Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Previous Page

Intercritical HAZ (Tp « 7800C):

Grain coarsened HAZ (Tp « 13500C):

From this we see that the brittle zones are located 3.5 and 0.5 mm from the fusion bound-
ary, respectively. A comparison with the procedure test results in Table 7.2 shows that the
measured CVN toughness after welding at these locations is slightly higher than that inferred
from the weld thermal simulation experiments. This observation is not surprising, considering
the fact the CVN specimens extracted from the procedure test weld, in practice, include a
wide spectrum of thermal regions which have undergone highly different temperature-time
programmes, whereas the microstructure within the thermally cycled CVN specimens is more
homogeneous due to a similar temperature-time pattern across the whole gauge length (see
Fig. 7.31). Hence, weld thermal simulation cannot replace procedure testing carried out on
real welds. Nevertheless, it is a useful method of evaluating the microstructural stability and
mechanical response of materials to reheating, as experienced in welding.

7.2.3 Hydrogen cracking


Hydrogen embrittlement as a problem is mainly associated with ferritic steels and the risk of
crack initiation in the grain coarsened HAZ following welding.5758 As shown in Fig. 7.32,
these cracks are usually situated at weld toes, weld root, or in an underbead position.
Occationally, hydrogen cracks can also develop in the weld metal. A characteristic feature of
hydrogen-induced cracking is that the process is time-dependent, i.e. the crack may first ap-
pear after several minutes or hours from the time of arc extinction. Consequently, the phenom-
enon is also referred to delayed cracking or cold cracking in the scientific literature.

7.2.3.1 Mechanisms of hydrogen cracking


Hydrogen embrittlement in steels in characterised by:59'60

(i) The crystal structure dependence


Hydrogen embrittlement is mainly associated with materials which exhibit a bcc
or a bet crystal structure, i.e. ferritic and martensitic steels. Austenitic stainless
steels and aluminium alloys with a fee crystal structure are usually not sensitive to
hydrogen.

(ii) The microstructure dependence


A martensitic steel is generally more prone to hydrogen cracking than a ferritic
steel, but a martensitic microstructure is not a requirement for crack initiation.
(a)

Specimen holder

[Homogeneous zone

CVN-specimen

Notch location

(b) Notch location Weld metal

CVN-
specimen Base metal

Grain coarsened HAZ


Grain refined HAZ

Fig. 7.31. Methods for evaluation of HAZ toughness (schematic); (a) Weld thermal simulation, (b) Weld
procedure testing.

(a) (b)

HAZ Transverse crack

Toe
Root irack
Toe crack Underbead
crack crack

Underbead HAZ
crack

Fig. 7.32. Schematic diagrams showing hydrogen-induced cracks in different types of welds; (a) Fillet
weld, (b) Butt weld. The diagrams are based on the ideas of Coe.57
5%/min 104%/min
Uncharged
Uncharged

Charged
True fracture strain

Charged

5x105%/min 1.9x106%/min
Uncharged

Charged
Charged
Uncharged

Test temperature, 0C

Fig. 7.33. Variation of true fracture strain with nominal strain rate and test temperature for charged and
uncharged specimens. Data from Brown and Baldwin.59

(iii) The strain rate dependence


Hydrogen embrittlement is most prominent at low strain rates typical of tensile
testing, as shown in Fig. 7.33. At high strain rates the hydrogen diffusion is not
fast enough to keep pace with the fracture development.

(iv) The temperature dependence


Hydrogen cracking occurs usually within the temperature range from -150 to
+2000C. This temperature dependence reflects the fact that both the hydrogen
concentration and the stress intensity at the crack tip must exceed some critical
value before crack propagation occurs.

(v) The time dependence


Since hydrogen embrittlement is a diffusion-controlled process, the cracks will
propagate in a stepwise manner to allow for supply of hydrogen from the sur-
rounding matrix to the crack tip (see Fig. 7.34).

Over the years a number of mechanisms have been proposed to explain the origin of hydro-
gen embrittlement. The three most important are:
Fracture
Resistance change x 10 ,Q.
-8

Fracture
Applied stress:
1240MPa
Applied stress:
1100MPa

Time, min
Fig. 7.34. Example of stepwise crack propagation in notched tensile specimens, as inferred from electri-
cal resistivity measurements. Data from Steigerwald et a/.60

(a) The hydrogen gas pressure model, originally proposed by Zapfee and Sims,61 which
postulates that atomic hydrogen will diffuse to microvoids where it recombines to
form molecular hydrogen. In ferritic steels the equilibrium H2(g) pressure within
the microvoids is typically of the order of 106 to 107 atm, which is more than
sufficient to bring about a local fracture development.

(b) The surface energy model (Petch62). According to this model hydrogen will re-
duce the effective surface energy of the crack. Under such conditions the crack
can propagate at a lower nominal stress in the presence of hydrogen, in agreement
with the Griffith's theory (equation (7-5)).

(c) The slip softening model of Beachem,63 which accounts for the experimental ob-
servation that hydrogen-charged specimens generally exhibit a lower flow stress
than hydrogen-free specimens. This suggests that hydrogen interfers with dislo-
cations in a manner which facilitates different types of fracture, including micro void
coalescence (or dimpled rupture), quasicleavage fracture, and intergranular frac-
ture.64

Currently, it cannot be stated with certainty which of these three mechanisms that are opera-
tive under the conditions existing in welding. However, this question is of minor importance
in the present context, since we here are mainly concerned with the factors responsible for
hydrogen cracking in steel weldments.
7.2.3.2 Solubility of hydrogen in steel
Since hydrogen is the smallest of all atoms, it is readily soluble in iron. In general, both
octahedral and tetrahedral lattice sites are potential traps for interstitials, as indicated in Fig.
7.35. In the case of hydrogen it is believed that the dissolved atoms are mainly present in
tetrahedral positions in the form of protons.65 Because of the pertinent difference in the size of
the fee and the bcc interstices (see Fig. 7.35), the hydrogen solubility in iron will change
stepwise with temperature following the bfe —> yFe and yFe —> aFe transformations, as shown
previously in Fig. 2.7(c) (Chapter 2).

(a)

Metal atom Metal atom


Octahedral interstices Tetrahedra! interstices
(Size: 0.15 aQ) (Size: 0.08 aQ)

(b)

Metal atom Metal atom


Octahedral interstices Tetrahedral interstices
(Size: 0.07 a0) (Size: 0.13 a 0)

Fig. 7.35. Schematic representation of octahedral and tetrahedral lattice sites in different crystal struc-
tures; (a) Face-centred cubic (fee) structures, (b) Body-centred cubic (bcc) structures.
In addition to the interstitial fraction, hydrogen may be present in the form of molecular
(gaseous) hydrogen trapped in micro voids or plane lattice defects. This amount is frequently
referred to as residual hydrogen, and can in many cases overshadow the equilibrium hydrogen
content. For example, at room temperature the maximum solubility of atomic hydrogen in the
iron lattice is estimated to be 0.001 to 0.01 ppm, while the analytical hydrogen content of
steels varies typically from 1 to 10 ppm. This supersaturation is formidable and provides the
necessary driving force for trapping of gaseous hydrogen in the microstructure.

7.2.3.3 Diffusivity of hydrogen in steel


Published data for the diffusivity of hydrogen in steels are summarised in Fig. 7.36. At high
temperatures, the diffusivity of hydrogen in ferritic steels is in reasonable agreement with the
reported value for lattice diffusion of hydrogen in bcc iron. However, when the temperature
drops below say 2000C, both the scatter and the discrepancy become more apparent due to the
phenomenon of hydrogen trapping. Inclusion of the trapping effect has led to the introduction
of an apparent diffusion coefficient for hydrogen in ferritic steels, D%pp , which according to
Oriani66 is given by:

(7-12)

where D^ is the lattice diffusion coefficient for hydrogen in bcc iron, K is the density of trap
sites (i.e. number of trap sites per number of lattice sites), and EB is the binding energy be-
tween hydrogen and the trap site.
A graphical representation of equation (7-12) is shown in Fig. 7.36. A closer inspection of
the graph reveals that the predicted temperature dependence of the apparent diffusion coeffi-
cient is in fair agreement with the reported diffusivity data for hydrogen in steel. Moreover, it
is interesting to note that the hydrogen diffusion coefficient in austenite is nearly two orders of
magnitude lower than the corresponding value for the ferrite phase at a given temperature.
This observation is not surprising, considering the pertinent difference in the packing density
between the fee iron lattice and the bcc iron lattice (74% and 68%, respectively). Thus, for
diffusion of hydrogen in austenite, we have:57

(7-13)

where T is the absolute temperature (in K).

7.2.3.4 Diffusion of hydrogen in welds


The thermodynamics and kinetics of hydrogen absorption in the weld pool have previously
been discussed in Section 2.8 (Chapter 2).
Since hydrogen is a very mobile atom (and therefore is easily lost to the surroundings), the
hydrogen concentration will vary both in the longitudinal direction and in the through thick-
ness direction of the weld, as shown in Fig. 7.37. This process will continue even after the
weld has cooled down to room temperature due to the characteristic high diffusivity of atomic
hydrogen in ferrite (see Fig. 7.38).
Temperature, 0C

Hydrogen diffusion coefficient, m2/s

Ferritic steels
Trapping
theory

Austenitic
steels

1000/T1K'1

Fig. 7.36. Summary of reported diffusion coefficients of hydrogen in iron and steel. Data compiled by
Coe57 and Yurioka and Suzuki.58

Several successful attempts have been made in the past to model hydrogen diffusion in
welds by means of numerical methods.68"70 Unfortunately, none of these solutions are simple
enough to get a good overall indication of the hydrogen redistribution during cooling and
subsequent PWHT. As an illustration of principles, we shall therefore present a simplified
analytical solution to the hydrogen diffusion problem in welding, based on an analogy be-
tween diffusion and heat conduction.

Diffusion model
The idealised model considers a butt weld of uniform hydrogen concentration in the longitudi-
nal direction, as shown in Fig. 7.39. The width of the fusion zone is 2L, while the initial
hydrogen concentration at the time of solidification (i.e. at t = 0) is Q. The hydrogen con-
centration in the base metal outside the fusion zone is C0. If element losses to the surroundings
are neglected, the problem can be treated as uniaxial diffusion in an isotropic solid analogous
to that described in Section 1.7 (Chapter 1) for heat conduction in thermit welding. Thus, in
the limiting case where the diffusion coefficient can be regarded as constant, the hydrogen
concentration ( Q as a function of time (t) and distance (y) is given by equation (1-22):
Deposited metal
ml H 2 /10Og

Fused metal

Position x, mm
ml H 2 /100g

Mean value

Diffusible
Residual
10 mm

Fig. 7.37. Measured longitudinal and lateral distributions of hydrogen in a single pass SMA weld quenched
right after welding. Data from Christensen et al.61

(7-14)

where D** is the hydrogen diffusivity, and erf(u) is the Gaussian error function (defined previ-
ously in Appendix 1.3, Chapter 1).
In practice, it is necessary to rewrite equation (7-14) in a differential form to allow for the
variation in the hydrogen diffusion coefficient with temperature. After some manipulation, we
obtain:

(7-15)
H I/cm3

15 mm
a)

p. I/cm3

b)

Fig. 7.38. Redistribution of hydrogen following welding (numerical calculations); (a) Right after weld-
ing, (b) After 12 h at room temperature (ljil cm"3 = 0.0115 ppm). Data from Christensen.68

Fusion zone

Fig. 7.39. Sketch of idealised hydrogen diffusion model.


This differential equation can be integrated numerically in temperature-time space when
the weld thermal programme is known. The boundary conditions are as follows:

when when

when when

Case Study (7.2)


Although the above model does not give a true physical picture of the hydrogen redistribution
in butt welds, it may provide valuable quantitative information about the extent of hydrogen
diffusion occurring during cooling from the solidification temperature under different welding
conditions. Figure 7.40(a) shows a sketch of a 2mm thick single pass butt steel weld made by
means of the GTA process. For the purpose of convenience we shall assume that the tempera-
ture field around the heat source is given by the simplified Rykalin thin plate solution (equa-
tion (1-100) in Chapter 1). Thus, at (T-T0)= 15000C a total width of the fusion zone of about
4.3mm is obtained for a net heat input of 67.2 J mm"2. The corresponding distance from the
weld centre-line to the 13500C HAZ isotherm is 2.5mm.
Figure 7.40(b) shows computed temperature and hydrogen concentration profiles at the
centre of the weld (y = 0) and in the grain coarsened HAZ (y = 2.5mm) for a chosen ambient
temperature of 200C (no preheating). As expected, the hydrogen concentration within the
fusion zone itself (v < L) decreases in a monotonic manner as the weld cools down. In contrast,
the hydrogen concentration outside the fusion zone (y > L) is seen to pass through a local
maximum. In the absence of preheating this maximum is attained after very long times, which
may initiate hydrogen cracking in the grain coarsened HAZ if the microstructure is martensitic.
The picture is completely changed if the ambient temperature is raised to 1000C (moderate
preheating). As shown in Fig. 7.40(c), the main effect of preheating is to decrease the cooling
rate in the low-temperature regime (i.e. below 5000C) after the completion of the austenite to
ferrite transformation. The HAZ microstructure is therefore not significantly altered, but in-
stead more hydrogen is allowed to diffuse out of the weld region before the temperature drops
below the critical value where hydrogen cracking may occur. This is seen as a shift in the peak
hydrogen concentrations towards lower absolute values and shorter times in Fig. 7.40(c).

7.2.3.5 Factors affecting the HAZ cracking resistance


Safety against hydrogen cracking is an important aspect of weldability. In spite of the knowl-
edge accumulated and the improvements made over the past decades, the current trends to-
wards stronger steels and heavier sections still require continuous attention to the risk of cracking.
Cold cracking test methods
In principle, a proper weldability criterion should enable the user to select combinations of
steel, consumables and operational conditions that will ensure sufficient crack safety at a mini-
mum of total cost. It should also enable him to examine the effects of these main variables
separately and establish a quantitative grading system for safety which takes into account the
consequences of failure.
In recognition of this situation, a number of empirical cracking test methods has been de-
Tp =
HAZ iaiMMlTOMsl 1 3500C

(b)
Temperature-time Solid curves: Centre - line
programme Broken curves: HAZ
T0 = 200C

(C-C0)Z(C1-C0)
Temperature, 0C

Hydrogen
concentration

Time, s

(C)
Temperature-time Solid curves: Centre - line
programme Broken curves: HAZ
T0 = IOO0C
(C-C0)Z(C1 -C 0 )
Temperature, 0C

Hydrogen
concentration1

Time, s
Fig. 7.40. Computed temperature and hydrogen concentration profiles during GTA butt welding of a
2mm thin steel sheet (Case Study (7.2)); (a) Sketch of weld, (b) Redistribution of hydrogen in the ab-
sence of preheating, (c) Redistribution of hydrogen after preheating to 1000C.
veloped over the years to study the mechanisms of hydrogen cracking in weldments (e.g. see
the review of Yurioka and Suzuki58). Broadly speaking, the cold cracking tests fall into either
one of the two categories, i.e. self-restrained tests or externally loaded tests. Examples of the
former type are the Tekken (oblique Y-groove) cracking test, the CTS (controlled thermal
severity) cracking test, and the cruciform cracking test. Well-known externally loaded tests
are the implant cracking test, the TRC (tensile restraint cracking) test, and the RRC (rigid
restraint cracking) test.

The implant method


The implant technique is a good example of a cracking test method which allows separate
assessment of the various metallurgical and operational factors that contribute to hydrogen
cracking in welds. As shown in Fig. 7.41 the Scandinavian version of the implant test employs
a cylindrical test bar made from the steel to be examined which is notched (threaded) at some
distance from one of its ends.71 The bar is inserted into a reamed hole in the backing plate of
a similar grade of steel, so that the threaded end is flushed with the plate surface. In order to
execute a test, the hole is sealed off with a single weld bead deposited on the plate under
carefully controlled conditions. After a specified delay of 60 s per kJ mm"1 gross heat input,
the implant assembly is quenched and put under a static tensile load at room temperature. The
stress applied is subsequently reported on the nominal cross-section of the bar. After a speci-
fied loading time (if no rupture has occurred), the assembly backing plate with the test bar may
be sectioned and examined with respect to microcracks in the HAZ.
The implant test may be performed at different stress levels, cooling rates and hydrogen
contents. The implant rupture strength, RIR, is obtained in a programme of stepwise loading,
going down or up by a fixed amount depending on rupture or not. It is then defined as the
nominal stress at which the statistically probability of rupture is 50% when the load is applied
for a long period.

Implant test results


The critical stress for rupture in the Scandinavian version of the implant test may be written in
a differential form as a sum of four contributions:

Test weld

Implant test Backing


plate

Tensile load

Fig. 7.41. Schematic representation of the implant test method.


(7-16)

The CEW parameter in equation (7-16) refers to the so-called HW carbon equivalent, origi-
nally developed for C-Mn steels:

(7-17)

Moreover, A/333 is a hydrogen diffusional parameter which takes into account variations in
the measured implant rupture strength after various thermal treatments (including preheating
and PWHT). According to Christensen and Simonsen,71 the extent of hydrogen diffusion
which occurs in the low-temperature regime can be reported in the form of an equivalent
isothermal hold time at 6O0C (or 333K), defined as:

(7-18)

where Tc refers to the local HAZ temperature at the moment of quenching (usually taken as
1000C).
It follows from equation (7-16) that the two first members reflect the influence of micro-
structure upon the implant rupture strength, and is therefore related to the HAZ peak hardness.
The two last members take into account the effect of analytical HFM and local hydrogen con-
centrations. As shown in to Fig. 7.42, the numerical values of the partial derivatives dRIR/
dCEw, dRIR/dAts/5, 3RIR/3 log HFM, and dRIR I B^At333 may vary within relatively wide limits,
depending on the steel chemical composition and the operational conditions applied. Never-
theless, the concept is still useful for quantitative predictions of the HAZ cracking resistance,
as illustrated below.

Example (7.4)
Experience has shown that conventional pipeline steels with carbon equivalent CEW up to
0.4% can be welded with basic electrodes (Af8/5 ~ 8-9 s) without the use of external preheating,
provided that the weld metal hydrogen content is kept sufficiently low (HFM ~ 4 ppm). Sup-
pose that the same procedure shall be employed in hyperbaric welding of pipeline steels at a
depth of 320 m (33 bar total pressure). Based on the implant test data in Fig. 7.42, estimate the
minimum reduction in the steel carbon equivalent (ACEW) which must be incorporated in the
specifications to compensate for the increased hydrogen absorption observed at such depths
(#FM-10ppm).

Solution
The concept of partial derivatives implies that we will have the same safety against hydrogen
cracking if there is no net change in the implant rupture strength (i.e. ARm = 0). Since the weld
(a)
HSLA
R|R,MPa
steels

CEW,%

(b)

HSLA steel
R|R, MPa

At8/5, S

Fig. 7.42. Examples of implant test results: (a) Rm vs CEW, (b) RIR vs Ar8/5.

cooling programme is similar in both cases, the variation in A/8/5 and ^At333 can be neglected.
Hence, equation (7-16) reduces to:
(C)
HSLA steel

R1R, MPa

HFM, ppm

(d)
R|R, MPa

Quenched and tempered steels

^ • • "

Fig. 7.42. Examples of implant test results (continued); (c) RIR vs HFM, (d) RIR vs -^Ar333 . Data from
Christensen and Simonsen.71

In the present example the total change in the weld metal hydrogen content between 1 and
33 bar is equal to:
Moreover, the numerical values of dRIR/dCEw and dRIR/d log HFM can be read from Fig.
7.42(a) and (c), respectively. When A/8/5 ~ 8.6 s, we obtain:

and

This gives:

The above calculations suggest that the CEW carbon equivalent of pipeline steels should not
exceed 0.35% if hydrogen cracking is to be avoided under hyperbaric welding conditions.

7.2.4 H2S stress corrosion cracking


Hydrogen sulphide (H2S) stress corrosion cracking is a well-known phenomenon taking place
in steels in environments containing sour oil and gas. As shown in Fig. 7.43 this type of
cracking arises from corrosion reactions with subsequent absorption of hydrogen in the metal.
The embrittlement mechanisms are therefore similar to those reported for hydrogen cracking
in steels, and can be evaluated from standard test methods.72

7.2.4.1 Threshold stress for cracking


For a given combination of steel, microstructure and H2S concentration there exists a lower
limit for the imposed stress where cracking no longer will occur. Experience has shown that
the threshold stress, vth, is related to the yield strength RpQ2 through the following equation:73

(7-19)

Equation (7-19) predicts that the threshold stress (and thus the steel cracking resistance)
passes through a local maximum as the yield strength increases. The locus of this peak stress
is obtained by setting dath/dRpo2 = 0, which gives Rpo2 ~ 600MPa and crth (max) ~ 360MPa.
In practice, a hardness criterion rather than a yield strength criterion is used for ranking of
steels with regard to H2S stress corrosion cracking resistance. According to Dieter,19 the fol-
lowing relation exists between Rpo and HV:

(7-20)

where m is the strain hardening exponent in the Ludwik equation.


The value of m may vary between wide limits, depending on the steel chemical composi-
tion and the heat treatment conditions applied, but for HSLA steels m ~ 0.15 is a reasonable
compromise.19 In that case the observed maximum in the cracking resistance at Rp02 ~600MPa
Short distance between
anode and cathode sites

Stress
Anode:

Crack Embrittled zone

Cathode:

Stress

Fig. 7.43. Mechanisms of hydrogen absorption in cathodic stress corrosion cracking (schematic).

corresponds to a hardness of about 250VPN. This value should be compared with the maxi-
mum hardness level of 22HRC (Rockwell C) or 248VPN incorporated in many offshore speci-
fications.

7.2.4.2 Prediction of HAZ cracking resistance


Prediction of the H2S stress corrosion cracking resistance based on equation (7-19) requires
quantitative information about the HAZ strength level. The peak strength for various combi-
nations of steels and welding conditions can easily be read from diagrams of the type shown in
Fig. 7.20 or calculated from diverse empirical formulae.2674 Figure 7.44 shows examples of
computed cr^ - At^ profiles for three different types of steels spanning a range in the CEn
carbon equivalent from 0.41 to 0.50%. It is evident from these plots that the cooling time
required to obtain the maximum threshold stress depends on the steel chemical composition.
In general, ultra-low-carbon steels will exhibit the highest resistance against stress corrosion
cracking, since a low HAZ hardenability will eliminate problems with martensite formation in
the grain coarsened region during the 7 to a transformation. On the other hand, these steels
suffer from a severe HAZ softening during high heat input welding, with consequent reduction
in the threshold stress. Under such conditions it may be safer to use ordinary low-carbon
microalloyed steels or C-Mn steels with a higher HAZ hardenability to compensate for the
observed strength loss.
Equation (7-19) can also be employed for assessment of the relevance of current hardness
requirements. Figure 7.45 shows the same types of plots as in the preceding figure (on a
normalised scale) where typical ranges for the HAZ hardness are indicated. It is evident that
the threshold stress is rather insensitive to small variations in the hardness level. Only in cases
where the HAZ hardness exceed 300VPN or drops below say 220VPN a significant deteriora-
tion in the H2S stress corrosion cracking resistance is to be expected. This suggests that the
maximum hardness requirement of 248VPN incorporated in many offshore specifications is
too stringent, since it imposes severe restrictions on the use of steels in welded structures
without improving the service performance to any great extent.
Next Page

Threshold stress, MPa

Ultra-low-carbon steel
Low-carbon steel
C-Mn steel

A
WS
Fig. 7.44. Computed a f/l -A% 5 profiles for selected steels.
Normalized threshold stress

Ultra-low-carbon steel
Low-carbon steel
C-Mn steel

A s
w

Fig. 7.45. Effect of peak hardness on the HAZ stress corrosion cracking resistance.

You might also like