36(d)), and a
limit on the beam current (or dose rate) will result in
10−4 a longer irradiation to achieve the specified dose.
Dose rate (dpa s−1)

Figure 37 summarizes how competing features of


10−5 an irradiation vary with beam energy, creating trade-
offs in the beam parameters. For example, while
10−6 greater depth is generally favored in order to increase
(a)
the volume of irradiated material, the higher energy
10−7
600 required leads to lower dose rates near the surface
Time to reach 1 dpa (h)

500 At 360 ⬚C
and higher residual radioactivity. For proton irradia-
At 400 ⬚C tion, the optimum energy range, achieved by balanc-
400
300 ing these factors, lies between 2 and 5 MeV as shown
200 by the shaded region.
100
0
(b) 400
Energy deposited (W)

500 400
300
Residual
activity Energy 350
200
maximum
Residual activity (arbitrary units)

400 Energy
minimum due to
100 residual 300
due to
depth activity Time to reach 1 dpa (h)
0 penetration 250
(c) 300
Range (μm)

100 Range
Beam current (μA)

80 Maximum current at 360 ⬚C 200


Maximum current at 400 ⬚C
60 200
150
40
Time to 1 dpa
100
20 100
0 50
5 10 15 20
(d) Energy (MeV)
0 0
1 2 5 10 20
Figure 36 Behavior of beam-target parameters as a
Energy (MeV)
function of beam energy proton irradiation at 360  C;
(a) dose rate, (b) time to reach 1 dpa, (c) energy deposition, Figure 37 Variation of ion range, residual activity, and
and (d) beam current limit to maintain a sample temperature time to reach 1 dpa as a function of proton energy.
of 360  C. From Was, G. S.; Allen, T. R. In Radiation Reproduced from Was, G. S.; Allen, T. R. In Radiation
Effects in Solids, NATO Science Series II: Mathematics, Effects in Solids, NATO Science Series II: Mathematics,
Physics and Chemistry; Sickafus, K. E., Kotomin, E. A., Physics and Chemistry; Sickafus, K. E., Kotomin, E. A.,
Uberuaga, B. P., Eds.; Springer: Berlin, 2007; Vol. 235, Uberuaga, B. P., Eds.; Springer: Berlin, 2007; Vol. 235,
pp 65–98. pp 65–98.
Radiation Damage Using Ion Beams 221

References 25. Wei, L. C.; Lang, E.; Flynn, C. P.; Averback, R. S. Appl.
Phys. Lett. 1999, 75, 805.
26. Fielitz, P.; Macht, M. P.; Naundorf, V.; Wollenberger, H.
1. Garner, F. A. J. Nucl. Mater. 1983, 117, 177. J. Nucl. Mater. 1997, 251, 123.
2. Mazey, D. J. J. Nucl. Mater. 1990, 174, 196. 27. Okamoto, P. R.; Harkness, S. D.; Laidler, J. J. ANS Trans.
3. Standard Practice for Neutron Irradiation Damage 1973, 16, 70.
Simulation by Charged Particle Irradiation, Designation 28. Okamoto, P. R.; Wiedersich, H. J. Nucl. Mater. 1974, 53,
E521-89, American Standards for Testing and Materials, 336.
Philadelphia, 1989; p D–9. 29. Schmitz, G.; Ewert, J. C.; Harbsmeier, F.; Uhrmacher, M.;
4. Was, G. S.; Andresen, P. L. JOM 1992, 44(4), 8. Haider, F. Phys. Rev. B 2001, 63, 224113.
5. Andresen, P. L.; Ford, F. P.; Murphy, S. M.; Perks, J. M. 30. Krasnochtchekov, P.; Averback, R. S.; Bellon, P. Phys.
In Proceedings of the Fourth International Symposium Rev. B 2005, 72(17), 174102.
on Environmental Degradation of Materials in Nuclear 31. Enrique, R. A.; Bellon, P. Phys. Rev. Lett. 2000,
Power Systems – Water Reactors; National Association 84, 2885.
of Corrosion Engineers: Houston, TX, 1990; pp 1–83. 32. Enrique, R. A.; Nordlund, K.; Averback, R. S.; Bellon, P.
6. Andresen, P. L. In Stress Corrosion Cracking, Materials J. Appl. Phys. 2003, 93, 2917.
Performance and Evaluation; Jones, R. H., Ed.; ASM 33. See e.g., Jung, P.; Schwarz, A.; Sahu, H. K. Nucl. Instrum.
International: Meals Park, OH, 1992; p 181. Meth. A 1985, 234, 331.
7. Kinchin, G. H.; Pease, R. S. Prog. Phys. 1955, 18, 1. 34. Mayr, S. G.; Averback, R. S. Phys. Rev. B 2003, 68,
8. Norgett, M. J.; Robinson, M. T.; Torrens, I. M. Nucl. Eng. 214105.
Des. 1974, 33, 50. 35. Brongersma, M. L.; Snoeks, E.; van Dillen, T.; Dillen, A.
9. ASTM E693-01. Standard Practice for Characterizing J. Appl. Phys. 2000, 88, 59.
Neutron Exposures in Iron and Low Alloy Steels in Terms of 36. Serruys, Y.; Trocellier, P.; Miro, S.; et al. J. Nucl. Mater.
Displacements Per Atom (DPA), E 706(ID); American 2009, 386–388, 967.
Society for Testing and Materials: West Conshohocken, 37. Matzke, Hj.; Lucuta, P. G.; Wiss, T. Nucl. Instrum. Meth. B
PA, 2007. 2000, 166–167, 920.
10. Zinkle, S. J.; Singh, B. N. JNM 1993, 199, 173. 38. Ronchi, C. J. Appl. Phys. 1973, 44, 3573.
11. Kulcinski, G. L.; Brimhall, J. L.; Kissinger, H. E. In 39. Gan, J.; Was, G. S.; Stoller, R. E. J. Nucl. Mater. 2001, 299,
Proceedings of Radiation-Induced Voids in Metals; 53–67.
Corbett, J. W., Ianiello, L. C., Eds.; USAEC Technical 40. Onchi, T.; Dohi, K.; Soneda, N.; Navas, M.;
Information Center: Oak Ridge, TN, 1972; p 453, Castano, M. L. In Proceedings of the 11th International
CONF-710601. Conference on Environmental Degradation of Materials
12. Naundorf, V. J. Nucl. Mater. 1991, 182, 254. in Nuclear Power Systems – Water Reactors;
13. Iwase, A.; Rehn, L. E.; Baldo, P. M.; Funk, L. J. Nucl. American Nuclear Society: La Grange Park, IL, 2003;
Mater. 1996, 238, 224–236. p 1111.
14. Mansur, L. K. J. Nucl. Mater. 1994, 216, 97. 41. Singh, B. H.; Eldrup, M.; Horsewell, A.; Ehrhart, P.;
15. Jung, P.; Chaplin, R. L.; Fenzl, H. J.; Reichelt, K.; Dworschak, F. Philos. Mag. A 2000, 80, 2629.
Wombacher, P. Phys. Rev. B 1973, 8, 553. 42. Golubov, S. I.; Singh, B. N.; Trinkaus, H. Philos. Mag. A
16. Vajda Rev, P. Mod. Phys. 1977, 49, 481. 2001, 81, 2533.
17. King, W. E.; Merkle, K. L.; Meshii, M. Phys. Rev. B 1981, 43. Singh, B. N.; Foreman, A. J. E. Philos. Mag. A 1992, 66,
23, 6319. 975.
18. Gibson, J. B.; Goland, A. N.; Milgram, M.; Vineyard, G. H. 44. Garner, F. A.; Sekimura, N.; Grossbeck, M. L.; et al.
Phys. Rev. 1960, 120, 1229. J. Nucl. Mater. 1993, 205, 206–218.
19. Lucasson, P. In Fundamental Aspects of Radiation 45. Wakai, E. Trans. J. Nucl. Mater. 1992, 33(10), 884.
Damage in Metals; Robibnson, M. T., Young, F. W., Jr., 46. Garner, F. A. J. Nucl. Mater. 1983, 117, 177–197.
Eds.; ERDA Report CONF-751006; 1975, p 42. 47. Lee, E. H.; Mansur, L. K.; Yoo, M. H. J. Nucl. Mater. 1979,
20. Corbett, J. W.; Smith, R. B.; Walker, R. M. Phys. Rev. 85&86, 577–581.
1959, 114, 1452. 48. Brailsford, A. D.; Mansur, L. K. J. Nucl. Mater. 1977, 71,
21. Burger, G.; Isebeck, K.; Volkl, J.; Schilling, W.; Wenzl, H. 110–116.
Zeitschrift Angew. Phys. 1967, 22, 452. 49. ASTM E521-96. Standard Practice for Neutron Radiation
22. Garr, K. R.; Sosin, A. Phys. Rev. 1969, 162, 669. Damage Simulation by Charged-Particle Irradiation;
23. Ehrhart, P. In Landolt –Bornstein New Series, Group III; American Society for Testing and Materials: West
Ullmaier, H., Ed.; Springer: Berlin, 1991; Vol. 25, p 115. Conshohocken, PA, 2009.
24. See e.g., Bacon, D. In Computer Simulations in Materials; 50. Ziegler, J. F.; Biersack, J. P.; Littmark, U. The Stopping
Kirchner, H. O., et al. Eds.; Kluwer: The Netherlands, and Range of Ions in Matter; Pergamon: New York,
1996; p 189. 1996.

You might also like