You are on page 1of 23

30.

Hydrogen Cracking in Specialty Steels

Stephen F. Floreen
Knolls Atomic Power Laboratory
Schenectady, New York

INTRODUCTION
Specialty steels could also be called ultra-high-strength
steels. For present purposes we define these materials as
steels that can be heat treated to yield strengths greater than
1000 MPa. The emphasis on heat treating is a practical one,
since for most service applications only heat treating is used
to produce high-strength parts or structures. While specifying
strengths over 1000 MPa is arbitrary, it helps focus on certain
types of steel and strength levels where hydrogen embrittlement
is always a problem.
Table I gives the nominal compositions of some of the
commercial steels of interest. Three different varieties of
steel are presented and it helps future discussion to briefly
describe their hardening behavior. The medium-carbon low-alloy
steels are generally older alloys characterized by carbon
contents of about 0.3-0.5 percent and by relatively low alloy
contents. Because these steels have limited hardenability they
commonly are solution treated at temperatures of about 800-
900°C, and then quenched to form a hard, brittle martensite.
The martensitic structure then is tempered for several hours at
temperatures ranging from 200 to 600°C, depending upon the
desired strength level. Tempering relieves transformation
stresses, eliminates retained austenite and produces a disper-
sion of epsilon carbide particles that are the major source of
strengthening. With higher tempering temperatures, the carbides
coarsen and eventually are replaced by cementite. Alloying
elements alter the tempering behavior, such as in 30OM, where a
Si addition to the chemistry of a 4340 steel is used to retard
cementite formation and to allow higher tempering temperatures.
The medium-carbon low-alloy steels are fairly inexpensive,
but for more critical applications they are often deficient in
fabricability, ductility, toughness and weldability. To miti-
gate these shortcomings, various grades of air-hardening steels
were developed. These alloys have lower carbon contents and
higher concentrations of alloying elements. The higher alloy
content improves the hardenability so that quenching is not
required. The higher Cr and Mo levels also promote the forma-
tion of alloy carbides during tempering and provide secondary
TABLE I
NOMINAL COMPOSITIONS OF SPECIALTY STEELS
Alloy
Designation C Mn Si Cr NJ. Mo Other
Medium Carbon Low Alloy Steels
4130 0.30 0.5 0.15 1.0 - 0.2
4340 0.40 0.7 0.25 0.8 1.8 0.25
30OM 0.45 0.8 1.60 0.8 1.8 0.35 .07V
D-6a 0.45 0.7 0.25 1.0 0.5 1.0 .07V
Air Hardening Steels
H-Il 0.40 0.3 0.9 5.0 - 1.3 0.5V
HP 9-4-20 0.20 0.3 0.1 0.75 9.0 1.0 O.IV,4.5Co
HP 9-4-30 0.30 0.3 0.1 1.0 7.5 1.0 O.IV,4.5Co
AF-1410 0.14 0.8 - 2.0 10.0 1.0 14.0Co
Maraging Steels
18 Ni (200) .03 max - - - 18,0 3.3 8.5Co,0.2Ti
18 Ni (250) .03 max - - - 18.0 5.0 8.5Co,O.4Ti
18 Ni (300) .03 max - - - 18.0 5.0 9.OCo,O.7Ti

hardening at tempering temperatures of about 500°C.


The maraging steels are distinctly different materials
because they do not rely on carbon for hardening. These steels
are austenitized at 820 C, air cooled to form martensite, and
then hardened by heat treating several hours at 480°C. Harden-
ing is produced by the precipitation of Ni3Ti and Ni3Mo inter-
metallic compounds in the low-carbon martensite. The absence
of carbon in these alloys minimizes the lattice distortions
during quenching and tempering, which is of considerable help in
fabricating parts with complex shapes. Also in contrast to
carbon-hardened steels, the as-formed martensite is relatively
soft and ductile; this improves the fabricability and weldabil-
ity. The higher alloy contents of the air-hardening and marag-
ing steels makes them more expensive, but their improved proper-
ties have led to extensive usage in demanding applications.
Several other specialized steel products such as TRIP
steels, high-strength wires, and maraging steel composites are
also discussed. These are interesting materials because of what
they teach us about methods for improving resistance to hydrogen
embrittlement, but they have found only limited or no practical
applications. We also briefly consider weldments, but do not
attempt to treat welding processes or filler metal compositions
in detail.
Several hundred papers have been written about hydrogen
cracking in ultra-high-strength steels. Much of this work has
been summarized in earlier reviews (1-10) and is not
repeated here. Also, no attempt is made to catalog the proper-
ties of the alloys. The aim instead is to focus on more recent
studies of the factors that influence hydrogen embrittlement
(HE).
PROPERTIES IN GASEOUS ENVIRONMENTS
Crack growth in hydrogen or gaseous hydrogen compounds has
been extensively studied in recent years. Part of this activity
arises from service failures, such as in an electrical generator
cooled by hydrogen gas. There has also been extensive interest
in determining the chemical or mechanical reactions controlling
the crack growth kinetics in these environments. Various
gaseous environments also have been used as convenient, reason-
ably well defined test media for examining metallurgical para-
meters. In this section we discuss only the effects of environ-
mental variables such as gas chemistry, temperature, pressure,
etc., and leave consideration of metallurgical factors until
later.
Figure 30-IA shows an example of the crack growth behavior
of a 4340 steel in dry hydrogen gas. There is commonly a thres-
hold stress intensity below which the cracking rate is zero or
extremely small. Above the threshold, in Stage I, the rates
rise rapidly with stress intensity. At higher stress intensi-
ties, in Stage II, the growth remains approximately constant
with increasing stress intensity. Finally, not shown in Figure
30-1, there is usually an accelerated rate in Stage III as the
stress intensity approaches Kj^. Note the temperature depen-
dence of the cracking rate in Stage II, and that the maximum
rate occurred at an intermediate temperature of 75°C. Crack
paths are frequently intergranular in ultra-high-strength
steels, particularly at lower stress intensities.

Stress Intensity Factor (ksi-in1/2)


Temperature (°C)

AISl 4340 STEEU


P N j- 133 UPo UOOO lorr)
I 90pc(-ond 95pel
Stage Il Crack Growth Rate (in/s)

confidence intervals
Stage Il Crack Growth Rate (m/s)
Crack Growth Rate (m/s)

Crack Growth Rate (in/s)

Stress Intensity Factor (MPa-m1/z)

Figure 30-IA: (Left) Crack growth rate data for 4340 steel in
dry H2 at 133 kPa (Ref. 21).
Figure 30-1B: (Right) Stage-II data from Figure IA plotted
versus reciprocal temperature (Ref. 21).
Earlier studies demonstrated that crack growth could take
place in hydrogen gas at atmospheric or sub-atmospheric pressures
(11,12). These results were of considerable interest because
they evidently refuted the pressure theory of hydrogen embrittle-
ment, since high pressure-deposits of molecular hydrogen gas
inside the steel should not exist in thermodynamic equilibrium
with low-pressure external hydrogen. Numerous subsequent experi-
ments confirmed that crack growth in high-strength steels can
occur at quite low hydrogen gas pressures, particularly if care
is taken to remove impurities from the gas system that interfere
with hydrogen absorption. Most of the current interpretations
of hydrogen cracking in ultra-high-strength steels usually are
based on Troiano-type mechanisms of localized weakening of
atomic bonds by hydrogen. Cracking due to internal hydrogen gas
sometimes can be seen in blisters, fish-eyes, etc., where
extremely high local concentrations of hydrogen can be achieved,
but this mechanism does not appear to be a common means of
hydrogen embrittlement in these steels.
Gas composition markedly affects the cracking behavior.
Nelson, et al. (13) showed that atomic hydrogen produced crack
growth rates several orders of magnitude faster than molecular
hydrogen. Various gaseous hydrogen compounds such as CH4, HCl,
HBr, H2S, HF and H 2 O also can cause crack growth, but not NH3
(14-19). Of these various compounds the most potent is H2S,
which can produce crack growth rates in excess of one mm/s (20).
Minor additions of a second gaseous species can markedly
accelerate or inhibit crack growth in hydrogen. Small amounts
of oxygen, in particular, halt crack growth in hydrogen (11), as
can SO2, CO, and NH3 (17). Conversely, additions of H2S
considerably reduce the threshold stress intensity and increased
the crack growth rate by 1000 fold (15). These potent effects
have been attributed to the additives either inhibiting or pro-
moting the supply of hydrogen atoms at the metal surface.
Oxygen can react rapidly and preferentially with steel to form
an oxide layer that inhibits the uptake of hydrogen (17-21).
The accelerating effect of H 2 S appears to be related to the
formation of a surface compound of sulfur that catalyzes the
absorption of hydrogen (17,20).
Attempts have been made to relate crack growth rates in
hydrogen-containing gases to the thermodynamic activities of the
hydrogen. However these predictions do not work well in gas
mixtures where strong inhibiting or catalytic effects are
present, such as from oxygen or sulfur, because of the difficul-
ties in estimating hydrogen activities in these circumstances.
In water vapor environments the crack growth rates in H-Il
steel increased with relative humidity values up to about 40
percent, but at higher pressures the results were identical to
those in liquid water (22). The saturation behavior at higher
humidity was attributed to capillary condensation of water vapor
at the tip of the crack.
Wei and co-workers have conducted a series of experiments
to identify the chemical reactions controlling the crack growth
rates in gaseous environments (20,21,23-26). Figure 30-1B shows
Stage-II (rate independent of K) data from Figure 30-1A plotted
versus reciprocal temperature. At lower temperatures the data
follow an Arrhenius- type relationship with an apparent activa-
tion energy of 14.7 kJ/mole. Similar activation energies have
been found with other steels in dry hydrogen. In water vapor
or in water the values are about 33-38 kJ/mole. For H2S
however, the activation energy was only 4.6 kJ/mole. Surface
chemistry measurements were also made to study the kinetics of
the surface reactions between the steels and the gases. The
results of these studies suggested that the surface reaction
between the hydrogen gas and the steel is the rate-limiting step
in the crack growth kinetics in pure hydrogen. In water vapor
the surface reaction associated with oxidation of the steel and
the associated liberation of hydrogen is rate controlling, with
the resulting surface oxide inhibiting the reaction. In H2S, on
the other hand, the surface reaction is extremely rapid and the
rate-limiting step is the transport of H2S gas to the crack tip.

The threshold stress intensity for crack growth depends


upon gas pressure. Figure 30-2 shows the results of Oriani and
Josephic for 4340 steel in high-purity hydrogen and in deuterium
(27). Note the low threshold values even at very low gas pres-
sures. The loading response of the steel in these tests was
extremely rapid; moving cracks changed velocity within one
second after a load change. Subsequent studies (28) showed that

K. ksi >/Jnch
Figure 30-2: Threshold stress intensities for crack growth of
4340 steel versus gas pressure in hydrogen and deuterium (Ref.
27).

after unloading a considerable boost in pressure over the


equilibrium value was required to restart a stationary crack.
This was attributed primarily to blunting of the crack tip by
the original loading. Cracking was found to be accompanied by
sonic emissions. Oriani and Josephic argued that the sporadic
nature of crack advance was not due to the intrinsic action of
hydrogen, but to plastic tearing of local ligaments that were
relatively unaffected by hydrogen.
Many studies of hydrogen cracking in a variety of environ-
ments have found evidence that crack growth is discontinuous.
Akhurst and Baker (29) observed that crack growth in hydrogen
gas at stress intensities slightly above the threshold was
intermittent and that a crack would decelerate and stop unless a
small load boost was applied. The discontinuous mode of crack
advance was associated with the need to furnish a sufficient
supply of hydrogen to inhomogeneities in the steel ahead of the
crack tip. Many other investigators also have modeled hydrogen
cracking in terms of achieving a critical hydrogen content at
some point ahead of the crack tip. In some cases this distance
is specified in metallurgical parameters such as one grain
diameter. In others it is not specifically defined or left as
an adjustable parameter. These theories commonly assume a
Troiano-type embrittlement, but the actual crack nucleation site
is usually uncertain. Trap theory, as discussed below, tries to
deal with this question. There is also a basic difference of
opinion in the treatments as to whether discontinuous cracking
is due to periodic bursts of hydrogen cracking, or due to
continuous hydrogen cracking and periodic plastic tearing of
ligaments.

The major factor controlling hydrogen cracking is the yield


strength. Figure 30-3 shows the threshold stress intensities
for crack growth versus yield strength for 4340 steel in H2S
(30) and 4130 steel in H 2 (31). Similar threshold-yield
strength trade-offs can be seen in hydrogen-precharged specimens
or in stress corrosion tests. Maraging steels have somewhat
better crack growth resistance in hydrogen than quenched-and-
tempered steels (24-26). However, all steels with yield
strengths above 1000 MPa must be considered susceptible to
hydrogen cracking.

• 50ptlg (344kPa)HeS, Bolt Load Test


o 5pslg(H^Pa)H2S. Rising Load
Test

Yield Strength , ksl

Figure 30-3: Threshold stress intensities for crack growth


versus yield strength. Left - 4340 steel in 344 kPa H 2 S
Left - 4340 steel in 344 kPa H2S (Ref. 30).
Right - 4130 steel in 77 kPa H2 (Ref. 31).
PROPERTIES OF HYDROGEN CHARGED STEELS
Hydrogen can be picked up during steelmaking from many
sources; moist air, damp charge materials, new refractory
linings, hydrocarbon fuels, etc. Pickling and electroplating
also can charge hydrogen into steels. Welding is another opera-
tion where hydrogen can be introduced from damp fluxes, hydro-
carbon preheating gases, and rust, grease or paint on the steels.
The consequent embrittlement and cracking from internal hydrogen
has been recognized for over 100 years and has been extensively
studied. Most sources of hydrogen are now well recognized, and
means of reducing hydrogen absorption by vacuum melting or
degassing, baking, use of special welding materials, etc., have
become established.
These precautions are necessary, because the amounts of
hydrogen needed to cause cracking under the static loading in
ultra-high-strength steels are very small. Figure 30-4 shows
the threshold stress intensities for crack growth versus hydro-
gen content for several steels that were hydrogen charged,
electroplated to retain the hydrogen, and then given a low-
temperature heat treatment to produce a uniform distribution of
hydrogen (32). The sensitivity of the maraging steels to hydro-
gen was less than the 30OM steel, but clearly all the alloys
were significantly affected by several ppm of hydrogen.

*Heat A
Critical Stress Intensity (MNm'2 \/m)

Critical Stress Intensity (ksi \/iri)

*Heat B
=Heat C
=Heat D

Parts Per Million Hydrogen

Figure 30-4: Threshold stress intensities for crack growth in


hydrogen precharged steels. Heat A = 2010 MPa yield strength
maraging steel; Heat B = 1870 MPa yield strength maraging
steel; Heat C = 1740 MPa yield strength maraging steel; Heat
D = 1700 MPa yield strength 300 M steel (Ref. 32).
In ultra-high-strength steel, absorbed hydrogen has little
effect on the yield strength. Tensile ductility is reduced, and
the effects are generally greater with higher yield strengths
and higher hydrogen concentrations. Hydrogen at large fugacity
also promotes plastic instability, as observed in the slip
behavior around U-notched bend specimens of 4340 (33). Sub-
critical crack growth due to hydrogen can cause deviations from
linearity in fracture toughness testing that give apparent
decreases in KJC values (34). Faster loading rate minimized
these effects. Clark and Landes (35) also found that slower
loading rates reduced the fracture toughness of 4340 in hydrogen
gas.
It has been recognized for some time that the test tempera-
ture and strain rate significantly influence the ductility of
hydrogen charged tensile samples. Originally, these effects
were interpreted in terms of hydrogen diffusion during loading,
but considerable evidence now shows that hydrogen can be trans-
ported by moving dislocations. Dislocation transport can
greatly magnify hydrogen mobility and produce hydrogen enrich-
ment at sites well beyond diffusion distances. Whether disloca-
tion transport can create very high concentrations of hydrogen
at local sites has been a point of contention. Nair et al. (36)
indicate that if the sites are adequate traps for hydrogen the
enrichment can be quite significant but this is disputed by
Hirth and Johnson (37). If Nair et al. are correct, certain
traps may nucleate hydrogen cracking very easily because rapid
enrichment of hydrogen can take place at these sites.

STRESS CORROSION CRACKING


Is stress corrosion cracking in aqueous environments the
result of hydrogen embrittlement? This also has been a point of
contention for many alloy systems. It should be noted that
cracking caused by anodic dissolution and as a result of hydro-
gen embrittlement are not necessarily separate and distinct
phenomena (38,39). Both processes can occur simultaneously, and
may operate in parallel or sequentially to promote cracking.
In ultra-high-strength steels, most evidence suggests that
cracking in aqueous environments is a result of HE. Some
earlier experiments showed that anodic polarization promotes
cracking. Later it was recognized that HE can take place over a
broad potential range because the localized dissolution of iron
and subsequent hydrolysis in pits or crack tips produce low pH
values that promote hydrogen generation and absorption (40).
Cathodic polarization has been found often to be very delete-
rious, and cathodic protection of even intermediate-strength
steels should be used with caution.
Stress corrosion tests under different states of stress
have been used to distinguish between anodic dissolution and
HE mechanisms. Triaxial tension is assumed to be the most
severe stress state for hydrogen embrittlement because, as dis-
cussed below, the volume expansion ahead of the crack tip leads
to a high local concentration of hydrogen that causes failure.
In agreement with this hypothesis, Mode-I loading was the most
severe condition but Mode-III loading did not produce cracking
in maraging and 4340 steels (41,42).

More evidence of HE during stress corrosion is afforded by


comparisons of threshold values in different hydrogen-containing
environments. Figure 30-5 shows the data of Sandoz (43) for
CODE
°K«
•KH
*s
* K «CC
OK000(Mg)
THRESHOLD STRESS INTENSITY (ksi /Sj

^K000(Cu)
OK
W«CHAJ«)
HYDROGEN LEVEL IS ABOUT I PPM
APPROXIMATE STRENGTH WHERE

YCLD STRENGTH (ksi)

Figure 50-5: Threshold stress intensities for crack growth


for 4340 steel in H 2 gas, after precharging with hydrogen, or
in stress corrosion (Ref. 43).

4340 specimens tested in a 3% percent NaCl solution, in hydrogen


gas, or after cathodically precharging with hydrogen. The
threshold values are somewhat different because of the differ-
ences in hydrogen fugacity and hydrogen absorption in the
various environments, but the overall pattern' is clearly consis-
tent. The crack paths in all cases were intergranular. A more
recent comparison of four steels in either a saline solution or
in one atmosphere hydrogen gas also showed higher threshold
values in the gaseous environments, but the rankings of the four
alloys were the same in both environments (44).

The influence of environmental factors in the stress corro-


sion cracking in medium-carbon low-alloy steels have been
reviewed by Carter and Hyatt (4) and in maraging steels by
Dautovich and Floreen (5). The chloride content generally has
little effect. The presence of sulfur-containing compounds
however, can promote rapid fracture. Temperatures ranging from
O to 7O0C have little effect. Apparent activation energies for
crack growth comparable to those for hydrogen diffusion have
been observed, but this is not always true. High pH values have
been observed to help prevent crack growth. Applied potential
can have a significant effect. Large cathodic potentials are
deleterious. In many cases significant improvements in crack
growth resistance have been reported at various slightly anodic
or cathodic potentials, but the optimum conditions appear to
depend strongly on the type of steel, specimen design, etc., so
it is difficult to make any general recommendations.
CORROSION FATIGUE
Hydrogen from gaseous or aqueous media can accelerate
fatigue cracking in ultra-high-strength steels. Several recent
reviews (45-48) have treated various aspects of this problem.
Perhaps the most noticeable environmental effects are seen in
comparisons of smooth-bar data in air and aqueous environments.
Because corrosion reactions such as pitting, anodic dissolution,
etc., can take place in aqueous environments, fatigue crack
nucleation and subsequent growth may take place at stress ranges
well below those found in non-aggressive environments. In
gaseous environments such damaging surface reactions rarely
occur, crack nucleation usually is more difficult, and the data
in the two environments are in closer agreement.
With notched or precracked samples it is useful to distin-
guish between environmental effects in different regions of the
da/dN curve. Figure 30-6, taken from Ritchie (49), shows these
three separate regimes. In regime A, at low stress intensity
ranges, the slower crack growth rates allow more time for hydro-
gen absorption and transport and thus the environment may have a
substantial effect on the threshold value and crack growth rate
(48,49). In this region higher R ratios, because the mean
stress is higher, also promote environmental damage. In 30OM,
for example, increasing the R ratio from 0.05 to 0.70 reduced
the threshold value by a factor of two and increased the crack
growth rate about two orders of magnitude at low AK values (49).
Crack closure effects, particularly in environments that form
extensive corrosion products, may also be significant at lower
stress intensities and higher R ratios also will minimize these
effects.

K
Regime A Regime B C
final
NON-CONTINUUM CONTINUUM MECHANISM failure
MECHANISMS (striation growth)
large influence of: little influence of.
i microstructure i microstructure
ii mean stress ii mean stress
iii dilute environment
iv thickness
da/dN , mm/cycle

Regime C
'STATIC MODE* MECHANISMS
(cleavage, inttrgranular and
fibrous)
targt influence of:
i microstructure
ii mean stress
iii thickness
little influence of:
iv environment

threshold AK 0
log AK

Figure 30-6: Illustration of three regimes of corrosion


fatigue (Ref. 49).
At higher stress intensities in regime B slower frequencies
considerably promote hydrogen cracking, again because increased
time at load allows HE effects to take place. Much of the crack
growth rate data for steels in both gaseous and aqueous environ-
ments generally indicates that in regime B an order of magnitude
slower frequency produces about an order of magnitude faster
crack growth rate. This one-to-one relationship is, of course,
what would be expected if time, rather than cycles, is the para-
meter controlling crack growth. At very slow frequencies the
environmental effects saturate, presumably because the time
necessary for the time-dependent processes to occur has been
provided, and still slower frequencies have no additional effect.
The cyclic waveform is also important. Since the earlier work
of Barsom (50), the common view has been that environmental
damage occurs primarily during rising loads, and thus slow load-
ing-fast unloading cycles are most damaging. Some recent work
suggests however that this may not always be true (51).
Corrosion fatigue tests have been conducted in a variety of
gaseous and aqueous hydrogen-containing environments. In
general, environmental effects in fatigue parallel those under
static loading, and as a first approximation the relative
severity of different environments under the two types of load-
ing appear comparable. Inhibiting effects of mixed gases on the
crack growth in fatigue have been observed that are similar to
those discussed earlier in static tests (52). Work by Wei and
co-workers also suggests that the rate-controlling reactions in
hydrogen or water vapor environments remain the same under both
types of loading (45,46). However, in most cases it does not
appear that the crack growth rate is simply the sum of pure
fatigue plus stress corrosion cracking. More generally, the
crack growth in corrosion fatigue (da/dN)^ should be expressed
(46):
(da/dN)E = (da/dN)I + (da/dN)SCC + (da/dN)CF (1)
where the subscript I refers to the crack growth in an inert
environment, SCC is the stress corrosion contribution, and CF
accounts for the interaction between cyclic loading and the
environment. Wei (46) has suggested that this third term is
proportional to the amount of hydrogen produced per cycle at the
crack tip. Other interactions that accelerate or possibly
retard cracking may also be possible.
In regime C, at high AK ranges, the growth rate accelerates
as KIC is approached. Environmental effects are usually small
in this region, but if the frequency is low enough some hydro-
gen-induced cracking may be observed.

METALLURGICAL EFFECTS
A great many studies have been made of the effects of
composition and microstructure on the HE behavior of steels. In
all steels the dominant factor is the yield strength. At lower
yield strengths around 1000 MPa, composition and microstructure
can be fairly important, and better cracking resistance has been
achieved through control of these variables. At higher
strengths however, the crack growth resistance falls off
sharply, and at yield strengths on the order of 1500 MPa all
steels possess comparably poor resistance to hydrogen cracking.

Carter and Hyatt (4) have summarized the results of a


number of investigations of alloy chemistry in medium-carbon
quench-and-tempered steels on the cracking behavior. At yield
strengths of 1000-1300 MPa there is fairly general agreement
that high levels of carbon and manganese have deleterious
effects. Variations in Cr, Ni, Mo, Co have either no effects or
give conflicting results in different tests. However, a high
total level of alloying elements, such as in maraging steels,
could be beneficial because the diffusion rate of hydrogen would
be reduced. Comparisons of the Stage-II crack growth rate data
in stress corrosion do, in fact, show that at comparable yield
strengths the maraging steels display order-of-magnitude lower
crack velocities compared to quench-and-tempered steels (5).
Work comparing the stress corrosion behavior of 30OM and 4340
steel indicates that silicon also lowers the crack growth rate
because it lowers the diffusion rate of hydrogen (53).
Impurity elements are very important. In maraging steels,
C is a deleterious impurity that lowers the resistance to stress
corrosion cracking (5). Sulfur and phosphorous often have been
observed to be detrimental in many grades of steel. Several
explanations for these deleterious effect have been put forward.
At higher levels of impurities the resultant inclusion particles
can act as crack nucleation sites. A recent study of the stress
corrosion behavior of a C-Mn steel shows very clearly the strong
relationship between hydrogen cracking and the elongated MnS
inclusion particles in the steel (54). It has also been pointed
out a number of times that elements such as P or S, either in
the steel or dissolved from the steel into aqueous environments,
can promote hydrogen absorption because they act as poisons to
the hydrogen recombination reaction at the steel surface.
A newer approach has called attention to the relationship
between temper embrittlement and hydrogen embrittlement, and has
suggested a common role in both types of cracking due to segre-
gation of elements such as P and S to the grain boundaries.
Earlier studies had shown that different tempering treatments
could sometimes produce surprisingly large changes in HE. For
example, in hydrogen-precharged 4340 samples, Gerberich, et al.
(55) noted that the Stage-II crack growth rates after different
tempering treatments showed much bigger variations than would be
expected simply from the strength changes. Subsequent work has
demonstrated that segregation of elements such as P or S to the
grain boundaries occurs during tempering, and that this segre-
gation markedly impairs the resistance to hydrogen embrittlement
(56-64).
This degradation in resistance to hydrogen often parallels
the deterioration in mechanical properties produced by temper
embrittlement. Increasing tempering time markedly lowers the
threshold stress intensity for cracking in hydrogen (56). In
addition to S and P, other impurities usually associated with
temper embrittlement such as Sb, Sn, etc., also are deleterious
(63). Additions of Mo to 4130 helped minimize hydrogen cracking
(61), but Mn and Si were found to be deleterious because they
promoted P segregation to the grain boundaries (58,63). The
effect of Ni is controversial, with current work suggesting that
deleterious effects are due to microstructural factors and not
Ni per se (64).
With suitable control of the Si, Mn and P levels the thres-
hold stress intensity for crack growth in hydrogen gas has been
dramatically increased, as shown in Figure 30-7 for a 5% Ni
steel from the results of Takeda and McMahon (63). It should be
APPARENT K f h MPoVTW
HEAT A

HEAT F

AGEING TIME (h) AT 480'C

Figure 50-7: Threshold stress intensities for crack growth in


hydrogen gas versus tempering time. Heat A = low P, Si and Mn
heat of HY-130 steel. Heat F = commercial heat of HY-130
steel (Ref. 63).

noted however, that these results were obtained at yield


strengths of approximately 1000 MPa.
McMahon, et al. (56) have suggested that as a first approxi-
mation the fracture strength Of after embrittlement may be given
as

°f = 0O - Aa imp - Aa H (2)

where a0 is the strength of the pure steel without impurities or


hydrogen, and Aai m p and AQH represent the losses in strength due
to impurity segregation alone and hydrogen alone. As these
investigators point out, this simple additivity is the minimum
effect that might be expected, and that synergistic interactions
between the impurities and hydrogen might promote further
damage. Also, microstructural changes during tempering may
influence Aa-[mp and Aa^ in different ways so that the combined
effects may not be simply additive (59).
Impurity effects are most noticeable at lower strengths.
At high yield strengths it is obviously necessary to keep
residuals below certain limits, but the evidence does not
suggest that extreme high purity will provide substantial
benefits. In their crack growth studies in hydrogen gas and
water vapor Wei, et al. (21) used a 4340 steel containing only
9 ppm P and 12 ppm S. No traces of either element, or any other
obvious impurity, were seen in Auger examinations of the frac-
ture surfaces, yet these samples cracked quite readily at stress
intensities of about 25 MPa/m. In maraging steels, earlier work
suggested a marked improvement in stress corrosion resistance if
the S levels were reduced from 150 to 30 ppm (5). Additional
studies on heats of 18 Ni (250) maraging steel, however, showed
that there was no further benefit in stress corrosion in heats
containing from 1 to 30 ppm S (65). At present, aside from the
steep costs of making extremely high-purity steels on a commer-
cial scale, there is no clear evidence as to what element or
combination of impurity elements must be eliminated, or what
improvements, if any, could be realized by such an approach.

One other very impractical but interesting alloy chemistry


effect is the use of noble metal additions to reduce hydrogen
damage. Wilde, et al. (66) showed that additions of about one
weight percent of either Pt or Pd to a 4130 steel tempered to
690 MPa yield strength considerably improved the threshold
stress intensities in a f^S-saturated NaCl solution. These addi-
tions did not alter the hydrogen absorption behavior. The
mechanism by which Pt improved the cracking resistance was not
clear. The MnS inclusions however, were found to be coated with
Pd, and it was speculated that the Pd inhibited the formation of
hydrogen gas and subsequent nucleation of fractures at these
sites. In principle one might achieve such beneficial segrega-
tion effects with less expensive alloying additions, but thus
far little work of this kind has been reported.

Microstructure can also pay a role, but the effects are not
always clear cut. It is agreed that twinned martensite is much
more susceptible to cracking than lath martensite (4,67).
Tempered martensite or bainite usually are considered the best
microstructures, and untempered martensite the worst (67).
Judging from the rather comparable KJSQQ values of the carbide-
hardened steels and the maraging steels, it would seem that
there is little effect of carbides versus intermetallic com-
pounds as precipitate particles. Grain size is difficult to
categorize, since different studies have found that finer grain
size is beneficial, deleterious, or has no effect. Gerberich
and Wright (68) have reviewed much of the earlier work and con-
cluded that the effects of grain size depend upon the alloy
purity. In high-purity heats increasing the grain size leads to
greater impurity segregation in the grain boundaries and thus is
deleterious. In "dirty" heats the grain boundaries are saturat-
ed with impurities at all grain sizes. Coarser grain sizes then
are beneficial because the micro-fracture zone size is larger
and suitably oriented grains will be able to retard crack growth.
Other microstructural processing effects are even less
certain. Ausforming has been examined several times with
mixed results. Various attempts also have been made to improve
the hydrogen resistance by introducing either untransformed
or reverted austenite into the microstructure. Austenite could
be beneficial for several reasons, e.g., slowing hydrogen diffu-
sion, absorbing harmful impurities, relaxing local stresses,
etc., and some studies have indicated that austenite improves
the stress corrosion resistance. However, the results have been
difficult to reproduce. In maraging steels, for example, heat
treatments intended to produce reverted austenite gave widely
different results in different heats of the same alloy because
minor variations in processing history, segregation, etc.,
drastically altered the austenite morphology (65).
More generally, it should be remarked that the stress
corrosion data for ultra-high-strength steels often show
considerable scatter. Part of this scatter is no doubt due to
variations in composition and process history. Scatter of this
kind may depend upon when the steel was made, since newer pro-
cessing and inspection techniques and better understanding of
the metallurgical effects should at least tend toward eliminat-
ing the production of bad heats. There is also some sense of
progress at lower-strength levels in that new steels, such as
AF 1410, may provide higher KJ^QQ values. At yield strengths of
about 1500 MPa however, no visible improvement in KJCJQQ has been
evident in many years.
Is it possible, assuming optimum control of composition,
processing steps, etc., to achieve significant improvements in
KJSQQ at these very high strength levels? One new way to
examine this question is through the use of trap theory (69-71).
Traps in steels were first noticed when permeability experiments
showed that significant amounts of hydrogen are retained at
certain sites within the lattice. Since that time various
experiments have shown that hydrogen can be preferentially
absorbed at traps such as grain boundaries, particle boundaries,
dislocations and solute atoms. In particular, the use of
tritium autoradiography has directly demonstrated hydrogen trap-
ping often occurs associated with C or C-containing particles
(72).
A fairly detailed methodology (69,71) been developed to
predict the redistribution of hydrogen atoms during straining.
Some sites, depending upon temperature and strain rate, are
shallow, reversible traps that will surrender hydrogen atoms to
dislocations. Other sites are irreversible traps that retain
hydrogen. Dislocations provide transport of hydrogen atoms
between the traps. Fracture begins when the combination of
local stress and hydrogen content at irreversible traps is
sufficient to nucleate cracking. Several interesting conclu-
sions arise from this analysis. One is the distinction between
exterior and interior sources of hydrogen. When hydrogen is
being supplied from an exterior source during loading it is
desirable to have a very high density of both reversible and
irreversible traps, because more hydrogen will then be drawn
away from crack nucleation sites into innocuous trap sites.
However, when hydrogen is initially within the steel, such as in
precharged material, reversible traps will be bad because more
hydrogen will be made available for cracking. Thus the differ-
ent forms of hydrogen supply require different alloy design.
The best overall microstructure is to have no reversible
traps and a high, uniform density of small, incoherent particles
that will act as irreversible traps (69,71). By providing a
large number of irreversible traps the hydrogen build-up at any
individual trap would thus be minimized.
Such a microstructure also should be effective in homo-
genizing slip, and thus reducing local stress concentrations at
the particles. Indeed, if one were to design a high-strength
steel with an optimum combination of strength and toughness the
desired particle morphology would be essentially the same. Thus
a steel optimized for hydrogen cracking resistance should have
excellent toughness. The reverse is not necessarily true. For
example, a steel could have good fracture toughness but a low
KISCC because of segregation of an impurity to the grain
boundaries. Many investigators have commented that there is
little overall correlation between Kj^ and Kjgg^ data. And yet
the maximum values of Kjgpp often seem to be limited by Kj^.
For steels in the 1500-2000 MPa strength range the data for
tests under freely corroding conditions in aqueous environments
generally show that the maximum Kjs££ values are about half of
Kj£. A relationship of this kind suggest that crack initiation
occurs at the same sites in both types of tests, which may be
plausible for materials with optimum resistance to HE. Regard-
less of the interpretation, the relationship indicates that to
double KJSCC would require doubling the KJQ. This is an
extremely difficult problem that would probably require an
entirely new type of steel. Seen in this light, the magnitude
of the problem of achieving a significantly higher Kjg^ is
apparent.
To date there have apparently not been any practical
attempts to produce a very high-strength steel using the teach-
ings of trap theory. Practical examples of this theory do
exist, however. Hydrogen flakes in plate steels, for example,
are much more numerous in heats with lower sulfur levels, pre-
sumably because the fewer sites for hydrogen occlusion lead to
higher local hydrogen contents in the low sulfur steels (73).

As mentioned several times, yield strength appears to be


the dominating factor controlling hydrogen sensitivity. The
reason that this is so is primarily because of the high local
concentrations of hydrogen that can be achieved at high
strengths. Ritchie (49) has shown that the triaxial stress a
ahead of a crack tip can be approximated by:

a = oy + 2 a K (3)

where ay is the yield stress and a is numerical constant. At


high values of Oy and low values of K this relationship simpli-
fies to:

a = ay (4)

Following Li, Oriani and Darken (74) the concentration of


hydrogen GJ.J in the triaxial stress field is:

CH = C0 exp (J^) (5)

where C0 is the equilibrium concentration in unstressed material,


V the partial molar volume of hydrogen, R the gas constant and
T the temperature. Thus at high yield strengths
a V
v
CH/C0 = exp (-J^) (6)

i.e., the local concentration of hydrogen is proportional to the


exponential power of the yield strength. This, of course, is a
thermodynamic argument that neglects the enrichment in hydrogen
that may take place at specific trap sites. Finally, assuming
that the loss in strength due to hydrogen, Aa^, is proportional
to CH:

Aa H = 3 CH (7)

where 3 is a constant, then from equation (2) the fracture


strength is approximately:
a V
of = a0 - Aa imp - 3C0 exp (-v^) (8)

More detailed versions of this relationship have been


derived by other investigators. For present purposes, however,
the potent effect of yield strength can be seen in equation (8).

OTHER SPECIAL STEELS

Thus far we have discussed conventional steels. Several


other ultra-high-strength steel products are also of interest:

1. Trip Steels

These steels were developed in the late 1960's (75).


Although extensively studied they apparently have not reached
commercial status. The steels are metastable austenitic steels
that transform to martensite during straining. By suitable
compositional adjustment and thermomechanical processing,
strengths about 1500-2000 MPa can be achieved in the austenitic
matrix. During straining, additional plasticity is provided by
the formation of martensite.
Earlier work suggested that steels of this kind had good
resistance to HE. Later work (76,77) showed that these materials
behaved well only if they remained austenitic, presumably because
of the much lower diffusion rate of hydrogen in austenite. When
martensite was formed however, the steels failed quite rapidly
due to cracking in the martensite or along the martensite-
austenite interfaces. Their hydrogen embrittlement susceptibil-
ity then looked comparable to that conventional steels. Whether
a more stable ultra-high-strength austenitic steel that does not
transform to martensite would possess significantly better resis-
tance to HE has evidently not been tested.
2. Cold Drawn Wires
Heavily cold-drawn wires of high-carbon pearlitic steels
have often demonstrated good resistance to HE. These improve-
ments have been related to the greater trap densities, the
decreased dislocation cell sizes, and the axially oriented com-
posite structures with grain boundaries, inclusions, etc., lying
along the wire axis (78). The relaxation of triaxial stresses
in small-diameter wires may also contribute to improve proper-
ties .
More generally, cold-worked structures may be more resistant
to hydrogen. Ryder et al. (79) found much better cracking resis-
tance in a cold-worked pearlite microstructure than when the same
steel was quenched and tempered to the same yield strength.
3. Maraging Steel Composites
Laminar composites of maraging steel with internal layers of
iron, nickel or low-alloy steel have demonstrated excellent
resistance to stress corrosion cracking in 3.5% NaCl solutions
(80). Stress corrosion cracks propagating through the specimens
were stopped when they reached a softer material. Delamination
fractures sometimes were responsible for stopping crack advance,
but in other instances the cracks simply blunted in the softer
layers. Composite welds prepared by alternating filler materials
also showed the ability to stop advancing cracks.
Materials such as these clearly are very specialized items
that often would be difficult to use. For certain applications
however, they may provide ways of achieving significantly better
resistance to hydrogen cracking.

Weldments
Much of what has been said about HE also applies to welds,
but the nature of the welding process causes further problems.
Hydrogen can be easily picked up from various sources during
welding, and precautions must be taken to decrease this absorp-
tion. The thermal history during welding can promote cracking
in the heat-affected zone (HAZ) of the base plate steel and/or
in the weld metal. The weld metal itself is, of course, a cast
structure, with the penalties such microstructures can impose.
Specialized tests such as implant tests (81) or augmented strain
tests (82) have been developed to study hydrogen cracking in
welds, and detailed analyses to predict the simultaneous distri-
butions of stress and hydrogen in weldments are being attempted
(83).
Gooch (84) has summarized the results of extensive stress
corrosion tests of a variety of ultra-high-strength steel welds
in 3% NaCl solutions. Tests of HAZ regions showed steels with
high C contents to be especially susceptible to cracking because
of the high hardnesses in these regions. Post-weld heat treat-
ments that softened the HAZ, particularly when the hardness was
lowered to that of the base plate, considerably improved the
properties. Preheating and post-weld heating also are exten-
sively used to promote hydrogen diffusion out of the weld region.
Cracking in weld metal was influenced somewhat by segrega-
tion during solidification (84) . More important was the micro-
structure. Twinned martensite in particular was found to be
very susceptible to cracking. Once again post-weld heat treat-
ments were helpful because of the tempering of the martensitic
structures. In maraging steels Kenyon, et al. (85) found that
solution annealing and aging after welding gave much better
stress corrosion resistance than just aging.
With suitable post-weld heat treatment it appears possible
to achieve hydrogen cracking resistance in weldments comparable
to base plate properties. With large structures however, such
post-weld heat treatments are not possible. In these cases,
weld properties usually are somewhat inferior to base plate.
Figure 30-8 shows the Kjgcc data for HY-130 weld metal and base
plate versus the cathodic potential. The weld metal values were
consistently about 20 MPa/m lower than the base plate. Maraging
steels at various yield strengths shows a comparable separation
between welds and base plate (85). Both sets of data were for
gas tungsten arc welds, and other welding processes can further
degrade weld properties. Welding processes that produce small
beads give better properties because of the finer-grained struc-
tures produced and because of the greater reheating of previous
deposits by the next weld bead (86). Weld metal chemistry as
such is less important as long as a fine-grained, tempered
martensite structure can be achieved.

DESIGN CONSIDERATIONS
Clearly, successful use of ultra-high-strength steels
require recognition of the potential problems of HE. It also
should be recognized that despite these limitations these steels
have been used successfully in a variety of applications.
Perhaps the best example of use in a critical application is the
employment in landing gears of 30OM steel tempered to tensile
strengths of about 1800 MPa. Such usage clearly requires care-
ful alloy selection, specification, and inspection both before
and during service. Attention must also be paid to the design
to reduce stress concentrations, and to machining and heat-
treating procedures to avoid introducing harmful residual
stresses.
Coatings of all sorts have been used on these steels. Most
coatings act as diffusion barriers to prevent or delay hydrogen
entry into the steel. More recently it has been shown that even
partial coatings may be extremely effective if they retard the
dissociation of hydrogen molecules to form adsorbed hydrogen
atoms at the steel surface (87,88). Thermal treatments that
alter the surface oxide can also have a considerable effect on
hydrogen permeability (89).
BASE METAL

WELD METAL

FRACTURE MODE TRANSITION


MVC INCREASING BRITTLE FRACTURE
PREDOMINANT CLEAVAGE/INTERGRANULAR

E VS A^AgCI (VOLTS)

Figure 50-8: Threshold stress intensities for crack growth


in HY-130 steel base plate and weld metal versus applied
potential (Ref. 86).

In aqueous systems a variety of different inhibitors have


been found effective in retarding hydrogen cracking (90-93). In
gaseous hydrogen the very potent effects of small amounts of
oxygen in preventing crack growth has been noted.
Proof testing may also be a means of retarding crack
growth. Several investigators have noted that preloading KJQ
samples in air considerably improved the KjsCC values in subse-
quent stress corrosion tests (5,94,95). The introduction of
residual compressive stresses at the crack tips is probably the
main cause for the improved properties.

CONCLUDING REMARKS
Hydrogen embrittlement of ultra high-strength steels con-
tinues to be an active area of research, and much has been
accomplished in recent years in achieving a clearer.understand-
ing of this problem. We now have a better picture of hydrogen
entry and transport within the steel, and the roles of trapping
and impurity segregation are becoming clear. The details
surrounding crack nucleation are less certain. Hopefully,
future work will better define the interactions of alloy chemis-
try, micr9Structure and stress state on the initiation of hydro-
gen cracking.
In the practical sense, less has been accomplished. At
lower strengths, of about 1000 MPa, composition and microstruc-
ture are important and some general ground rules have been
established that have helped to produce better properties. At
higher yield strengths it appears that strength itself is the
overwhelming factor that dominates the metallurgical variables.
There has been no improvement in properties to speak of in many
years, and it is difficult even to suggest how one might try to
achieve any significant advance. Composite-type steels seem to
offer this best possibility for combining high strength and
resistance to hydrogen, but these materials clearly would have
limited usage.
While hydrogen cracking is an ever present danger, these
steels have found widespread usage. The key to their success no
doubt resides in the fact that the dangers are well known, and
that procedures have been developed in the production and employ-
ment of these materials that reduce the hazards to acceptable
levels.

REFHRENCEiS

1. Phelps, E. H., "Fundamental Aspects of Stress Corrosion


Cracking," NACE, Houston, TX, p. 398 (1969).
2. Johnson, H. H., "Fundamental Aspects of Stress Corrosion
Cracking," NACE, Houston, TX, p. 439 (1969).
3. Kerns, G. E., Wang, M. T. and Staehle, R. W., "Stress
Corrosion Cracking and Hydrogen Embrittlement of Iron Base
Al_loy^s," NACE-5, Houston, TX, p. 700 (1977).
4. Carter, C. S. and Hyatt, M. V., "Stress Corrosion Cracking
and Hydrogen Embrittlement of Iron Base Alloys," NACE-5
Houston, TX, p. 524 (1977).
5. Dautovich, D. P. and Floreen, S., "Stress Corrosion
Cracking and Hydrogen Embrittlement of Iron Base Alloys,''
NACE-S, Houston, TX, p. 798 (1977).
6. Bernstein, I. M. and Thompson, A. W., Int. Met. Reviews 21:
269 (1976) .
7. Oriani, R. A., Ann. Rev. Mat. Science 8: 327 (1978).
8. Thompson, A. W. and Bernstein, I. M., Adv. in Corrosion
Science § Tech. 7: 53 (1980).
9. Hirth, J. P. and Johnson, H. H., Corrosion 32: 3 (1976).
10. Hirth, J. P., Met. Trans. UA: 861 (1980).
11. Hancock, G. G. and Johnson, H. H., Trans. Metall. Soc. AIME
236: 513 (1966) .
12. Williams, D. P. and Nelson, H. G., Metall. Trans. 1: 63
(1970) .
13. Nelson, II. G., Williams, D. P. and Tetelman, A. S., Metall.
Trans. 2: 953 (1971) .
14. Kerns, G. E. and Staehle, R. W., Scripta Metall. 6: 631
(1972) .
15. Kerns, G. E. and Staehle, R. W., Corrosion 34: 306 (1978).
16. Opoku, J. and Clark, W. G., Corrosion 36: 251 (1980).
17. Srikrishnan, V. and Ficalora, P. J., Metall. Trans . 7A:
1669 (1976).
18. Clark, W. G., J. Mat, for Energy Systems 1: 33 (1979).
19. Bradhurst, D. A. and Heuer, P. M., Corrosion 37: 63 (1981).
20. Lu, M., Pao, P. S., Weir, T. W., Simmons, G. W. and Wei, R.
p
- > Metall. Trans. 12A: 805 (1981).
21. Simmons, G. W., Pao, P. S. and Wei, R. P., Metall. Trans.
9A: 1147 (1978) .
22. Johnson, II. H. and Wiliner, A. M., App. Mater. Research
4: 34 (1965).
23. Lu, M., Pao, P. S., Chan, N. II., Klier, K. and Wei, R. P.,
"Proc. Second JIM Int. Symp. Hydrogen in Metals," Jap.
Inst. of Met. 21: 449 (1980).
24. Hu^ak, S. J. and Wei, R. P., Metall. Trans. 7A: 235 (1976).
25. Gangloff, R. P. and Wei, R. P., Scripta Metall. 8: 661
(1974).
26. Gangloff, R. P. and Wei, R. P., Metall. Trans. 8A: 1043
(1977).
27. Oriani, R. A. and Josephic, P. H., Acta Metall. 22: 1065
(1974).
28. Oriani, R. A. and Josephic, P. H., Acta Metall. 25: 979
(1977) .
29. Akhurst, K. N. and Baker, T. J., Metall. Trans. 12A: 1059
(1981) .
30. Moon, D. M. and Landes, J. D., Scripta Metall. 10: 121
(1976).
31. Stowell, M. J., Davenport, J. and Gradwell, K. J., Metal
Science 12: 192 (1978).
32. Dautovich, D. P. and Floreen, S., Metall. Trans. 4: 2627
(1973).
33. Lee, T. D., Goldenberg, T. and Hirth, J. P., Metall. Trans.
1OA: 439 (1979).
34. Kobayashi, H., Hirano, K., Kayakabe, H. and Nakazawa, II.,
"Proc. Second JlM Int. Symp. Hydrogen in Metals,'' Jap.
Inst. of Met. 21: 477 (1980). ' ~
35. Clark, W. G. and Landes, J. D., "Stress Corrosion - New
Approaches," ASTM-STP 610, ASTM, Philadelphia, PA, p. 108
(1976).
36. Nair, S. V., Jensen, R. R. and Tien, J. K., Metall. Trans.
14A: 385 (1983).
37. Johnson, H. II. and Hirth, J. P., Metall. Trans. 7A: 1543
(1970)
38. Latanision, R. M., Gastine, O. H. and Compeau, C. R.,
"Environment Sensitive Fracture of Engineering
Materials," Metall. Soc. of AIME, Warrendale, PA, p. 48
IT979) .
39. Thompson, A. W. and Bernstein, I. M., "Proc. Second JIM
Inst.__Symp.
__ ^ 1980Hydrogen
^^ in Metals," Jap. Inst. of Met.,
40. Smith, J. A., Peterson, M. H. and Brown, B. F., Corrosion
26: 539 (1970).
41. Hayden, II. W. and Floreen, S., Corrosion 27: 429 (1971).
42. St. John, C. and Gerberich, W. W., Metall. Trans. 4: 589
(1973).
43. Sandoz, G., Metall. Trans. 3: 1169 (1972).
44. Charbonnier, J. C. and Margot-Marette, H., Corrosion
Science 20: 821 (1980).
45. Pao, P. S., Wei, W. and Wei, R. P., "Environment Sensitive
Fracture of Engineering Materials," Met. Soc. of AIME,
Warrendale, PA, 538 (1979).
46. Wei, W. P. , "Environmental Degradation of Eng. Mat. in
Aggressive Environments," Virginia Tech.Printing~T)ept.,
Blacksburg, VA, p. 73 (1981).
47. Shahinian, P. and Sadananda, K., "Pressure Vessels and
Piping Design Technology - 1982 A Decade of Progress,"
ASME, New York, NY, p. 519 (1982).
48. Ritchie, R. 0., "Environment Sensitive Fracture of
Engineering Materials," Metall. Soc. of AIME, Warrendale,
PA, p. 538 (1979).
49. Ritchie, R. 0., Metal Science 11: 368 (1977).
50. Barsom, J. M., "Corrosion Fatigue," NACE-2, Houston, TX,
p. 424 (1972) .
51. Nakasa, K., Takei, H. and Itoh, II., Eng. Fract. Mech. 17:
449 (1983).
52. Frandsen, J. D. and Marcus, H. L., Metall. Trans. 8A: 265
(1977).
53. Ritchie, R. 0., Castro-Cedeno, M., Zackay, V. F. and
Parker, E. R., Metall. Trans. 9A: 35 (1978).
54. VenkatasubramaniaiT^ T. V. and Baker, T. J., Metal Science
16: 543 (1982).
55. Gerberich, W. W., Chen, Y. T. and St. John, C., Metall.
Trans. 6A: 1485 (1975) .
56. Briant, C. L., Feng, H. C. and McMahon, C. J., Metall.
Trans. 9A: 625 (1978).
57. Banerji, S. K., McMahon, C. J. and Feng, H. C., Metall.
Trans. 9A: 237 (1978) .
58. Mabuchi, H. and McMahon, C. J., "Proc. Second JIM Inst.
Symp. Hydrogen in Metals," Jap. Inst. of Met. 21: 441
(1980).
59. Costa, J. E. and Thompson, A. W., Metall. Trans. 12A: 761
(1981) .
60. Craig, B. D., Metall. Trans. 13A: 907 (1982).
61. Craig, B. D. and Krauss, G., Metall. Trans. UA: 1799
(1980).
62. Ritchie, R. 0., Metall. Trans. 8A: 1131 (1977).
63. Takeda, Y. and McMahon, C. J., Metall. Trans. 12A: 1255
(1981) .
64. Craig, B. D., Corrosion 38: 457 (1982).
65. Floreen, S., Inco Alloy Products Co. Research Center,
unpublished Work.
66. Wilde, B. E., Kim, C. D. and Turn, J. C., Corrosion 38:
515 (1982).
67. Thompson, A. W., "Environment Sensitive Fracture of
Engineering Materials," Metall. Soc. AIME,Warrendale,
PA, p. 378 (1979) .
68. Gerberich, W. W. and Wright, A. G., "Environmental
Degradation of Eng. Mat. in Hydrogen," Virginia Tech.
Print Dept., Blacksburg, VA, p. 183 (1981).
69. Pressouyre, G. M. and Bernstein, I. M., Metall. Trans.
12A: p. 835 (1981).
70. Pressouyre, G. M. and Bernstein, I. M., Acta Metall. 27:
89 (1979).
71. Pressouyre, G. M., Acta Metall. 28: 895 (1980).
72. Aucouturier, M., "Current Solutions to Hydrogen Problems
in Steels," ASM, Metals Park, OH, p. 407 (1982) .
73. Ryall, J. E., Barrett, J. F. and Dyer, L. P., Aust. Inst.
of Metals, Metals Forum, 2: 174 (1979).
74. Li, J. C. M., Oriani, R. A. and Darken, L. S., Phys. Chem.
29: 271 (1966).
75. Zackay, V. F., Parker, E. R., Fahr, D. and Busch, R., ASTM
Trans. 60: 252 (1967).
76. Vandervoort, R. R., Ruotola, A. W. and Raymond, E. L.,
Metall. Trans. 4: 1175 (1973).
77. McCoy, R. A. and Gerberich, W. W., Metall. Trans. 4: 539
(1973).
78. Langstaff, D. C., Meyrick, G. and Hirth, J. P., Corrosion
37: 429 (1981).
79. Ryder, D. A., Davies, T. J. and Strecker, E., "Second
International Congress on Hydrogen in Metals," Pergamon
Press, London, Paper 3B2 (1977).
80. Floreen, S., Kenyon, N. and Hayden, H. W., J. Eng. Mat. 5
Tech. 96: 176 (1974) .
81. Neumann, V., Florian, W. and Schonherr, W., "Current
Solutions to Hydrogen Problems in Steels," ASM, Metals
Park, OH, p. 147 (1982) .
82. Savage, W. F., Nippes, E. F. and Tokunaga, Y., Weld. J.
Res. Sup. 57: 126s (1978).
83. Anderson, B. A. B., J. Eng. Mat. § Tech. 104: 249 (1982).
84. Gooch, T. G., Weld. J. Res. Sup. 53: 287s (1974).
85. Kenyon, N., Kirk, W. W. and Van Rooyen, D., Corros ion 27:
390 (1971).
86. Zanis, C. A., Holsberg, P. W. and Dunn, E. C., Weld. J.
Res. Sup. 59: 355s (1980).
87. Chatterjee, S. S., Ateya, B. G. and Pickering, W. H.,
Metall. Trans. 9A: 389 (1978).
88. Kargol, J. A., Fiore, N. F. and Paul, L. D., "Environmental
Degradation of Eng. Mat, in Hydrogen," Virginia Tech.
Print Dept., Blacksburg, VA, p. 55 (1981).
89. Heubaum, F. H. and Berkowitz, B. J., Scripta Metall. 16:
659 (1982).
90. Lynch, C. T., Vahldiek, F. W., Bhansali, K. J. and Summit,
R., "Environment Sensitive Fracture of Engineering
Materials," Metall. Soc. of AIME, Warrendale, PA, p. 639
(1979) .
91. Saito, A., Tokuhiro, U., Yoshizawa, S. and Yamakawa, K.,
"Proc. Second JIM Inst. Symp. Hydrogen in Metals," Jap.
Inst. of Met. 21: 469 (1980).
92. Baker, H. R. and Singleterry, C. R., Corrosion 28: 340
(1972).
93. Parrish, P. A., Chen, C. M. and Verink, E. D., "Stress
Corrosion - New Approaches," ASTM-STP 610, ASTM,
Philadelphia, PA, p. 189 (1976).
94. Carter, C. S., Metall. Trans. 3: 584 (1972).
95. Jonas, 0., Corrosion 29: 299 (1973).

You might also like