Hic

You might also like

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

hydrogen-induced cracking

Hydrogen-induced cracking (HIC) is still a challenging issue that requires careful


assessment and utilization of test validation measures.

From: Handbook of Materials Failure Analysis with Case Studies from the Oil and
Gas Industry, 2016

Related terms:

Carbon Steel, Pipelines, Corrosion, Stress Corrosion Cracking, Hydrogen, Hydrogen


Sulfide

View all Topics

Learn more about hydrogen-induced cracking

General Requirements for the Purchase


of Pipes for Use in Oil and Gas Indus-
tries
Alireza Bahadori PhD, CEng, MIChemE, CPEng, MIEAust, RPEQ, in Oil and Gas
Pipelines and Piping Systems, 2017

8.17.1 Hydrogen-Induced Cracking Tests


Hydrogen-induced cracking (HIC) tests should be performed and reported in ac-
cordance with NACE TM 0284-96. Test Solution “A” should be used. Values of
crack-length ratio (CLR), crack-thickness ratio (CTR), and crack-sensitivity ratio (CSR)
should be reported and photomicrographs of specimens showing any blister, or
alternatively dimensioned sketches, should be provided with the report.

Temperature 25±3°C
H2S concentration (2300–3500 ppm) Saturated condition
pH Value-initial 2.9–3.3
pH Value-final 3.5–4.0
Test period 96 hours

> Read full chapter

Defining the problem


N BAILEY, ... R J PARGETER, in Welding Steels Without Hydrogen Cracking (Second
Edition), 2004

Hydrogen-induced cracking in welds


Hydrogen-induced cracking is also known as cold cracking, delayed cracking or
underbead cracking. Unfortunately, the term ‘hydrogen-induced cracking’, usually
abbreviated as HIC, has been introduced to designate cracking sometimes encoun-
tered in pipelines or vessels as a result of hydrogen picked up in service; the term
‘fabrication hydrogen cracking’ is therefore preferred for the subject of this book
whenever there may be any doubt. It also occurs in steels during manufacture,
during fabrication and in service. It is thus not confined to welding, but when it
occurs as a result of welding the cracks are sited either in the HAZ of the parent
material or in the weld metal itself. Brief descriptions are followed by a discussion
of the factors which are responsible for cracking and of the means by which control
may be achieved.

Cracking in the HAZ


Hydrogen-induced cracking occurs when the conditions outlined in 1–4 (below)
occur simultaneously

1 Hydrogen is present to a sufficient degreeThis is inevitably present, derived from


moisture in the fluxes used in welding and from other sources. It is absorbed
by the weld pool and some is transferred to the HAZ by diffusion.
2 Tensile stresses act on the weldThese arise inevitably from thermal contrac-
tions during cooling and may be supplemented by other stresses developed
as a result of rigidity in the parts to be joined.
3 A susceptible HAZ microstructure is presentThat part of the HAZ which expe-
riences a high enough temperature for the parent steel to transform rapidly
from ferrite to austenite and back again produces microstructures which are
usually harder and more susceptible to hydrogen embrittlement than other
parts of the HAZ. Hydrogen cracks, when present, are invariably found in these
transformed regions.
4 A low temperature is reachedThe greatest risk of cracking occurs when tempera-
tures near ambient are reached and cracking may thus take place several hours
after welding has been completed. Cracking is unlikely to occur in structural
steels above about 150°C, and in any steel above about 250°C.

Cracks in the HAZ are usually sited either at the weld toe, the weld root or in an
underbead position. These positions are shown schematically for fillet welds and
butt welds in Fig. 1.1. In fillet welds, HAZ cracks are usually oriented along the weld
length, but in butt welds subsurface cracks can be transverse to the weld. Hydrogen
cracks examined in sections of a weld under the microscope may be intergranular,
transgranular or a mixture with respect to the transformed microstructures in which
they lie.

1.1. Hydrogen-induced cracks in HAZs of (a) fillet and (b) butt welds.

Intergranular cracking is more common in harder, higher carbon and more highly
alloyed steels. Cracks may vary in length from a few microns to several millimetres.
Some typical HAZ cracks are shown in the photomacrographs, Fig. 1.2.

1.2. Heat-affected zone in C-Mn steel (a) hydrogen crack at root of single-run fillet
weld, and (b) crack at toe of multipass fillet weld.
Cracking in the weld metal
Hydrogen cracking can occur in the weld metal as well as in the HAZ. Weld metal
hydrogen cracks can be orientated longitudinally or transverse to the weld length,
while in the transverse orientation they can be either perpendicular or angled, typ-
ically at approximately 45° (often referred to as chevron cracks), to the weld surface.
The cracks may be buried or may break the weld surface. Under the microscope
they are usually recognised as being predominantly transgranular, although in more
alloyed deposits there is an increasing proportion with intergranular morphologies.
Typical examples may be seen in Fig. 1.3.

1.3. Weld metal hydrogen cracks in (a) single-run manual metal-arc fillet weld, (b)
root bead of Y groove welding test, (c) submerged-arc weld (longitudinal section).

The same factors which influence the risk of cracking in the HAZ also apply to
weld metal, namely, stress, hydrogen, susceptible microstructure and temperature.
However, weld metal hydrogen cracking can occur at much lower levels of weld metal
hardness than is generally the case for HAZ cracking.

> Read full chapter


Nonmetallic Inclusions and their Ef-
fects on the Properties of Ferrous Alloys
R. Kiessling, in Encyclopedia of Materials: Science and Technology, 2001

7 Hydrogen and Inclusions


Hydrogen-induced cracking (HIC) in steel is a risk problem when steels are charged
with hydrogen in service as in various hydrogen sulfide environments like sour oil
and gas wells and some power plants. Also high strength structural steels which are
welded are exposed to a certain hydrogen pick-up which means a risk at fracture and
blistering due to HIC.

Sulfide inclusions are known to affect HIC (Kiessling 1989, p. 135). The interfaces
between these inclusions and the steel matrix provide sinks (microvoids) in which
hydrogen present in the steel may be trapped and hydrostatic pressure may build up.
Achievement of good HIC resistance requires desulfurization to low levels, typically
below 0.002% sulfur, with inclusion modification. Calcium treatment has been used
with success but control of the Ca/S ratio has been found to be important. An excess
of calcium will lead to the formation of clusters of CaS inclusions reducing the HIC
resistance. See also Fig. 1.

Steels treated with rare earth metals also exhibit a low HIC caused by the sulfide
modification, probably explained by a decrease in the area of the sulfide–matrix
interface. A second effect of the sulfide inclusions is that they poison efficiently
the reaction of atomic hydrogen in forming H2 molecules. Thereby the pick-up of
hydrogen at a steel surface subject to corrosion that creates atomic hydrogen is
promoted. The hydrogen absorption of the steel and the pressure build-up are so
large under many conditions that HIC occurs without applied stress, for example,
during exposure of a steel to H2S-saturated seawater.

> Read full chapter

Corrosion
Michael J Schofield BSc, MSc, PhD, MIM, CEng, in Plant Engineer's Reference Book
(Second Edition), 2002

33.1.3.3 Hydrogen-induced cracking


Hydrogen-induced cracking (HIC) is most commonly encountered in steels but oth-
er metals are susceptible, as shown in Table 33.3. The presence of hydrogen atoms in
a metal degrades some of its mechanical properties, especially its ductility, leading in
some cases to embrittlement. Additionally, hydrogen atoms diffuse through metals
and coalesce to form hydrogen molecules at certain preferred locations such as
inclusions.

Table 33.3. Hydrogen-induced cracking

Metal Agent
Steel Hydrogen gas
Atomic hydrogena
Hydrogen sulphide-containing aqueous environ-
mentsb
Waterc
Martensitic Hydrogen sulphide-containing
stainless steels aqueous environments
Copper alloys Sea waterd
Nickel alloys Hydrogen sulphide-containing aqueous environ-
ments
Titanium alloys Anhydrous alcohols
Zirconium alloys High-temperature water

a Created by overprotection during cathodic protection, evolved during electro-


plating or pickling processes and during welding with wet welding consum-
ables.
b Steels of strength in excess of 550 MPa.

c Steels of strength in excess of 900 MPa.

d High-strength alloys, cathodically protected.

Hydrogen atoms are soluble in ferrite (which is the major phase of most steels).
At discrete inclusions in the steel (e.g. manganese sulphide, which is present in
many steels) the hydrogen atoms combine to form hydrogen molecules. Hydrogen
molecules are insoluble in the steel lattice and these molecules associate, produc-
ing high tensile stresses which can initiate hydrogen cracking. When laminations
are present in the steel hydrogen molecules form at the interface between them,
creating blisters. At high temperatures hydrogen atoms can react with carbides in
the steel producing methane gas. Methane is insoluble in the steel structure and
causes blistering or cracking. In clean steels containing none of these features the
hydrogen atoms pass straight through the steel, causing no damage.

Hydrogen can originate from several sources:

1. Electroplating (e.g. zinc, cadmium or hard chromium plating);

2. Acid pickling (e.g. prior to galvanizing);


3. Cathodic protection, especially if overprotected, through inadequate potential
control;
4. Corrosion, where H+ reduction is a cathodic reaction (Section 33.1.2);

5. The process inside the plant.

The amount of the hydrogen that is liberated on or near a metal surface which then
enters the metal varies according to the environment and condition of the metal.
The main factor that promotes the entry of hydrogen into a metal is the presence
on the metal of a surface poison such as sulphide or other species which inhibit the
hydrogen recombination reaction.

> Read full chapter

The role of grain boundaries in hydro-


gen induced cracking (HIC) of steels
C.J. McMahonJr., in Gaseous Hydrogen Embrittlement of Materials in Energy Tech-
nologies: Mechanisms, Modelling and Future Developments, 2012

6.1 Introduction: modes of cracking


There are two modes of hydrogen induced cracking (HIC), in steels, and they are
fundamentally different. The more common one is intergranular. An example is
shown in Fig. 6.1(a). It is stress-controlled, can occur in bursts, and involves the loss
of cohesion along prior austenitic grain boundaries. Except in specially prepared
laboratory heats, the loss of cohesion is exacerbated by the presence of segregated
elements of the kind that produce temper embrittlement in steels, as well as other
embrittling elements, like sulfur, tellurium, and oxygen.1

6.1. (a) 4340 steel in 10 psi (68.9 KPa) H2 and (b) HY 130 steel in 1500 psi (10.33 MPa)
H2.
The other mode of HIC is strain-controlled and therefore occurs in a stable manner
at a rate dependent upon the rate of straining.2 The cracking occurs within the prior
austenite grains in a quenched-and-tempered steel, rather than along the grain
boundaries, and the morphology is difficult to characterize, as illustrated by Fig.
6.1(b). The presence of flat facets has led some workers to call it quasi-cleavage.
It has been suggested that such facets are examples of glide-plane decohesion
caused by the collection and transport of hydrogen by dislocations moving along a
martensite lath or plate, plus the tensile stress across the slip band when it becomes
blocked, as shown schematically in Fig. 6.2.3 Other areas have the appearance of
separated boundaries between colonies of tempered martensite. None of this can
be considered definitive, although it has been established that this plasticity-related
HIC occurs preferentially along planes of maximum shear stress.2

6.2. Schematic representation of (a) hydrogen transport in dislocation cores in a


blocked slip band and (b) hydrogen-induced decohesion along a lath boundary or
the glide plane in a steel with a microstructure of tempered martensite.3

> Read full chapter


Assessing and modelling the perfor-
mance of nuclear waste and associated
packages for long-term management
T.M. Ahn, in Radioactive Waste Management and Contaminated Site Clean-Up,
2013

7.4.1 Hydrogen-induced cracking of carbon steel


Carbon steel may be susceptible to hydrogen-induced cracking (Kobayashi et al.,
2011). Hydrogen is likely to be produced by radiolysis of groundwater and by water
reduction during the corrosion process in reducing conditions. Figure 7.6 shows
reduction in area versus applied potential in constant extension rate testing (CERT)
for ASTM A216-Grade WCA steel in concentrated synthetic groundwater at 80 °C
(176 °F) (Ahn and Soo, 1995). This hydrogen-induced cracking can be regarded as a
variation of SCC. Recently, the surface opening area resulting from SCC container
damage has been assessed for various candidate container metals including carbon
steel. Generally, the maximum opening area is approximately 0.1% of the total
surface area of the waste package (Gwo et al., 2011). This original assessment was
made with impact stress in the deformed area of the container as an effect of seismic
impact. However, this assessment may be applicable to the normal static case too.
The weld residual stress and a weld area can be used instead of impact stress and
deformed area from seismicity. The models for estimating opening area due to SCC
are described in the following section for stainless steel canisters. This approach is
equally applicable to the hydrogen-induced cracking of carbon steel.
7.6. Reduction in area versus applied potential in constant extension rate testing
(CERT) for ASTM A216-Grade WCA steel in concentrated synthetic groundwater at
80 °C (Ahn and Soo, 1995). SCE: standard calomel electrode. Used with permission
from Elsevier.

> Read full chapter

Industrial Consequences of Hydrogen


Embrittlement
Laurent Briottet, ... Flavien Vucko, in Mechanics - Microstructure - Corrosion Cou-
pling, 2019

10.2.1 Oil and gas industry


Regarding applications in the oil and gas industry, hydrogen-induced cracking can
occur on a very wide variety of steel equipment: drill strings, underground hydro-
carbon extraction tubes (Oil Country Tubular Goods), pipelines, casing maintenance
pieces, etc. The steel grades (carbon, stainless), as well as nickel-based alloys, used
cover a broad material spectrum and are selected according to the environment and
mechanical stress applied. The hydrogen can come from the equipment’s cathodic
protection or from corrosion induced by the aqueous medium generating hydrogen.

Hydrogen sulfide, H2S, found naturally in the deposit or released by sulfidogenic


bacteria, acts as a catalyst for hydrogen ingress in the steels. Cracks in the presence
of H2S have punctuated the history of oil and gas production on many occasions,
such as in the development of the Lacq gas field in France in the 1950s [SUD 17],
but are still an issue nowadays as evidenced by the break of a pipeline in the giant
Kashagan deposit in September 2013 [NCO 14].

10.2.1.1 Cracking in carbon steels in the presence of hydrogen sulfide

10.2.1.1.1 Mechanism elements and fracture modes encountered

Hydrogen sulfide promotes the absorption of hydrogen by steel. A first degradation


mode is commonly known as hydrogen-induced cracking (HIC) and occurs by
the generation and propagation of internal cracks, following hydrogen diffusion
towards specific sites of the microstructure (band of ferrite/pearlite microstructures,
elongated MnS-type inclusions, etc.) where its accumulation and recombination into
hydrogen gas causes internal pressures of several hundred bars, enough to crack the
steel core (HIC) and create blistering on the surface.
A second mode of degradation is sulfide stress cracking (SSC) with the presence of
residual and/or applied stresses. This cracking initiates at the surface and propagates
perpendicular to the direction of the main principal stress (Figure 10.1). Following
the nucleation of a surface crack, the combination of the stress triaxiality at the crack
tip and the hydrogen content causes SSC development. This basic description of
the fracture mode has inspired the main types of corrosion tests used to evaluate
the sensitivity of a steel to SSC (tests that can be found in the NACE TM0177 and
EFC 16 documents): smooth (cylindrical or prismatic test piece) or notched (DCB -
Double Cantilever Beam) corrosion test specimens are used to determine the ability
of a material to withstand SSC crack initiation and the minimum stress intensity
factor required to propagate SSC respectively.

Figure 10.1. Example of SSC on a cylindrical uniaxial tensile specimen in contact with
a saturated aqueous solution of hydrogen sulfide

(source: Vallourec)

Regardless of the fracture mode, the environmental conditions entailing the highest
risks are those that combine the highest H2S partial pressures and the lowest pH
values (aqueous medium acidities); these parameters establish four severity regions
(Figure 10.2). The tests are usually carried out at room temperature, which is an
unfavorable condition to the material resistance to SSC.

Figure 10.2. Regions of increasing environmental severity from 0 to 3 [NAC 15]

10.2.1.1.2 Carbon and low-alloy steels, Sour Service

Steel grades and manufacturing processes have been specifically developed to


ensure high resistance to SSC of carbon and low-alloy steels; these materials are
known as Sour Service and combine a quenched and tempered microstructure,
whether martensitic or bainitic, a very low content of residual elements and con-
trolled additions of carbide-forming elements (e.g. Cr, Mo).

The most critical factor regarding resistance to SSC is undoubtedly the mechanical
properties: the higher the yield strength and hardness of a steel, the higher its
sensitivity to SSC. The need to extract oil and gas from increasingly deeper deposits
implies the use of steel grades with increased mechanical properties which can reach
up to 1 GPa of yield strength. This induces the development of new Sour Service steel
grades that are more resistant to SSC. As a consequence, the evaluation conditions
of resistance to SSC are required to be increasingly similar to the actual operating
conditions and lead to the use of Fit-For-Purpose (FFP) solutions, with higher pH
and lower H2S partial pressure.

10.2.1.2 Nickel-based alloys with precipitation hardening


These Ni-based alloys, with Cr contents of 20% or more and Mo, are hardened by
the precipitation of ’ (Ni3Ti) and/or ” (Ni3Nb) phases, which provide them with
high mechanical properties (yield strength greater than 950 MPa). They have been
developed based on alloy 718, for which it has been shown that the microstructures
obtained after aeronautical-type thermal treatments (favorable to creep resistance
at high temperatures) are sensitive to hydrogen-induced stress cracking (HISC).
Indeed, in operating components, cracks have been observed at the screw thread
[HUI 03] (highest strained areas), associated with hydrogen embrittlement (hydro-
gen produced for example by galvanic coupling with carbon steel). The intergranular
fractures obtained on alloy 718 have shown that the presence of phases, such as
phase Ni3Nb, at the grain boundaries is a determining factor regarding sensitivity
to HISC.

Intergranular cracking mechanisms have been well studied on pure Ni, highlighting
the effect of the grain boundary nature (low or high disorientation) on hydrogen
diffusivity and solubility [OUD 14]. When hydrogen is present, the HELP (hydrogen
enhanced localized plasticity) mechanism has been experimentally demonstrated as
well as the formation of microcavities located at the grain boundaries. The contribu-
tion of the HEDE (hydrogen enhanced decohesion) mechanism is suspected but has
not been experimentally proven. Depending on the microstructure and, in particular,
on the precipitate grain boundaries’ coverage, intra or intergranular fractures can be
observed on alloy 718. Moreover, there is strong evidence of hydrogen-dislocation
interactions in the case of intragranular fractures [ZHA 16].

Considering the challenges associated with these materials, which are mechanically
resistant and resistant to corrosion, standards have been proposed that enable
the elimination of materials that are not sufficiently resistant to HISC based
on microstructural examinations and taking into account the relationship between
HISC and intergranular precipitation. However, it appears that this approach is
insufficient since embrittlement mechanisms are still poorly established and may
differ from one grade to another [DEM 17].

> Read full chapter

Environmentally assisted cracking


(EAC) in nuclear reactor systems and
components
K. Wolski, in Nuclear Corrosion Science and Engineering, 2012

5.4.3 General equation for stress assisted diffusion


The formalism presented below comes mainly from the field of hydrogen induced
cracking and is of interest for low temperature EAC, i.e. low enough for hydrogen
to concentrate ahead of the crack tip. The aggressive environment is considered
here as a source of embrittling species, typically hydrogen from a liquid phase.
Once adsorbed, hydrogen must dissociate at a surface in order to be absorbed.
Its distribution within the material depends on its actual concentration and stress
gradients and is described by the stress assisted diffusion equation, which can be
derived for any impurity, starting from the definition of the chemical potential μ in
a dilute solid solution (Magnin and Combrade, 2000):

[5.1]

where μ0(T) is a standard chemical potential, C is impurity concentration, VM is


impurity molar volume and p = Tr( )/3 is a hydrostatic pressure.

This pressure decreases the chemical potential of impurity atoms for tensile stresses.
The first Fick’s law allows to express the flux J as a function of the gradient of chemical
potential, where D is the diffusion coefficient, R is the gaz constant and T is the
absolute temperature:

[5.2]

The variation of concentration can be related to the flux gradient via the following
equation ∂C/∂t = –ΔJ and combined with the flux equation which leads to the stress
assisted diffusion equation in its widely used form:

[5.3]
This equation allows the impurity distribution to be calculated as a function of the
concentration and stress gradients.

> Read full chapter

The effect of H2S on the corrosion of


steels
C. Duret-Thual, in Understanding Biocorrosion, 2014

16.2.2 Hydrogen embrittlement: influence of microstructure


In the case of carbon and low alloy steels, two main forms of hydrogen embrit-
tlement can be distinguished: hydrogen induced cracking (HIC) and hydrogen
assisted cracking (HAC). In H2S containing solutions, HAC is referred to as sulphide
stress cracking (SSC), an acronym traditionally used in the international standards
dedicated to the oil and gas industry. Typical standards are EFC 16 (2009), NACE
TM0284 (2011) for HIC resistance and NACE TM0177 (2005) for SSC resistance.

When hydrogen is absorbed in the material, it is accepted that it is present in the


metal lattice as protons (the electron left is in the electron cloud) found in interstitial
positions. Due to charge effects, strong interactions take place between the iron
lattice and the protons as evidenced by the proton induced lattice distortion, higher
than the one induced by interstitial carbon atoms (Lunarska and Zielinski, 1985).

The diffusion coefficient of hydrogen in the iron bcc (body centred cubic) structure
is high when compared to fcc (face centred cubic) structures as shown in Table 16.1.
Conversely, the hydrogen solubility is low in bcc structures versus fcc ones (Table
16.1) (Fallahmohammadi et al. 2013; San Marchi et al., 2007). Consequently, in bcc
structures, hydrogen will tend to segregate or to be trapped in specific sites with a
higher binding energy than in the lattice. These traps are for example, dislocations,
foreign elements present in the lattice, grain boundaries, interfaces between bcc
matrix and precipitates. The trapped hydrogen may be susceptible to moving from
traps to lattice (reversible traps) or it can be stuck (irreversible traps). Traps will
gradually release hydrogen as steel temperature is steadily increased, the highest
binding energy would indeed require the higher temperature. The total hydrogen
concentration in steels can be measured by extraction of hydrogen from the speci-
men heated at high temperature up to the melting point. With thermal desorption
allowing the measurement of released hydrogen as a function of temperature, the
different kinds of traps which are present can be quantitatively evaluated.
Table 16.1. Diffusion coefficients and solubility of hydrogen in iron-based materials
with bcc, fcc and martensitic structures at room temperature. These coefficients
are affected by the presence of traps (apparent diffusion coefficient as compared
to diffusion coefficient restricted to diffusion in the lattice)

Phase Diffusion coefficient (cm- Solubility (molH/m3) Typical


2/sec.) Typical value (apparent, value (apparent, lattice + traps)
lattice + traps) –25 °C –25 °C
Bcc (ferrite) 10−5–10−6 0.1 to 0.5
Fcc (austenite) 10−9–10−11 300
Martensite 10−7 –

Specific effects such as embrittlement and cracking will then take place according to
the distribution of hydrogen in the microstructure, with or without applied stress.

HIC develops without applied stresses. Cracking results both from the stresses
induced by the internal pressure due to the recombination of hydrogen on specific
sites in the microstructure, and from the hydrogen embrittling effect, weakening the
cohesive forces within the metal. For example, a banded structure in ferrite–pearlite
carbon steels favours cracking along the interface between ferrite and pearlite as
shown in Figure 16.1. Non-metallic inclusions such as sulphides and hard phases
have also been identified as crack initiation sites.

Figure 16.1. Cross section of a carbon steel (ferrite–pearlite) specimen showing


hydrogen induced cracking (HIC) damage.

SSC takes place when stresses (residual + external) are present with the occurrence
of hydrogen/dislocation interactions. In addition, hydrogen will tend to segregate
to plastically strained areas which will favour crack propagation due to embrittling
effects located at the crack tip (Figure 16.2). An example of SSC cracking is given in
Figure 16.3, initiated from a pit.
Figure 16.2. Sketch of hydrogen interactions with dislocations at the crack tip.

Figure 16.3. Cross section of a carbon steel (martensitic) specimen showing SSC
damage initiated from a pit.

The consequences of these hydrogen specific interactions with the microstructure


are the following:

• homogeneous structures close to equilibrium (minimum internal stresses) will


be more resistant than the others
• in general, hard materials will be more sensitive to HIC and SSC

• the carbon steels used in the oil and gas industry have generally low hard-
ness to decrease their HIC and SCC susceptibility (recommendation to use
carbon steels of hardness less than 22 HRC, Rockwell C hardness scale, NACE
MR0175-ISO 15156 standard (2003)).

> Read full chapter


UNUSUAL CASES OF WELD-ASSOCI-
ATED CRACKING EXPERIENCED IN
A HIGH TEMPERATURE CATALYST
REDUCTION REACTOR
M.L. HOLLAND, in Failure Analysis Case Studies II, 2001

3.3 Discussion
All three of the cracks examined above are typical of heat-affected-zone cracking
which is also referred to as “hydrogen-induced cracking”, “weld cracking”, “delayed
cracking” or “underbead cracking”.

Cracks in the HAZ are usually sited either at the weld toe, the weld root, or in an
underbead position. The interaction between the factors responsible for cracking
and the ways in which control over them may be achieved are discussed below [4].

3.3.1 Hydrogen level.


During welding, hydrogen is absorbed by the weld pool from the arc atmosphere.
During cooling, much of this hydrogen escapes from the solidified bead by diffusion
but some also diffuses into the HAZ and the parent metal. The amount which does
so depends on several factors such as the original amount absorbed, the size of the
weld, the decreasing solubility, and the time-temperature conditions of cooling.

In general the more hydrogen present in the metal the greater the risk of cracking.
Control over this hydrogen level may be achieved either by minimising the amount
initially absorbed or by ensuring that sufficient is allowed to escape by diffusion
before the weld cools. Frequently a combination of both measures provides the best
practical solution.

3.3.2 Stress level.


Stresses are developed by thermal contraction of the cooling weld and these stresses
must be accommodated by strain in the weld metal. The presence of the hydrogen
appears to lower the stress level at which cracking will occur. In rigid structures the
natural contraction stresses are intensified because of the restraint imposed on them
by the different parts of the joint. These stresses will be concentrated at the toe and
root of the weld and also at notches constituted by inclusions and other defects. The
higher degrees of strain which result produce higher risks of cracking for a given
microstructural hardness.
The stress acting upon a weld is a function of weld size, joint geometry, fitup, external
restraint, and the yield strengths of the plate and weld metal.

3.3.3 Microstructure.
The heat affected zone (HAZ) of the parent metal adjacent to the weld is raised to
a high temperature during welding and subsequent rapid cooling (quenching) by
the surrounding parent metal causes hardening by transformation to martensite.
Close to the fusion boundary the HAZ is raised to a sufficiently high temperature
to produce a coarse grain size. This high temperature region, because of its coarse
grain size is not only more hardenable but also less ductile than regions further from
the fusion boundary. It is thus the region in which the greatest risk of cracking exists.
As a general rule, for both carbon-manganese and low alloy steels, the harder the
microstructure the greater is the risk of cracking. Soft microstructures can tolerate
more hydrogen than hard before cracking occurs.

3.3.4 Temperature.
Hydrogen embrittlement of ferritic steels occurs only at low temperatures, close
to ambient. It is therefore possible to avoid cracking in a hard, i.e. susceptible,
microstructure by maintaining it at a sufficiently high temperature, either until
hydrogen has diffused away or until the microstructure is softened by tempering,
to render it less susceptible. This principle is employed in multipass welding and in
post-weld heat treatments.

An increase in temperature increases the rate of diffusion of hydrogen and thus


accelerates its removal from the weld. Any measure which slows down the weld
cooling rate is therefore helpful in reducing the hydrogen level. Preheat, for example,
by slowing the cooling rate, not only softens the microstructure but also helps
hydrogen to escape. As a result, higher hardness levels can be tolerated without
cracking than if preheat had not been used.

For welds in those steels with hardenability so high that soft microstructures cannot
be produced at all, and where preheat cannot remove sufficient hydrogen, (such as
the Cr-Mo steels) a weld interpass temperature, or a post-weld heating temperature,
high enough to avoid cracking must be held for a sufficiently long time to allow
hydrogen to diffuse away before the weld cools.

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like