You are on page 1of 64

Review of type IV cracking of weldments in 9

12%Cr creep strength enhanced ferritic steels


D. J. Abson and J. S. Rothwell*
TWI Ltd
International Materials Reviews, November 2013, 58(8), 437-473.

Abstract
The improvement of thermal efficiency of power plants has provided the incentive for the
development of the martensiticferritic 912%Cr creep-resistant steels. Good progress has been
made in developing such steels, which are being used particularly in the wrought form as tubes
and pipes for fossil fuelled power stations. They are also finding use in high temperature process
plant within the oil and gas sector, and are being considered for use in generation IV nuclear
designs. The high temperature conditions that these steels operate under in fossil fuelled power
stations induce type IV cracking. This type of cracking occurs in the intercritical or fine grain region
of the heated affected zone via a creep mechanism, and results in fractures with relatively little
total cross-weld strain. Despite the occurrence of type IV cracking experienced in lower alloy
predecessors, successor alloys have been introduced and widely used with insufficient
consideration given to the consequences of welding them. Unfortunately, the newer steels suffer
from reduced cross-weld creep strength due to type IV cracking to a greater degree in the
temperature range of operation expected of them, and thus many failures by this mechanism have
occurred. The subject of type IV cracking has been an area of active research interest. This review
aims to serve as an update, drawing selectively on some of the vast amount of literature that has
been published over the last 30 years.

Introduction
Improved thermal efficiency of power plant has been the main driver for the development of
ferriticmartensitic 9 12%Cr creep-resistant steels that are also commonly known as creep
strength enhanced ferritic (CSEF) steels. The target operating temperature for these steels is
650uC, with a common target design life of 100 000 h. Increasingly, the demand for efficiency is
linked to efforts to reduce CO2 emissions, in order to meet environmental obligations and minimise
any form of punitive environ- mental tax. Figure 1 illustrates how CO2 emissions vary with thermal
efficiency. 1 Government agencies such as the Department of Energy in the USA are funding large,
ambitious projects for the development of ultra- supercritical (USC) power plant designed to
operate at 760 C and conversions from supercritical to USC opera- tions at 700 C have been
carried out in Japan.2Retrofitting of existing plant is a popular option for improving output with
minimal investment, since existing fittings can be used. The inclusion of CO2 capture and storage
technology for fossil fuel power plants represents a significant advance in terms of combating CO2.
However, it also reduces overall output efficiency and therefore there is an additional incentive for
raising operating temperatures and pressures in order to maintain the same amount of saleable
energy per power station.
An example of the materials used in these state-of-the- art projects is illustrated in Fig. 2. This
Alstom showpiece demonstrates achievements in dissimilar metal joining necessary for minimising
capital costs.3 Alloy develop- ment programmes that involve the collaboration of many partners
from industry and academia feed into the high temperature plant initiatives and an example of this
is the European Co-operation in Science and Technology (COST) programme that has developed
alloys for use in the AD 700 initiative and other state-of-the-art fossil power plant such as the
Neurath plant to be constructed in Germany.4 Many of these technically challenging new power
plants incorporate the latest ferritic alloy developments and an increasing amount of austenitic and
nickel- based materials, while more conservative constructions rely on established, coded steels
that have been available for some time, such as grades 91 and 92 that are popular pressure vessel
materials, and that represent the current highest performing types in the 9%Cr family of steels. In
the developing economies, such as that of China, where construction of new power plant is very
rapid, the demand for these kinds of materials has recently been beyond the rate of supply.5 This
huge demand for such materials continues, despite significant problems experienced with them
during fabrication and service.

1 Plot showing the expected reduction in CO2 emissions by increased plant efciency per unit energy1

Table 1 and Fig. 3 illustrate how, over the years, good progress has been made in developing
creep-resistant steels which are being used particularly in the forged form as tubes and pipes in
boilers, but which are also being considered for rotors and their cast variants for turbine casings.
Such steels are also being considered for the new breed of nuclear reactors (generation IV), and
beyond the power industry, the oil refinery industry has become more interested in this family
of steels for pressure vessels and piping. The improvement of tensile and creep rupture strength
is attractive to designers who want to take advantage of the higher hot tensile and creep strength
to reduce pipe wall thicknesses, and thereby minimise thermal stresses for a more reliable plant.
This is increasingly important for todays power plants, which are subject to temperature cycles in
an effort to respond to the peaks and troughs of demand, and improve profitability. A schematic
illustration of the reduction in pipe wall thickness possible through the use of more advanced
materials is presented in Fig. 4.
Much less attention was paid initially to cross-weld testing of the 9%Cr steels, not without quite
discomfort- ing consequences for the power industry. Also, incorrect heat treatment of grade 91
steel in particular and an initially detrimental allowable aluminium level have contributed to poor
creep performance. However, the most significant, weld-related problem that has plagued the
power industry for over 40 years and that still has not been overcome, despite alloy development,
has been the phenomenon of type IV cracking. This is creep failure that occurs in the heated
affected zone (HAZ) of welded joints, particularly in high alloy precipitation- strengthened grades
that are popular today, such as grades 91 and 92. These grades have a tempered martensite
microstructure that is solid solution strength- ened, but with much of the strength derived from
fine
[ Zoom ]

2 Demonstration showpiece illustrating the materials and joining techniques that are under consideration for
the new USC plants3

Table 1 Development of 912%Cr steels67,178

600 C/105 h
creep
Example
rupture
alloys
strength/MP
a

Maximum
metal use
temperature/
C

Years

Alloy
modification

1960
1970

Addition of Mo
or Nb, V to
60
simple 12Cr
and 9Cr steels

EM12, HCM9M,
HT9, Tempaloy 565
F9, HT 91

1970
1985

Optimisation of
100
C, Nb, V

HCM12, T91,
HCM2S

1985
1995

Partial
substitution of 130
W for Mo

P92, P122,
P911 (NF 616, 620
HCM12A)

Increase W and
Emergin addition of Co,
150
g
B and
controlled N

NF12, SAVE
12, MARN,
MARB2,
MARBN

593

650

[ Zoom ]

3 Chart showing the progressive development of 912%Cr steels64,67

MX precipitates such as vanadium nitride (VN). The composition of popular creep-resistant steels is
given in Table 2.162,163
Safety is, of course, of primary concern in the design and manufacture of plant intended for steam
service. Improvements in design, weld integrity and inspection techniques have all contributed to
improved safety, with the knowledge gained being incorporated into relevant codes and standards.
An increased understanding of the role of defects in failure has been reflected in assessment
codes, with an improved ability to detect and size defects
[ Zoom ]

4 Illustration showing possible reduction in wall thickness achievable through use of advanced 9%Cr parent
alloys: temperature5600uC; pressure530 N mm22; 100 000 h life94

contributing to improved safety arising from their use. An understanding of the phenomenon of
type IV cracking and of the extent of long term creep strength degradation, and strategies for
improved performance, would constitute a useful addition to existing knowl- edge, and would help
designers and operators to take the right decisions when designing, operating or repairing plant, to
maximise overall safety, operating efficiency and plant availability.
In this document, the phenomenon of type IV cracking, particularly with respect to weldments in
9 12%Cr steels, has been reviewed, and associated technology gaps that exist in current
materials and practices have been discussed.

Type IV cracking characteristics and causes


Introduction
The problem of type IV cracking has existed for many years in power plant operated at elevated
temperatures. As early as 1974, in tests carried out at 550 C, Schuller in Germany reported the
low cross-weld strength of weldments in DIN 10CrMo9-10 (equivalent to AISI A182 grade F22
225%Cr1%Mo).6 He observed that, in tests extending to more than 104 h, the cross-weld
strength fell more than 20% below that of the parent steel. In the UK, many power stations built
in the 1960s and 1970s incorporated a large amount of 0 5%Cr0 5%Mo0 25%V steel (BS
3604 grade 660) that allowed a substantial rise in the operating temperature (from around 540C
up to 565C) for coal- fired plant.7 Problems with fabrication cracking were followed by long-term
failure mechanisms such as type IV cracking. Meanwhile, continental Europe deployed the 12%Cr
steel X20CrMoV121 (X20), with modest service temperatures in thinner section sizes, and
encountered few problems, thanks to an inherent resistance to reheat cracking and a resistance to
type IV cracking at the lower service temperatures.7 Grade 91, provided a cheaper and stronger
alternative to X20 that could be deployed in higher temperature plant with target temperatures of
600C. Unfortunately, at these higher temperatures, the resistance of both X20, grade 91, and of
subsequently developed creep-resistant steels to type IV cracking, falls in a more dramatic way
than for their lower alloy predecessors. Early determinations of the ratio of cross-weld creep
strength to that of the parent steel revealed values substantially below unity. Townley gave a 10
000 h/570C value of 065, and Etienne and Heerings gave values at 550, 600 and 650C,
determined by extrapolation to the design stresses of 10, 068 and 070 respectively.8,9 (These
authors also listed values for many other materials in elevated temperature service, including
X20.) A warning note was sounded by Bru hl et al., who measured cross-weld creep strengths
that were substantially lower than parent steel values.10,11 This and the following warning note of
Middleton appear to have been largely to no avail: The operation of super- critical plant
constructed from either grade 91 or X20 steels at y600uC would therefore involve a degree of risk
of type IV cracking, the level depending on the medium to long term strength loss found for crossweld testing, unless a modest overdesign were applied.12 From a comparison of known creep
rupture strengths and observed failure times of power plant welds, Middleton also deduced the
likely level of system stresses. He commented that At 600uC the superior rupture strength of P91
renders it more resistant to type IV failures than X20 beyond y30 kh. The presence of a creepweak HAZ leads to a cross weld rupture strength that is severely reduced, however, leading to a
much lower tolerance of system stress, to the extent that P91 systems operated at 600uC are
predicted not to sustain system stresses > 16 MPa without risk of failure for a 40% strength loss.
Bru hl et al.10,11 also noted that, while weld metal, HAZ and parent steel are subjected to the
same strain for circumferential welds, the lower HAZ creep strength of P91 weldments is of
concern for longitudinal welds, as they are subjected to the same stress as the parent steel (which
is twice the cross-weld pressure stress to which the girth welds are subjected).
Almost all testing in alloy development programmes is for parent steels, and extensive testing
programmes have been used to optimise parent steel compositions. 13 Any cross-weld testing
appears commonly to be an after- thought, although the presence of a creep-weak HAZ in a 9%Cr
steel was reported, as discussed above, and demonstrated on simulated HAZs as long ago as 1990
by Middleton and Metcalfe, who carried out creep rupture testing on grade 91 samples that had
been heated to temperatures between Ac1 and Ac3 (inter- critical and austenitic phase fields); see
Fig. 5.12 Since then, similar studies have been carried out highlighting the poor creep strength of
material heated above the Ac1 temperature; see Figs. 6 and 7. A growing body of data, reviewed
in 1993 by Etienne and Heerings, revealed the lower creep strength of weldments, with the ratio of

cross-weld creep rupture strength to that of the parent steel falling in the range from 05 to
unity.9 However, as discussed in the section on Design codes, published weld strength factors and
extrapolation of data, with a few exceptions, such information has been slow to find its way into
design codes.
Interest in type IV cracking has intensified in recent years. Service experience with grade 91 steel
components has revealed many early failures in the HAZ. 14,15 The need to assess HAZ creep
rupture strength has been recognised in Japan where extensive cross-weld testing or the
imposition of simulated HAZ heat treatments, followed by creep rupture testing, have been carried
out. European institutes that have particular interest in the problem of type IV cracking include the
University of
Table 2 Chemical composition of new 912%Cr steels and experimental European steel
FB2179

Element

Grade
91

NF616
(grade
92)

HCM12A TB 12M

Grade
911

Grade
122

FB2

0.080.12

0.070.13

0.070.14

0.100.15

0.090.13

0.070.14

0.13

Mn

0.200.60

0.300.60

0.70

0.400.60

0.300.60

0.70

0.82

Si

0.200.50

0.50

0.50

0.50

0.100.50

0.50

...

0.010 0.010

0.010 0.010 0.010 0.010 ...

0.020 0.020

0.020 0.020 0.020 0.020 ...

Cr

8.009.50

8.509.50

10.0012.50

11.011.30

8.5010.50

10.012.5

9.32

Mo

0.851.05

0.300.60

0.250.60

0.400.60

0.901.10

0.250.60

1.47

...

1.502.00

1.502.50

1.601.90

0.901.10

1.502.50

...

Co

...

...

...

...

...

...

0.96

Ni

0.40

0.40

0.50

0.701.0

0.40

0.50

0.16

Cu

...

...

0.301.70

...

...

0.301.70

...

0.18-

0.15-

0.15-

0.15-

0.18-

0.15-

0.20

0.25

0.25

0.30

0.25

0.25

0.30

Nb

0.060.10

0.400.09

0.090.10

0.040.09

0.060.10

0.040.10

0.030- 0.0300.070 0.070

0.0400.100

0.040.09

0.040- 0.0400.090 0.100

Al

0.04

0.040

0.040 0.010 0.04

0.04

...

0.0010.006

0.005 ...

0.00050.0083
0.005

Sn

...

...

...

0.010 ...

...

...

As

...

...

...

0.010 ...

...

...

Sb

...

...

...

0.005 ...

...

...

(115)*

(115)

[150
(105
h)]

(115)

...

...

(60)

[80
(105
h)]

(65)

...

105 h
creep

600uC 94

rupture
strength 650uC 50
/MPa

(60)*

...

0.05
0.019
...

(): estimated; ...: not specified.


*Robertson and Holdsworth give 113 MPa for 100 000 h at 600C and 56 MPa for 100 000 h at
650C.180

5 Inuence of a brief hold at different austenitising temperatures, plus subsequent PWHT at 570 or 600C on
creep rup- ture life for grade 91 steel12

6 Plot of creep rupture stress displaying data for P91 parent material, simulated ne-grained heat affected
zone and welded joints: all specimens were subject to a 760C/2 h PWHT and were tested at a temperature of
600C53

Nottingham, the University of Loughborough, Imperial College, MPA Stuttgart and the Technical
University of Graz.
The poor creep strength of welded joints greatly undermines the advantages gained through
alloy developments, and has resulted in many unexpected repairs, several large-scale failures and a
subsequent loss of confidence in the new materials. Particularly in the UK, existing plants into
which grade 91 steel has been introduced in replacement components, commonly headers, or
power plant constructed in recent decades may contain under-designed components that need
careful monitoring and/or repair. It is only recently that valuable long-term data are starting to be
incorporated into design codes to help prevent premature failures. Initiatives such as the European
Creep Collaborative Committee (ECCC) and the state-funded analysis of data carried out at the
National Institute of Material Science (NIMS) in Japan have been set up to collect data for design
purposes. However, although reliable perfor- mance may now be designed in more easily by way of
increased thickness of material, the poor creep strength of welds that leads to type IV cracking
remains, and prevents the impressive properties of the parent alloys being taken advantage of to
their full extent.

Failure location
The classification of service cracking in weldments devised by Schu ller et al., which describes the
location of cracking relative to the weld, is still used today and is shown in Fig. 8. 6,16 Types IIII
are associated with the fabrication of a joint and may be solidification, hydrogen, reheat,
temper embrittlement and occasion- ally long-term creep-related cracks. Type IV cracking forms
towards the outer edge of the visible HAZ, beside the parent metal (Fig. 9), and is exclusively a
creep cracking mechanism that occurs after long durations.
To explain the location of type IV cracks more precisely in relation to the HAZ microstructure, it is
useful to illustrate the different subzones that commonly exist in the HAZ. Figure 10 shows a
section of the phase diagram for grade 91 steel, and how different parts of

7 Rupture times for HAZ simulated coupons for ASME grade 122 steel: open points are furnace heat treated,
and closed points represent coupons treated in a weld simulator108

8 Classication of cracking in weldments from Brear and Fleming,16 according to Schu ller et al.6

the phase diagram relate to the different microstructural regions of the HAZ. 17 While it is well
established that type IV cracking occurs towards the outer edge of the visible HAZ, the exact
location, in terms of micro- structure has been reported to occur in both the intercritically (IC)
heated HAZ and the grain-refined or fine grain (FG) HAZ. Thus, there is not a single HAZ
microstructural region associated with type IV cracking.

9 Type IV cracking in electron beam welded 9%Cr1%Mo steel

a macro; b micro image. mm scale shown


The distinction between the two zones is not a simple one to observe in practice. Figure 11 shows
the HAZ microstructure of a recently developed boron-containing 9%Cr steel (FB2), and illustrates
the difficulty of distinguishing the different parts of the HAZ. 18 Different materials, welding thermal
cycles and tem- peraturestress regimes may bring about differences in the observed failure
location, although the failure location in long-term (low stress) service is most commonly reported
to be the fine-grained heat affected zone (FGHAZ), if stresses transverse to the weld are the most
significant. It is worth noting that, in multiple pass welds, the fine-grained and intercritical HAZ
regions are essentially almost continuous through thickness, unlike the coarse-grained heat
affected zone (CGHAZ), which may exist in disconnected regions beside the fusion boundary. This
provides a continuous region that is susceptible to the phenomenon. It should be noted that the
number density of cavities is not uniform across the section. The highest density of cavities is
expected to occur at midsection, as observed by Yaguchi et al., where the triaxiality is
highest.19 This observation clearly has implications for the use of surface replicas to investigate
the extent of creep cavitation.
Type IIIa is another long-term creep failure mechan- ism, associated with carbon migration from
the CGHAZ near the fusion boundary over the lifetime of the weldment, resulting in a carbidedepleted zone with lowered creep strength that subsequently leads to cracking. 20 This type of
cracking is associated with differences in the carbon activity of the weld and HAZ, which is often
due to differences in the Cr concentration. It tends to be more prevalent in butt welds where less
inspection is focused, and consequently more time is available for diffusion and development of

cracks at typically longer durations than type IV cracks. The most commonly associated
combination of materials affected
[ Zoom ]

10 Schematic representation of the subzones of the HAZ corresponding to the calculated phase diagram of
X10CrMoVNb9-1 (grade 91 type) steel.17

The HAZ associated with a single weld bead is shown. Type IV cracking is associated
with the intercritical zone or the adjacent fine grain region
by type IIIa cracking are 1/2Cr1/2MoV steel welded with a 2Cr1Mo weld metal. However,
denuded CGHAZs are of much concern for 9Cr joints involving lower alloy bainitic or higher alloyed
martensitic and austenitic steels.21 As discussed in the section on Technology gaps and future
trends, EPRI and Metrode Products Limited have developed a filler for the avoidance of this
problem which has a lower carbon activity, and is especially useful for welding austenitic steels to
ferritic or martensitic steels.

Circumstances giving rise to type IV cracking


Service failure times for type IV cracking have been reported to be typically 610 years, but have
been as short as 35 000 h.7,15 While much of the earliest type IV cracking detected occurred in
heats with low N/Al ratios (15), many examples were found for which cracking was observed after
58 000 h service in heats with much higher N/Al ratios (some > 10, at nitrogen levels > 005%).
Consequently, Brett commented that Because
the cracking observed to date has occurred at such an early stage, there is unfortunately ample
scope for this process to continue with even material with quite high N/Al ratio potentially cracking
within the design life (typically 150,000 hours). Whether or not type IV cracking will occur is
influenced by the microstructure developed by the quality heat treatment, the operating
temperature and most importantly the stress state of the component. For transverse tests, the
mechanism is favoured at higher temperatures and lower stress levels. At higher stresses and
lower temperatures, failures are more common in the weld metal or parent metal. As the
temperature is increased and the stress reduced, there comes a point at which type IV creep
cracking becomes the dominant mechanism, as illustrated schematically in Fig. 12 and for
experimental steel FB2 in Fig. 13. This implies that, where the primary stress is across the weld,
below a critical or threshold stress level, the creep strain in a weldment is localised, and creep
failure occurs in the type IV region of the HAZ much earlier than would be
[ Zoom ]

11 Heat affected zone region of a ruptured creep cross-weld specimen that failed by the type IV mechanism.

The specimen was extracted from a flux cored arc weld made in a piece of FB2 rotor
steel. The test conditions were 625C at 110 MPa. 18 The rupture time was 9664 h
a HAZ region on non-ruptured side. Etched in 1 : 5 v/v Hcl (35%) and distilled water with 051 g
K2S2O5, 2 g NH4HF2 per 100 g of solution; b HAZ region beside rupture with an attempt at defining
the different regions of the HAZ. Etched in 25% picric 25% Hcl v/v in ethanol

12 Schematic representation of the reduction in creep rupture strength beyond the onset of type IV cracking
for weldments at a xed temperature.

The expected fracture location for a cross-weld specimen is indicated. Type IV cracking
occurs much earlier than for a parent or weld metal specimen under the same test
conditions
expected for the parent steel. Figure 14 shows the dependence of creep life on temperature for
weldments in ASTM/ASME grade 91 steel, with a line distinguishing the boundary conditions for
type IV cracking. An approximate equation for this boundary is

log10(tf )=00235|(733-T) (1)


where log10(tf) is the logarith to base 10 of the time to rupture in hours, and T is the temperature
in C.
Similar behaviour is observed in girth welds subject to high system stresses that increase the axial
loading, and seam welds subject to pressure loading with hoop stress as the main component. 12 It
was noted in the section on Introduction that longitudinally welded pipe is not specified in USA
and most European countries, due to historic failures of pipes that resulted in fatalities. However,
longitudinal seams in headers and other pressure vessels are inevitable.
From their experimental measurements and subsequent computation of the creep of an X20 vessel
loaded at 610C with an internal pressure of 200 bar, Steen et al. concluded that the normally
expected failure mode in a pipe under pure internal pressure will be HAZ cracking perpendicular to
the hoop direction.22 The lifetime of the weld joint will fall below the lower bound of the base
material scatter-band, although when no weld efficiency factor is adopted, the time safety margin
may be very small.
Coussement and de Witte investigated these effects in X20 and P91 pressurised vessel tests, and
found that
[ Zoom ]

The experimentally observed dependence of type IV cracking on temperature and stress for weldments made
by various processes in FB2 steel.18

The points bounding the darker shaded red area to the upper left represent type IV
failures, i.e. fractures located mainly in the outer HAZ region and featuring limited
ductility; those bounding the grey area represent failures located in the parent or weld
metal, with higher ductility

14 Failure locations for cross-weld samples of grade 91 steel for different rupture lives and temperatures
describing the temperature dependence of type IV cracking7

poor placement of welds, e.g. on tight radii of end caps, resulted in reduced failure times due to
increased axial stress across the welds.23 Girth welds that were subjected to pure pressure failed
due to transverse cracking of the weld metal or parent material rather than the HAZ region, and
the importance of creep ductility in such situations was highlighted, i.e. the capacity to strain,
offload and redistribute stresses over the weld. Evidence of the importance of creep ductility was
highlighted when comparing girth welds in X20 made with matching and high Ni fillers. The
matching filler failed by axial cracking in the weld metal, while the higher ductility Nibased filler
weld failed in the parent, again by axial cracking, but with a 54% improvement in time to rupture;
cracking did not occur in the HAZ. Similar conclusions were drawn by Smith et al., who reported
that if welded components are subjected to a loading condition where bending predominates, and
type IV creep strain accumulates in a localised region, failure through discrete crack growth will
prevail, whereas with less constrained loading, longer plant lifetimes will result, with the tendency
for net section creep strain to accumulate in the type IV zone. 24
For welds under pure pressure, the principal stress for the seam welds, which is in the hoop
direction, is twice that to which the girth welds are subjected. Moreover, the failure of a seam weld
is likely to have much more serious consequences than the failure of a girth weld. However, the
superposition of system stresses can and does allow type IV failure to occur in the HAZs of girth
welds, for example those attaching header end caps. Multiaxial constraint is expected to increase
with specimen size and improve creep rupture times. It follows that high strength, low ductility
weld metals and high angle fusion lines should increase constraint and creep rupture times for
transversely loaded crossweld specimens. These effects have been demonstrated by Masuyama,
who considered the effects of specimen size, shape and loading configuration in high chromium
martensitic steels.25 It was reported that double penetration butt welds, X type, perform worse
than U groove geometries for large-scale (30 mm2) uniaxial cross-weld specimens, and when
tested in hot pressure vessels (seam welds subject to hoop stresses) for grade 122 steel at 650C.
The results presented also showed a large reduction in time to rupture for seam welds compared
with the large uniaxial specimens, although no explanation for this was offered. In contrast,
Yaguchi et al. observed similar creep rupture lives for conventional creep rupture specimens and
internally pressurised welded pipe, illustrating the conflicting findings that can found in the
literature.19 Their study involved narrow-gap longitudinal seam welds in a U-shaped preparation in
9%Cr and 12%Cr steels. Type IV failure was reported as occurring in the FGHAZ. The plane
stress/plain strain finite element modelling of Kimmins and Smith indicated only a small change in
the strain in the type IV region when its inclination to the stress axis, initially perpendicular, was
changed by 1015.26 The influence of weld angle, and thus the inclination of the fusion boundary
to the stress axis (between 0 and 45), on creep life was modelled by Tanner et al., who showed
up to a 60% reduction in a creep life, with the minimum at an inclination of ~25. They also
compared their computed values with the experimental data of Francis et al.. 27,28

Hyde et al. have also used element modelling of biaxial loading on 2 Cr1Mo girth welds in
Cr Mo V pipe to consider several axial (end load) and hoop stress configurations for weld
angles between 0 and 375 for a weld with uniform HAZ properties. 29 The analysis suggested an
interaction between applied stress state, weld angle and time to rupture. At a low axial to hoop
stress ratio, a high weld angle gave improved creep rupture life. A cross-over point, where no
benefit was derived, was identified (at an axial to hoop stress ratio of y0?6), and the trend was
reversed at higher ratios, but with a lesser difference between the predicted creep lives. Further
similar studies have since been made for grade 91 steel, highlighting the effect of stress state. 30

Weld strength factor (WSF) and strength reduction factor


(SRF)
As noted earlier, Bruhl et al. pointed out that the lower HAZ creep strength of P91 is of concern for
longitudinal welds, as they are subjected to the same stress as the parent steel. 10,11 In short-term
laboratory tests, the cross-weld and parent creep strengths are usually very 14 Failure locations
for cross-weld samples of grade 91 steel for different rupture lives and temperatures describing
the temperature dependence of type IV cracking7 Abson and Rothwell Review of type IV cracking
of weldments in 912%Cr creep strength enhanced ferritic steels International Materials Reviews
2013 VOL 58 NO 8 445 similar until the onset of type IV failure. With the onset of type IV failure
comes a degradation of cross-weld creep performance compared with that of the parent steel
under the same conditions. On a plot of creep rupture stress versus duration, an inflection is
observed for cross-weld specimens, as illustrated in Fig. 12. This inflection occurs at a threshold
stress level for a given temperature and material. Below the threshold stress, the difference
between parent and cross-weld creep strength widens, so that a progressively greater discrepancy
exists at lower stresses and longer durations. The difference between the weldment and parent
creep strength at a given duration is expressed in a number of ways, but commonly as the WSF or
the SRF. These are often defined as
WSF(t, T)=Ru(w)/t/T/Ru/t/T (2)
SRF(t, T)=(R/ut/T - Ru(w)/t/T)/Ru/t/T (3)
where Ru/t/T is the creep rupture strength of parent material specimens at time t and temperature
T, and Ru(w)/t/T is the creep rupture strength of cross-weld specimens at time t and temperature
T.31 The terminology for describing the reduced creep strength of weldments compared with that of
the parent steel is sometimes confusing. It is desirable to achieve a high WSF and a low SRF.
Examples of some of the alternative parameters are given in Table 3. Allen et al. identified that
there is a broad (but by no means perfect) correlation between HAZ and parent material creep
strength between heats.32 However, there is also a trend for the WSF to decrease with improved
parent performance between grades, thereby strongly reducing the effective creep strength of a
fabrication compared with a single piece component. In some cases, the improved creep strength
of the parent material has been shown to be completely countered by the detrimental effect of
welding, when compared with lower alloy predecessors. For example, in their in-depth study of
data from parent and cross-weld specimens of the 912%Cr family, Kimura et al. indicated that
employing a higher strength parent steel will bring no further advantage in terms of cross-weld
creep performance at the higher temperature ranges desired for operation.33 They demonstrated
that for grade 92 (which was the strongest of the coded CSEF steels) at 625C, the estimated
105 h cross-weld creep rupture strength is 462 MPa compared with welded joints in grade 91 (that
has a lower parent creep strength) that have a creep rupture strength of 499 MPa for the same
temperature and duration. Such findings have made the further development of creep-resistant
steels less attractive unless susceptibility to type IV cracking is eliminated.

Design codes, published WSFs and extrapolation of data


The relatively poor creep strength of weldments at long durations was not understood, and
generally not fully accounted for by design codes until many years after the introduction of steels
such as grade 91 into the market place. Figure 15 shows creep rupture data for parent and crossweld grade 92 specimens, and illustrates the point that if plant were designed to the allowable
stress specified by ASME alone (before 2008), weldment failure could be expected beyond 20 000
h of service for welded grade 92 material, if the allowable stress for 100 000 h service was
employed.34 Several of the standards that relate to out-ofcore nuclear components (ASME Section
III Code Case N47, the French Code RCC-MR and the Japanese code ETSDG) were reviewed by

Bhoje and Chellapandi, who observed that comprehensive creep SRFs for weldments are provided
in the French code RCC-MR.35 Weldments are discussed in Section 4/5 of the R5 assessment
procedure. In addition to stating that A particular problem with weldments is the presence of
residual welding stresses, there is a short section on type IV cracking in 05 CrMoV weldments,
which discusses crack growth in the type IV region of girth welds, with no explicit mention of low
cross-weld creep strength. The time to rupture of the complete weldment is taken to be the lowest
of the times to rupture for the constituent zones. 36
As noted in the section on Introduction, there was a comparatively early awareness in Germany
and in the UK that the cross-weld creep strength can be substantially lower than that of the parent
steel. This awareness was reflected as early as 1988 in the German code. 37 As noted by Gomes et
al., this code incorporated a factor in the stress of 80% for fully loaded longitudinal seam
welds.38 However, in the light of the data now available for grade 91 steel, it is clear that this
allowance was not sufficiently adequate. Until recently, no factor accounting for poor cross-weld
creep strength was incorporated into more widely used design codes such as the ASME code,
where allowable stress is based upon safety factors for parent material performance. The ASME
code relating to high temperature pressure vessels was, however, updated in 2008 to reflect crossweld performance. More precisely ASME I introduced a range of weld SRFs w used for calculating
component thicknesses.34 The general equation for calculating the thickness of drums and headers
has not changed, but the definition of the efficiency factor E has been expanded to incorporate w
for the seamwelded condition

(4)

where P is the pressure, D is the outside diameter, S is the maximum allowable stress value, E is
the efficiency, y is
Table 3 Some examples of terminology used for WSF or SRF

Description given by
authors

Designation according to
equations (2) and (3)

Weld reduction factor

SRF(t, T)

SRF
WSF(t, T)]
Weldment parent creep
WSF(t, T)]
strength ratio
Weld SRF

Weld creep SRF

WSF(t, T))

Author(s)
Middleton et al.,
20017
ASME B31.3
Abson et al.,
200783
Kimura et al.,
200833
EN 13445-2,
200244

*Weld strength reduction factor (WSRF)50?86cross-weld creep strength/allowable stress 33

15 Creep rupture plot showing the performance of ASTM/ASME grade 92 parent material and weldments.

The allowable stress S at 10 000 h for P92 seamless pipe, and S multiplied by weld SRF
w for N+T (077) and post-weld heat treatment (PWHT) (05) are illustrated. The w
multiplier was introduced into ASME I 2008 code in recognition of the poor cross-weld
creep performance of joints.34 This factor effectively increases the thicknesses required
by the code for seam welded pressure components
the temperature coefficient and C is the minimum allowance for threading and structural stability.
The efficiency factor E for seam-welded components is equal to w, or the ligament efficiency,
whichever is lower. In the case where the weld seam is penetrated by the openings forming the
ligament, E is taken to be the product of w and the ligament efficiency.
The value for w is provided in a table in the code, and is dependent on heat treatment of the
component after welding, the steel type and temperature range of operation. For grade 91 and
similar steels in the PWHT condition, between 510 and 649C, the value of w is 05. Over the
same temperature range, for the same steels in the normalised and tempered condition, the value
falls from 086 to 077. The introduction of w effectively reduces the maximum stress (S6w), as
indicated in Fig. 15, and increases the required thickness for seam-welded components, as
indicated in Fig. 16.
No specific guidance is given in the code for other types of welds, such as girth welds. However, for
girth welds, the cross-weld stress arising from the internal pressure is half that to which a
longitudinal weld is subjected. Moreover, the deformation of a girth weld in response to the
pressure stress is limited by the strain produced in the adjoining parent steel. Hence, girth welds
generally present a substantially lower risk of

16 Schematic showing the calculated pipe thickness according to ASME I 2007 and 2008

highlighting the difference that a weld SRF E makes for seam-welded pipe: P55 MPa,
T5572C, y507, C50, D5864 mm (Ref. 181)

17 Creep rupture WSFs for grade 92,115 grade 122107 and grade 91.68 Values derived from creep rupture
data of crossweld or simulated FGHAZ specimens tested at 650uC. A linear extrapolation up to 300 000 h has
been made

suffering type IV cracking, although sCh welds could be affected adversely by local stress
concentrations and, as discussed by Middleton, by high system stresses.12 A weld joint SRF W was
introdCed into ANSI B313 in 2004 [clause 30235(e)], as discussed by Becht IV.39 With a
general dearth of experimental cross-weld data, it specified linear interpolation between values of
10 at 510C and 05 at 815C, for all materials. Since the 2008 issue (Table 30235), values of
W range from 1 to 05 for different materials and service temperatures. For socalled CSEF steels in
the post-weld heat treated condition, it is 05 for the temperature range from 510 to 649C. For
weldments in the N & T condition, it decreases from 1 to 077 over the same temperature range,
varying linearly with temperature.
Unfortunately, the WSF is worse for the newer high chromium creep-resistant materials at their
target operating temperature range than it was for the older less alloyed steels, as indicated in
Figs. 17 and 18, which shows an extrapolated 100 000 h WSF of y04 for grade 122 (12%Cr) steel
at 650C. Grades 92 and 122 show broadly similar behaviour, namely a stronger downward trend
than that shown by grade 91.40 As noted above, this is supported by data presented more

18 Weld strength factors predicted for various power plant steels after 100 000 h41

recently by Kimura et al.33 Schubert et al.41 show 600C/ 100 000 h WSF values of
approximately 075 for 9%CrMoV steels and 05 for 12%CrMoV steels.41 The graphs show clearly
that the WSF is not constant, but rather decreases with increasing rupture life (decreasing applied
stress), and with increasing temperature, reaching levels well below unity for exposure times of
the order of 100 000 h at temperatures 600C.
This behaviour is also illustrated by the analysis of ECCC cross-weld data for grade 91 and E911
steels carried out by Holmstrom and Auerkari.42 Their plot of stress versus WSF for grade 91
steel at temperatures ranging from 575 to 650C, Fig. 19, shows WSF values falling from ~09 to
~06 as the stress decreases. Data points added to the original graph are the WSFs given

Stress versus WSF for grade 91 steel derived by Holmstrom and Auerkari42 from 2005 ECCC data, to which
the following data points have been added:

sun = 100 000 h WSF values from Schubert et al.,41 which indicate associated strength
values; star = 2009 ECCC 100 000 h strength values, with assumed WSF plotted in
accordance with underlying data
by Schubert et al.,41 which fit the underlying data set well, and 2009 ECCC 100 000 h strength
values, from which, relevant WSF values can be estimated. These added data points reveal clearly
the progressive decrease in WSF at 100 000 h with increasing service temperature. Similar
behaviour is reflected in graphs presented by Laha et al. 43 They showed the SRF (of equation (3)),
derived from data extrapolated to 100 000 h, increasing up to approximately 30% at 600C and to
approximately 45% at 650C. In the light of this decrease in WSF, it is clear that values of WSF
cannot be determined as a simple ratio from short-term creep test data. Rather, careful
extrapolation must be made to stresses and lifetimes that are relevant to the intended application.
The implication of the poor cross-weld creep strength and the weld SRFs introduced by ASME
described above is that either operating temperatures or pressures will have to be lowered or the
pipe wall thicknesses will need to be increased, unless the problem of type IV and associated poor
cross-weld creep strength is resolved.34
Allowance for the cross-weld creep strength at elevated temperature being lower than that of the
parent steel is likely to feature in further standards in the future. Towards this end, reliable
extrapolations of WSF are needed, preferably up to 200 000 h. EN 13445-2:2002 (E) Issue 35
(2009-01), Annex C, Procedure for determination of weld creep strength reduction factor
(WCSRF), which is based on VdTU V-Merkblatt 1153, requires testing at stresses selected to give
durations up to one-third of the creep design life at two test temperatures within a range of
30C of the mean design temperature.44,45 Hence, if steels are to have a design life of 20 years,
then a 7-year testing programme will be required. The standard states that if the failure is located
in the HAZ extrapolation is not allowed without further testing at longer times showing no further
apparent decrease (in the WSF); see Appendix 1. However, for materials operating at the high
temperature end of their application ranges, there is ample evidence to show that the
weldment/parent creep strength ratio continues to decrease with increasing time (and hence
decreasing stress). Hence, test durations for qualification to this code are likely to be similar to
that of the intended service life!
The guidance incorporated in the Volume 5 Part IIb of the ECCC Recommendations (2001) is more
realistic. The criterion for allowing extrapolation of the creep rupture strength R u(W)/t/T by a factor of
3 on life, beyond the life of the longest test, tu(W),max, is:

If WSF(t, T)>01. WSF(t, T) between 086xtu(W),max


and tu(W),max then Ru(W)/t/T may be extrapolated to 36tu(W),max If DWSF(t, T) .01. WSF(t, T)
between 086 x tu(W),max and tu(W),max then extrapolation is not advisable. While the R5 procedure,
developed by British Energy and Serco Assurance, and reviewed by Ainsworth, has a section
(Volume 4/5) that provides two methods for calculating creep/fatigue crack growth. It incorporates
fatigue SRFs to modify the stress range, but it does not appear to treat explicitly the lower crossweld creep strength compared with that of the parent metal.46 Further discussion of the
allowances made in national codes has been given by Ref. 47.

Microstructural degradation and failure mechanism


Type IV failure is the result of localised strain and void formation in the outer region of the HAZ
(FG/IC HAZ). This creep-weak region accommodates the vast majority of strain detectable in the
creep specimen and, as such, specimens fail with markedly low total strain to failure compared
with parent steel specimens or simulated HAZ specimens. Voids nucleate on precipitates mainly at
prior-austenite grain boundaries and, at an advanced stage, the voids coalesce or unzip to form a
crack. Once this final stage begins, creep crack growth progresses rapidly to failure, and
immediate repair or replacement is required. Investigation of creep crack propagation in grade 91
steel will ultimately help utilities to take decisions on the safest and most economical way to
operate, once a type IV crack has been detected. 48
Creep-resistant steels rely on alloy and precipitate strengthening, with the most important fraction
of strengthening understood to be due to fine MX precipitates such as VN and NbC. During the life
of the steel, many intermediate, metastable precipitates exist which, during tempering, PWHT or
service, transform to equilibrium phases, sometimes via another intermediate precipitate.
M23C6 accounts for a large volume fraction of the precipitates present early in the life of such
steels, but it is known to coarsen quickly, thereby removing any precipitate strengthening it
contributed. Similarly, M2X precipitates grow quickly and also have a passing contribution to
strengthening. The dislocation density reduces during tempering, PWHT and service, and the
martensitic laths break down into softer equiaxed sub-grains as the martensite tempers further to
ferrite.49 The MX precipitates NbC and VN are relied upon as the main strengthening phases in
these steels, due to their low coarsening rates and their ability to pin grain boundaries and
dislocations. NbC forms at high temperatures, where it can prevent grain growth and is therefore
important during the steelworking operation, as well as during service. VN forms during tempering,
precipitating throughout the matrix. MX precipitates are obstacles to dislocation movement, but
possibly of similar significance is their role in pinning sub-grain boundaries and preventing loss of
strength through grain recovery and growth.50
A high chromium content can prevent MX precipitates being the equilibrium phases, and instead Zphase can form. Dissolution of the strengthening MX phases into a complex nitride Cr(V,Nb)N
known as Z-phase can take many years, and is usually accompanied by a sharp fall in creep
strength.51 Another precipitate that forms during service is the intermetallic known as Laves
phase, Fe2(W,Mo). This phase is found in Mo- and Wcontaining steels, and is generally thought to
weaken the steel through reduction in solid solution strengtheners, but a degree of precipitate
strengthening from Laves phase may counteract this to some degree. The range and scale of the
phases that exist at any time are dependent on the temperature of operation and the sequence of
precipitation reactions.52,53 Coarsening kinetics lead to sharp changes in creep behaviour, for
example, dual phase steels (containing retained -ferrite) demonstrate a sudden drop off in creep
strength at long durations, associated with the accelerated growth of

20 Creep rupture strength of the COST development steels FB2, FB6 and FB8 as a function of time to rupture
at 650C. FB6 and FB8 both suffered from shortterm onset of Z phase causing the creep rupture strength to fall
dramatically40

precipitates at the delta ferrite/martensite interface. Also Z-phase, precipitated in various high Cr
steels, is known to form after extended creep service durations as the final equilibrium phase. Until
recently, this phase was unknown in creep-resistant martensiticferritic steels, and none of the
thermodynamic databases available even predicted its occurrence in these steels. The onset of Zphase formation after long-term exposure results in a rapid reduction in creep properties; see Figs.
20 and 21. Significant efforts towards developing a series of high Cr alloys, with improved creep
resistance, have been thwarted due to the inability to predict the existence and onset of Z-phase
formation. This example shows how the debilitating effects of unexpected phases can invalidate
predictions made from short-term data or data acquired under different operating conditions. It is
for just such reasons that development programmes should always employ extensive long-term
testing.
The thermal cycle imposed by welding that produces the FG/ICHAZ type IV region has a strong
adverse effect on the optimum distribution of the precipitates and their interparticle spacing, which
the manufacturer generated by judicious alloying and careful processing, and which are vital for
high creep strength.54
Commonly, the FGHAZ is reported as the most creep-weak region and location for type IV
cracking. This is often supported by minimum creep rupture strength data from simulated
specimens. In this region, where temperatures are typically between 900 and 1100C during
welding,55 complete reversion to austenite occurs. Partial dissolution of the M 23C6 phase is
expected followed by rapid growth according to equilibrium and kintetics calculations and diagrams
for 912%Cr steels. MX particles would also be expected to grow at a somewhat increased rate.
An in-depth TEM study of the distribution and composition of precipitates within this and other
regions of the HAZ via simulated thermal cycles, in the as-welded, post-weld heat treated and
creep exposed condition, confirmed that the dissolution of M 23C6precipitate in the as-welded state
was more evident in the FGHAZ compared with ICHAZ and parent specimens. The study showed
that precipitate distributions in the PWHT state were similar for all specimens, but after further
temperature exposure during creep tests, the number density decreased rapidly, i.e. growth rates
were comparatively high, in the FGHAZ. Furthermore, they found that the MX particles contained
relatively high amounts of Cr, leading to the conclusion that Cr diffusion (rather than V) was
controlling the growth rate of MX particles following greater dissolution of Cr containing
M23C6 particles during simulation heat treatment. Hence, the faster growth rate and demise in
coherency and creep resistance of the FGHAZ. It should be noted that the steel contained
approximately 3%Co and 3%W and no boron. The reorganisation of austenite grains is also an
important factor as it is the preferred site for M23C6 nucleation. During the welding thermal cycle,
the boundaries change position, and the carbide locations remain the same, and can no longer
provide grain boundary hardening benefits.56

As mentioned previously, the ICHAZ region has also been suggested, albeit in fewer references, as
the least creep resistant and therefore most likely type IV

21 Predicted reduction in backstress with time associated with the growth of different precipitates for COST
alloy CB8. The marked reduction in backstress due to Z-phase explains the poor creep strength of alloys that
suffer from precipitation of this phase during extended service101

location. The appearance of M23C6 in this region has been found to be much more spheroidised
compared with those in the FGHAZ or parent steel, making for a striking difference between the
TEM images of the regions.68 In this region, some austenite exists; some precipitates are
dissolving, while others are coarsening at relatively higher rates compared with the surrounding
parent steel. Concentration gradients between austenite and a quickly tempering martensite are
formed and a less homogenous microstructure results on cooling. Lee et al. reported a
substantially higher growth rate for M23C6 precipitates in the ICHAZ.57 The growth of creep cavities
that nucleate on these particles is assisted by the creep deformation occurring in this HAZ region.
Francis et al. and Smith et al. quantified the number of creep cavities as a function of the fraction
of rupture life.24,58 Kimura et al. attributed the low creep rupture strength of grade 122 steel crossweld specimens to the concentration differences arising between austenite and martensite existing
in the intercritical range.59,60 This increases the driving force for diffusion, and promotes recovery
of martensite that contains less solid solution strengthening and less effective precipitates at
boundaries defined by the phase interfaces that existed during welding.
In summary, the microstructure of creep-resistant steels is a finely tuned dynamic system, with
changes in the character of the precipitates occurring alongside the movement of dislocations,
subgrain boundaries and grain boundaries during PWHT and subsequent creep service. The weld
thermal cycle imposed upon the HAZ creates a severe disturbance in the finely balanced
microstructure, with a severe detrimental effect on creep strength. Although simulated specimens
demonstrate fairly consistently that the FGHAZ has the poorest creep resistance across the HAZ, in
practice, it is difficult to differentiate between FG and ICHAZ when trying to define the exact
microstructure that a type IV crack has nucleated in and propagated through in a real welded joint.

Hardness
Hardness is often used as a quick quality assessment for parent steel, before and during service.
Measuring the hardness in different parts of the HAZ and linking this to performance is more
difficult.61 Hardness measurements made on polished samples will reveal changes quite well. For
grade 91 parent steel, it has been asserted that, if the correct heat treatment has been received,
then hardness results should fall in the range 200 270 VHN. 62 For a properly heat treated grade
91 weldment, Cohn et al. indicated that, the normal hardness range is 200295 HV, i.e. quite
similar to the parent steel range.63 Hardness generally decreases across the weld from the fusion
line towards parent material with respect to the peak temperature of the welding thermal cycle

that the region has experienced. However, defining the particular region of the HAZ (CGHAZ,
ICHAZ, etc.) that the indent is made in is often not straight forward. Some weldments do not
display the typical HAZ characteristics as clearly as others. Relationships between room
temperature hardness and the time to rupture have been reported for grade 91 parent steel.
Masuyama demonstrated that the instantaneous (room temperature) hardness H of the gauge
portion of a creep test specimen, when normalised by being divided by the hardness H0 of the
(aged but not strained) grip region, is as follows64
H/H0=098-015t/tr (5)
where t is the test duration and tr is the time to rupture. This equation gives an approximate
indication of the hardness change that corresponds to end of life, which occurs when the hardness
has fallen by approximately 17% (when t/tr=1 and H/H0=083). The creep rupture life tr can be
predicted from a knowledge of the test duration t and determination of the hardness values H and
H0
tr=015t/(098{H/H0) (6)
Endo et al. have given a further example of an equation for the (room temperature) hardness of
grade 91 steel
H=207{294(t/tr) (7)
This equation can be written as
H/H0=1-014t=tr (8)
with the value of H0 (207 HV for their particular steel) being the initial hardness, with end of life
corresponding to a 14% drop in hardness from its initial value. 65 Sposito et al. state that Vickers
hardness correlates very well with creep life if care is taken to prepare the surface before
inspection, but large measuring errors limit the applicability of this technique in the field.66
In a study of creep degradation in grade 91 steel weldments, Masuyama noted that the lowest
hardness in the HAZ was approximately 10 HV below that of the parent steel, and that this
hardness difference persisted throughout almost the whole of the creep life (for creep life fractions
from 02 to 09), as both parent steel and HAZ softened, displaying a linear relationship with the
creep life fraction (t/tr).64,67
Hardness testing of specimens subjected to thermal cycles representative of different regions in
the HAZ, followed by PWHT indicated that the lowest hardness following PWHT is in the
intercritically heated specimens (Ac1>temperature<Ac3) corresponding well to the type IV location
in the same test specimens.68 However, the creep rupture life of simulated HAZ specimens has not
always been shown to correspond with the hardness minima. It has been shown that the test
stress has an influence on the ranking of the effect of heat treatment with respect to rupture time.
For grade 122 steel subjected to a simulated HAZ thermal cycle plus PWHT, it has been
demonstrated that not only does increasing the stress reduce the overall time to rupture, but it
shifts the shortest time to rupture from specimens simulating the FGHAZ to those simulating the
ICHAZ. Thus, the higher stressed specimens rupture sooner in lower temperature simulated
specimens of low initial hardness, as seen in Fig. 22. 69,70
This may go some way to explaining why type IV failure has been reported in both regions, but it
follows that the FGHAZ is the weakest in creep for the low loads expected during service
conditions. Albert et al. also observed creep rupture minima in specimens heated above AC3 for
grade 122 steel, and detected hardness minima at intercritically heated temperatures. 71 By
contrast, Laha et al. found that, after PWHT of simulated grade 91 specimens, the creep-weak
region

Plot of Vickers hardness (open symbols) and creep rupture time (closed symbols) for grade 122 HAZ peak
temperature simulated specimens after PWHT70

corresponded to the lowest hardness and an intercritical heat treatment. 68


In summary, hardness minima across the HAZ do not always represent the most creep-weak
region. However, the overall fall in HAZ hardness during service may be used to estimate
remaining life in a way similar to that in parent material. The estimation of remaining life is
discussed in Appendix 2.

Reported examples of failures, mainly by type IV cracking


As noted above, the cross-weld pressure stress in longitudinal seam welds is twice that in a girth
weld. It is therefore the former that is potentially at greater risk of type IV cracking and that poses
the greatest risk to safety (although the weld strength factor will be higher for the seam weld than
for a girth weld in the same pipe). However, type IV failures have often occurred in other weld
configurations, including at least one example of a weld attaching an end cap to a header.72
The deployment of grade 91 steel in utility power boilers as thick section components began in the
UK in the late 1980s on the basis of the performance of the parent metal strength. 7274 Only since
the late 1990s has attention moved to weldment performance, once the susceptibility of this steel
to type IV cracking was recognised. Type IV failures in grade 91 components have been more
common in the UK than in the USA, as this steel grade has been in service longer in the
UK,62 although it has found widespread use in power plant in the USA, where several failures have
occurred, although these have largely been in lower alloy steel grades. 62,75 At least one instance of
type IV cracking in a grade 91 longitudinal seam weld is known to have occurred in Japan.
Reports of at least six failures at the West Burton power plant in the UK and over a hundred
instances of type IV cracking are known. It is also clear that older units designed for base-load
operation and used in this capacity over many years are very susceptible to component failure
when they are then cycled regularly. One of the West Burton failures occurred in a forged header
end cap, and four occurred in bottle type joints, the first after only 20 000 h of service at 565C. A
contributory factor was reported to be the comparative strength of the parent steel, implied by the
low hardness of the grade 91 steel. The steel contained a low nitrogen to aluminium ratio, which

gave rise to AlN precipitates, and reduced the amount of nitrogen available to give the fine MX
intragranular precipitates, thereby reducing the creep strength of the steel. The specification for
the aluminium content has since been revised from #8804;004 to #8804;002% in the ASTM
material standards, but not yet in the equivalent European standards. On a retrofit header
installed on a 500 MW unit in 1992, and operating for 58 000 h at 568uC, over 100 stub welds are
reported to be affected by type IV cracking. Again a low nitrogen to aluminium ratio was
implicated.7678 A sharp change in section was a further contributory factor to an end cap failure
after 36 000 h of service. The seriousness of the problem and its frequent occurrence is illustrated
by the report of 100 instances of cracking during 2004 in outages in power stations operated by
Innogy/RWEnpower.61 Type IV cracking in grade 91 retrofit headers that have been fabricated to
the highest standards have been reported more recently, with the predominant location being the
stub welds.15,79 Clearly, in some instances, the composition and geometry of steel constructions
have been responsible for poor performance, and incorrect heat treatment is also suspected to
have contributed in some instances. However, as noted earlier, warnings of likely problems from
failure in the type IV region, if adequate provision for the system stresses was not made, were not
heeded. Moreover, the construction codes do not yet all require that adequate provision be made
for the reduced cross-weld creep strength compared with that of the parent steel, and hence
systems in steam service have suffered failures, and others may still be at risk of type IV failure,
even though they were code-compliant when fabricated.
A comprehensive review on type IV cracking has been presented by Ellis and Viswanathan. 80 They
included consideration of service experience of type IV cracking in girth welds and in seam welds,
with separate consideration of steam lines and headers. Seam-welded pipe failures have occurred
in the USA, for example the 1985 failure of a 30 inch (760 mm diameter) steam reheat line of the
Mohave power station that had been in service for 14 years. This failure, however, was not type IV
cracking, but rather occurred because of the poorer creep performance of the weld metal
compared with that of the parent pipe. However, the third and fourth failures at the Mount Storm
power plant were attributed to type IV cracking.80 While seam-welded pipe is not in common use
in the UK, clam-shell seam-welded grade 91 elbows have been installed in at least one UK power
station (presumably because this required one mould for the pressing, rather than two). Since the
seam welds occur along the extrados and intrados, stresses on the welds arising from flexure of
the pipe induced by any temperature fluctuations will be higher on the extrados than if the welds
were situated on the neutral axis, and will be in addition to the pressure stress. Such a design of
elbow requires careful evaluation to ensure that it does not give an unacceptably short life. Several
failures due to creep rupture were reported in the USA as early as 1996, 81 and the midspan of the
weldment on the extrados has been identified as a region for the early formation of creep
cavities.82
In China, there have been a total of six deaths in three separate incidents due to failures of seam
welds in grade 91 pipe. The best documented failure occurred during commissioning of unit 2 at
Datong Power Station in 2006, and resulted in two deaths. The main stem line was not
manufactured in grade 91 steel, as had been believed at the time of installation. Similar failures
have occurred elsewhere, with an additional four deaths, and there are believed to be an additional
30 plants containing this pipe. Such short-term failures are unlikely to have been the result of type
IV cracking; however, they do illustrate the potential danger arising from the use of seam-welded
pipe in power plant.

Type IV creep life prediction


Introduction
The ability to predict accurately the creep strength of specimens/components subject to long-term
exposure is highly desirable, since it would avoid the need for longterm testing during the
development of creep-resistant steels, and also allow plant operators to make more confident
decisions on maintenance schedules. Despite advances in methods of prediction, confidence in a
particular alloy is only built up after a degree of verification through testing. Clearly, adequate
appropriately long-term cross-weld testing is vital before the acceptance of any new creepresistant alloy.

Parametric studies
Parent steel

The equilibrium phases that exist in a steel and the kinetics of formation vary according to the
composition, temperature and stress, which in turn affect creep performance and may cause
significant changes in performance. Therefore, extrapolating results obtained at one temperature
to predict results at a different temperature will not always be valid, as different phases may exist,
and different mechanisms may be operational. For these reasons, the use of the LarsonMiller (L
M) parameter to predict performance at different temperatures is not encouraged, as it can yield
nonconservative lifetime predictions. Thus, the LM parameter, although useful in some instances,
do not always provide a reliable method for the interpolation and extrapolation of data over long
durations or interpolation between different conditions of temperature and stress. This is
fundamentally due to the differences in microstructure developed under different conditions, and
over the lifetime of a specimen that can give differences in creep resistance and differences in
fracture mechanism between short-term and long-term creep rupture tests or service, as
discussed below. However, the LM parameter is widely used
L-M=T(C+logt)/1000 (9)
where T is the temperature (K), t is the time (h) and C is the material constant, commonly set at
20, as empirically values very close to this have been found to be quite accurate for 912%Cr
creep resistant steels. Using values of 30 and above, as is sometimes done for CrMo steels,

23 Creep rupture test results for grade 122 showing a distinctive sharp inflexion in creep rupture stress. This
inflexion is associated with the presence of delta ferrite and Z-phase88

can introduce large differences in the estimated value of equivalent times at different
temperatures.83
In his discussion of parametric methods for the extrapolation of high temperature data, Goldhoff
favoured determination of the LM constant from the dataset being analysed. 84 However, he
obtained smaller errors in predicted values of rupture stress from the MansonHaferd (MH)
analysis, than from the LM or Dorn methods. The MH parameter is given by
MH=(T-Ta)/(logtr-logta)
It is derived from plots of temperature, T versus logt r, at constant stress, where tr is the time to
rupture, and Ta and logta are material constants which describe the point of convergence of the
isostress lines.85 One form of the Dorn equation, relating strain rate and stress, has been given by
Evans86

(10)

where DL is the lattice diffusion coefficient, is the shear modulus, b is the Burgers vector, a is
the effective stress, k is Boltzmanns constant, T is the absolute temperature and A and n are
numerical constants determined by experiment.
By splitting the data into regions, and assigning different constants to each part, time/temperature
predictions become more accurate and useful, but difficulties still exist when trying to predict
behaviour at different temperatures where different microstructural regimes are operating. Kimura
and researchers have coined the term region splitting analysis for one of the very first methods
used to separate out different regions of a creep rupture plot. 59,60,87,88 They observed that, when
creep rupture strength is plotted against time, there is an inflection in the curve, which occurs at a
stress that is approximately 50% of the 02% offset yield strength; see Fig. 23. Creep deformation
in the low stress regime is governed by diffusion-controlled phenomena, while in higher stress
regimes, it is controlled by dislocation glide. For dual-phase steels, i.e. those containing -ferrite,
such as grade P122, this inflection is more pronounced.

24 Normalised rupture stress as a function of time for failure for grade 122.

is creep rupture stress, TS is tensile strength at test temperature, tf is time to failure,


Q is activation energy for diffusion through the matrix (300 kJ mol -1), R is universal gas
constant (8314 J K21 mol-1), T is temperature89
In their analysis of grade 122 (parent steel) creep rupture data, Wilshire and Scharning noted that,
not only are long-term tests required in order to provide extrapolated design data, but that the
allowable (parent material) stress estimates have been reduced progressively as longer-term
measurements have become available. 89 They showed that, if creep rupture stress values for
grade 122 steel are normalised by dividing by the tensile strength at the test temperature, data
derived at different temperatures can be superimposed using a function including time to rupture,
temperature and the activation energy for matrix diffusion (300 kJ mol21); see Fig. 24. This
appears to be a more reliable and fundamental way of relating data collected at different
temperatures to each other than the usual approach of applying the LM parameter. Taking
logarithms yields a straight line plot that allows simple extrapolation of data. For grade 91 parent
steel creep rupture data, they produced extrapolations from 5000 h and from 30 000 h out to
times up to 200 000 h, and presented comparisons with measured data at 550, 600 and 650C. 90

Such an approach could potentially reduce the number of creep rupture tests required to
characterise a parent steel, and give greater confidence in extrapolations to longer lifetimes. A
similar analysis, with stress normalised by dividing by (initial) Vickers hardness, has been applied
to weldments of 225%Cr1%Mo steels by Brear et al. 93 Complex polynomial parameters are being
refined using extensive datasets in a rigorous mathematical fashion to give a single best fit curve.
However, such an exercise is particular to the steel grades involved, and seems to add little to the
accuracy of extrapolations when there is already a vast dataset available. Such a parametric
analysis serves only to give the most probable lifetime estimates, based on existing data.
Extensive numerical manipulation of data has been performed by organisations such as the ECCC.
As a consequence of the availability of increased amounts of data and of more reliable
extrapolation methods, greater confidence can increasingly be given to long-term creep strength
extrapolations for parent steels. It is important to establish reliable representations of parent steel
creep behaviour not only for design purposes, but also to assist in the derivation of more reliable
extrapolations of weld strength factors; see the section on Design codes, published weld strength
factors and extrapolation of data.
92

Weldments
The inflection observed in parent steels, as mentioned above for dual phase steels, is relatively
minor compared with the more pronounced change in performance seen at the type IV threshold
stress for welded specimens. As noted earlier, a graph showing the rupture life at which the
transition occurs in grade 91 steel weldments as a function of temperature has been presented by
Middleton et al.; see Fig. 14.7 Equations defining the transition to type IV in terms of time,
temperature and stress have been created for grade 91 steel by various authors. Bell gave an
equation that defined the transition from parent metal to type IV failure of cross-weld creep
specimens in 9%Cr steel; see Fig. 25a 94

(11)

Nath and Masuyama, having had sight of Bells review before its publication, derived a broadly
similar expression to represent their experimental data95

(12)

where tf is the rupture life in h, Tk is the temperature in K and s is the (uniaxial) creep rupture
strength in MPa. The reported application ranges are 8431005 K (570 732C), 4075 MPa and
411 600 h. The equation was subsequently recast in a form similar to that of Bells original
equation by Brear and Fleming;16 see Fig. 25b

(13)

As Fig. 25b shows, equation (12) gives a better fit to the specific experimental data presented
than does Bells original equation. Nevertherless, Bells simple version developed using many data
points from different weld types is often used as a good rough guide for studies of type IV
cracking, with good reason.

Modelling
The limitations of parametric methods have meant that more fundamental physical models based
on microstructural evolution are being developed for lifetime prediction. Various institutions have
been involved in the development of physical microstructural models for creep-resistant steels that
predict the evolution of microstructure and properties at specified times and temperatures. The

fusion zone itself in multiple pass welds is recognised as an extremely inhomogeneous entity. This
is reflected in the studies conducted by Hyde and co-workers, for example Hyde and Sun, whose
models reflect the presence of columnar as-deposited and equiaxed reheated weld metal
regions.96 The studies of these authors include the determination of longitudinal and transverse allweld metal creep data for incorporation into their model.
Much of this review has concentrated on the lower creep strength of the HAZ, and the consequent

25 Creep rupture data and the lower bound creep rupture strength for parent steel and type IV cracking

a the equation due to Bell;94 b the equation due to Brear and Fleming16
premature type IV failures. Hence, modelling studies that address this issue are of particular
relevance. Kimmins and Smith reviewed experimental studies and finite element
models.26 Recognising that the creep damage is confined to a narrow region in the HAZ, they
devised a novel finite element model of the behaviour of a creep-weak layer beside a single-V weld
(simulating the type IV region) inclined to the stress axis. Their model gave good agreement
between theory and experimental observations when they allowed for relaxation of constraint via
the sliding of adjacent elements. They concluded that the creep-weak type IV layer experiences
no measurable constraint from the adjacent material. In a later experimental programme, Smith
and co-workers concluded that cavitation in the type IV region is a consequence of grain boundary
sliding, leading to relaxation of constraint and multiaxial rupture governed by the von Mises
stress.24 Hence, rather than using conventional continuum damage models in finite element
analyses, alternative models involving mechanisms of grain boundary sliding require development.
They found that the maximum fraction of cavitated boundaries in the type IV region, at times close
to failure, was only approximately 1% (4000 cavities per mm2); this proportion must therefore
increase rapidly before rupture. With such a small proportion of a the equation due to Bell;94 b the
equation due to Brear and Fleming16 25 Creep rupture data and the lower bound creep rupture
strength for parent steel and type IV cracking, according to two different equations, from Brear
and Fleming16 Abson and Rothwell Review of type IV cracking of weldments in 912%Cr creep
strength enhanced ferritic steels International Materials Reviews 2013 VOL 58 NO 8 455 cavitated
boundaries, it is likely to be beneficial to keep the HAZ as narrow as possible, which may be one of
the reasons for the better cross-weld performance of EB welds compared with tungsten inert gas
(TIG) welds, as noted in the section on Improved performance through welding procedure.
The finite element-based creep continuum damage mechanics (CDM) method for modelling the
high temperature creep damage initiation, evolution and crack growth behaviour of cross-weld
specimens has also been used by Hayhurst and co-workers and by Hyde.30,97,98 Bauer et
al.99 have been modelling the use of matching, overmatching and undermatching filler metal,
together with the effects of pipe wall thickness (4247 mm) and weld edge angle (022) for
longitudinally welded pipes in E911 steel. Their modelling has indicated that using an
undermatching weld metal is beneficial by reducing the multiaxiality of the stress state, and thus
the extent of creep damage. Thus, modelling studies are now capable of describing several aspects
of cross-weld behaviour.
Continuum damage mechanics can be combined with the microstructural models to predict creep
strength and other properties. The University of Loughborough and the Technical University of Graz
appear to be leading the development of such models, which use thermodynamic and kinetic
models to predict the growth and distribution of precipitates considered most important for
performance. TU Graz have gone one step further by the application of back stress calculations to
their CDM model.101 Clearly, progress is being made in modelling the growth of precipitates which,
in the fine grain region of the HAZ is a precursor to void formation and linkage to form cracks.
Hence, it should eventually be possible to introduce into the models parameters derived from the
characteristics of a particular material, and to predict its type IV behaviour. Among the
complexities that do not yet appear to have been incorporated into the models is the change in
carbon content beside the fusion boundary that occurs during elevated temperature exposure of
weldments with differences in alloy content, as a consequence of the migration of carbon across
the fusion boundary.

Neural network analyses


If a sufficiently large dataset exists, a neural network analysis can be carried out, to predict the
value of one or more parameters (depending on the software) if the value of all the other
parameters is specified. Additionally, by fixing the value of all but one of the input variables, the
way in which the output variable changes as the remaining variable are changed can be explored.
The quality and applicability of the data-base being drawn on is of great importance in order to be
able to make firm conclusions.
This approach has been adopted by Francis et al., whose analysis indicated a strong influence on
predicted creep rupture strength of welded joints, not only of the test temperature and time, but
also of the normalising and tempering temperatures, the tungsten content, the preheat

temperature and the PWHT temperature.58,102Experimental support for some of this work is
discussed in Appendix 1. A subsequent study demonstrated the beneficial effects of increasing the
preheat temperature and having a steep side to the weld preparation, with the influence of heat
input, in the range 0824 kJ mm-1, being small.28 It is, of course, recognised that a range of
each of the input variables is required in order for trends to be revealed. As more data become
available, this approach will be capable of demonstrating the effects of the variables with greater
clarity.
Since grade 91 steel is strongly resistant to tempering, a PWHT temperature of 760C is generally
used for piping and pressure vessels to reduce the level of residual stress and to improve weld
metal toughness. However, with a lesser improvement in weld metal toughness, PWHT at 680C
has been used for its derivatives in exploratory studies when welding turbine rotors. While this
gives an initial benefit in terms of increased shortterm creep rupture stress in laboratory tests, it is
likely that at least some of the benefit will have been lost when subjected to lower stresses in
long-term service.
Despite difficulties in predicting long-term behaviour of creep properties, it is reasonable to
assume from the available evidence that, as long as only type IV data are used, the extrapolation
of these data and ignoring other short-term data should provide a simple and effective way of
predicting failure to longer durations. Data demonstrating a clear departure from parent material
creep performance should be used for a single particular temperature, if such a simplification is to
be employed. Data from tests somewhat in excess of 100 h at 650C or 300 h at 625C will
display the characteristic departure from parent material data for grade 91 steel; see Fig. 14.

Mitigation of type IV cracking


Introduction
There are four principal ways of improving the crossweld creep rupture strength of creep-resistant
alloys:
(i) normalising and tempering the weld
(ii) adjusting the welding procedure
(iii) modifying the tempering treatment
(iv) alloy development to increase inherent resistance to type IV mechanisms.

Normalising and tempering


This heat treatment may be carried out before incorporation into the main piping network and it
effectively eliminates theHAZ of the weld, causing the microstructure to revert back to a state
similar to that of the parent steel. The same regime of temperatures and times used in the
production of the parent material may be employed and, as with parent manufacture, care in
controlling the heat treatment is essential. Because the entire welded component, which may take
various shapes, needs to be normalised to prevent an HAZ-like microstructure in any part of the
component, there are significant difficulties in carrying out a successful treatment, and various
companies have invested heavily in the optimisation of the process. One such example is
Mitsubishi Heavy Industries in Japan who have reported the successful normalising and tempering
(NzT) of narrow-gap seamwelded P91. The reported results show little or no difference between
cross-weld and parent steel creep test results at durations up to 30 000 h and temperatures of up
to 700C.103 Figure 26 shows the basic details of the process used and the results obtained. There
are also companies in Germany known to be practising NzT for grade 91 pipe that is too thick to be
produced as seamless. The technology is also applicable to thin-walled large diameter pipes. Since
ASME introduced

26 Heat treatment, welding details and cross-weld creep results for normalised and tempered P91
weldments103

the weld strength reduction factor mentioned in the section on Weld strength factor (WSF) and
strength reduction factor (SRF), there has been greater interest in NzT treatments, because this
allows a strength factor of 08 instead of 05, and thus reduces the thickness of material needed.
Obviously, the advantages of thinner pipe need to be weighed against the extra time and cost of
theNzT process. It is not generally practical to normalise the attachment and site welds. However,
such welds are generally circumferential welds; hence, there will still be welds withHAZs in the
systemthat could suffer fromtype IV cracking, and could therefore compromise plant integrity. The
pressure stresses on such welds are half those experienced by the longitudinal welds. Therefore,
unless inadequate allowance has been made for system stresses or local stress concentrations,
they should present a lower type IV cracking risk.
In some instances, it may be possible to butter the ends of grade 91 pipe work with a weld metal
of higher creep strength and then carry out a PWHT, as noted by Coleman and Hainsworth. 104 If the
welding is carried out in a fabrication shop, it may even be possible to normalise the whole

component. The subsequent welding would then leave the site weldment with its HAZ lying within
the buttering layer of weld metal.

Improved performance through welding procedure


Although an HAZ susceptible to type IV cracking will always be produced by a welding thermal
cycle in currently used creep-resistant steels, there appear to be opportunities to improve the
performance by adjustments to the welding procedure.
A novel approach to improving the creep rupture strength of the type IV region, as noted by Bell,
is to carry out a partial tempering of the parent steel before welding and to complete the
tempering treatment as the PWHT, as described by Sikka. 94,105 This approach almost doubled the
short-term (< 10 000 h) creep life, compared with conventionally heat-treated 9%Cr steel
weldments.106 It would be interesting to see the effect over a longer test period. Although it seems
logical that partial tempering of the parent during steelmaking would reduce the size of
precipitates in the steel, this approach is unlikely to be adopted widely. It would require code
changes, and it would be necessary to persuade steelmakers to change the tempering treatment,
and accept a change to the specification requirements. Unless only short lengths of parent steel
were involved, the final tempering, in the field, would presumably require a mobile induction
heating unit,

27 Plot showing improved cross-weld creep performance

and thus this approach would have limited application. However, it may prove of benefit where the
creep rupture life cannot be increased by any other means. Other process variables suspected to
have a strong influence on the type IV creep life of weldments have been highlighted by Francis et
al. with their Baysian neural network analysis.28 These were preheat temperature and PWHT time.
The effect of PWHT on interparticle spacing has also been modelled, showing a negative effect on
creep properties.
A further strategy, which could presumably be combined with the half tempering approach, is to
join modified 9%Cr steels by welding processes that can create a very narrowHAZ (and therefore a
shorter thermal cycle and less time for overageing of precipitates). Abe 69 reported that EB welding
which produces a very narrow HAZ, had a beneficial effect on the creep rupture performance of

the type IV region for HCM12A steel, with more recent data confirming the behaviour; see Fig.
27.107 The approximate HAZ widths of EB and TIG welds were 0 5 and 2 5 mm respectively and
the cross-weld creep rupture life of EB welds was found to be approximately twice that of TIG
welds. Since reduced pressure EB welding eliminates the need for the fabrication to be housed in a
vacuum chamber, this welding process appears worthy of consideration for the field welding of
9%Cr steels.
Albert et al. have shown that simulated FGHAZ and real weldments behave significantly differently
in terms of their creep strain.108 This observation was attributed to the triaxiality introduced by the
different creep properties in the various regions of the HAZ. They pointed out that changing the
weld preparation angle can alter the stress state of the joint, and influence creep results
significantly. They concluded that by reducing HAZ width or the groove angle of the joint, this
stress state can be altered to achieve significant improvement in rupture life of the weld joints. In
their report of studies of the HAZ in grade 91 steel, Laha et al. discussed measures to improve
type IV cracking resistance, namely heat treatment, changing the HAZ width and modifying the
steel composition.68
Modification of the weld metal properties alone has been shown to be ineffective in avoiding type
IV failure in grade E911 steel.109 Below a certain stress, all failures were type IV. Nevertheless,
there are a number of considerations when selecting a weld metal for use with creepresistant
steels. Similar steam oxidation resistance is required to that of the parent steel. However, since
the creep-week region is likely to be the type IV zone produced by the welding, the creep strength
of the weld metal need only to be comfortably in excess of the cross-weld strength. Santella
reported the elimination of a hardness minimum in surveys across the HAZ when the tempering
treatment was carried out at 621C, followed by welding and a conventional 760C PWHT.110 In
subsequent creep testing, failure occurred in the parent steel. While this gives some hope that the
type IV zone may have been largely eliminated as a result of the revised tempering treatment, the
creep test conditions were not disclosed fully, and it appears possible that the testing was of too
short a duration for type IV cracking to develop. It is, however, possible that the extent of
precipitates nucleation and growth during the tempering treatment was so small that the weld
thermal cycle plus subsequent PWHT did not cause them to coarsen greatly beyond their optimum
size. This approach clearly has some similarities with the partial tempering proposed by Sikka,
which was discussed above.105

Alloy development
Boron
In view of the severe degradation in creep rupture strength that welding introduces in currently
available steels, it is desirable to develop steels in which such degradation is much diminished.
One approach to improving the creep rupture strength of the type IV region is to modify the parent
steel composition. The most promising step towards the elimination of poor cross-weld creep
strength has been the addition of boron as an alloying element, shown to improve creep strength
in general for 9%Cr steels; low concentrations are present in coded steel grade 92, for example
(the specified range is 00010006 wt-%).111 Higher concentrations have been shown to
enhance the creep strength, with the TAF steels developed in the late 1990s by Fujita at NIMS
(004%B) still being unsurpassed.112,113 Boron is reported to combine with M23C6 precipitates and
prevent their coarsening and the onset of tertiary creep.114 Abe explained that theoretically,
boron will segregate to grain boundaries during austenitising at 1100C. 115,116 During tempering at
~800C, boron at the boundaries is incorporated into the precipitating carbides, giving
M23(C,B)6precipitates. Confirmation of the segregation of boron to grain boundaries and its
presence in the carbides was obtained by autoradiography.114 Das et al. compared a grade 91 steel
and a boron-containing variant. 117 In addition to a uniform grain size across the HAZ for the boroncontaining steel and a smaller trough in the hardness profile, they claimed that the consequent
reduction in the metallurgical notch contributed to the avoidance of type IV failure.
Unfortunately, TAF steels were not conducive to welding or forming, but recent experimental
compositions appear to have overcome these difficulties. These new experimental alloys include
MARBN steels developed at NIMS and a similar version NPM1 produced at TU

28 Creep rupture data of Abe for parent steel and crossweld specimens of 92 grade and a boron-containing
steel tested at 650C 115

Graz.109 The most beneficial aspect of these steels, which contain 3%Co, 3%W, high boron and low
nitrogen, is their apparent immunity to type IV cracking. The creep strength of the weldment in
boron-containing steel appears similar to that of the parent steel, as illustrated in Fig. 28 by Abe
and by Kondo et al., who found that no type IV failure occurred in the high boron steel. 115,118 These
steels have carefully controlled boron and nitrogen additions and have been subject to a
normalising treatment at 1150C. These measures prevent the formation of coarse boron nitride
precipitates that are known to be detrimental for creep properties. Also, the HAZ region is not
refined and appears virtually unchanged from its state before welding, i.e. almost the exact parent
microstructure that existed before welding is present post-welding down to the location of prior
austenite grain boundaries and martensitic laths; see Figs. 29 and 30. 119 The propensity for this
behaviour is not well understood but is thought to depend on the relative concentrations of boron
and nitrogen. Figure 31 shows the solubility line for boron nitrides, as determined by observation
of the fracture faces of 9%Cr steels with varying amounts of boron and nitrogen, after tensile
testing.120It shows that, for a given boron concentration, the nitrogen must be kept below a level
defined by the line in order to avoid the formation of boron nitrides

29 Microstructure in steel 'NPM1' before and after welding thermal cycle simulation showing that the location
and orientation of features remain the same119

30 Heated affected zone microstructure in steel 'NPM1' after PWHT showing a clear resemblance to the parent
steel microstructure 109

31 Solubility limit for BN at 1150C for high Cr steel, with the corresponding information for the FeBN system,
derived by Fountain and Chipman 120 superimposed

Also shown is the composition specification for P92. The figure has been adapted from
Sakuraya et al.121
and thus the formation of a refined HAZ. The solubility limit of BN at 10501100C was expressed
by the following equation121
log(%B)={245log(%N)681 (14)
However, the true position of this line is not well defined and alloys with relatively high levels of
boron and nitrogen have been found to exhibit the same advantageous qualities as those with
compositions on the left side of the diagram. Mayr et al. 122 have attempted to further characterise
the solubility of BN for different austenitising temperatures using thermodynamic calculations and
examination of additional melts. Furthermore, the effect of varying the welding thermal cycle has
not been investigated in these, rather special, boron-containing steels. Nevertheless, there is much
excitement in industry surrounding the development of these steels because of their potential to
eliminate problems with type IV cracking through much improved cross-weld creep strength. The
European COST group are planning to scale up a cast of NPM1 (a 9%Cr steel containing B and N,
and high levels of Co and W), and to also adjust B and N levels in the front running steels FB2 (a
9%Cr steel with B and N and lower levels of Co and W) and CB2 (the cast equivalent of FB2 that is
of essentially the same chemical composition), in an effort to take advantage of the positive effects

of controlling these elements. It appears that the production of large heats of steels with these
modified compositions is likely. Information on the performance of welded joints in these materials
is therefore clearly required. Unfortunately, the presence of boron in 9%Cr steels precludes their
use in nuclear applications.

912%Cr steels strengthened by Z-phase


A separate parent steel development involves the formation of fine-scale precipitates of the
equilibrium phase (Z-phase), which is the phase that, when present as coarse precipitates, is
normally associated with the rapid degradation of creep strength at long durations in creepresistant steels. Previously, a 12%Cr steel would have been expected to suffer from the reduction
in creep strength associated with dissolution of finer-scale strengthening precipitates and the
formation of coarse Z-phase precipitates during long-term service, as noted earlier; now it is
possible to precipitate Z-phase very quickly, and in a fine stable dispersion, with higher Cr levels
actually serve to aid this process. In their European patent, Danielsen and Hald described the
production of a high Cr steel in which creep strengthening is achieved via a fine dispersion of Zphase particles, with an average size of < 400 nm.123 This is an exciting development in which the
most stable phase is precipitated during a tempering treatment to give a small interparticle
spacing. This means that the steel cannot suffer from the competitive growth of different phases
during its lifetime, and only the coarsening kinetics of the very stable Z-phase can cause the longterm degradation of creep properties. Another advantage of this type of steel is that higher Cr
levels can be tolerated without limiting the creep life. Higher Cr levels give the additional
advantage of improved steam oxidation resistance. Information on the cross-weld creep
performance of such

32 Creep rupture strength at 600C of some experimental flux-cored wire deposits after PWHT, containing
varying amounts of Ti and Nb, showing improved creep strength at Ti levels of ~0.06%Ti. Adapted from Abson
et al. 83

steel has not been published at the time of writing, but it will clearly be of great interest to
establish the level of susceptibility to type IV cracking and to confirm their expected good steam
oxidation resistance. In view of the stability of the fine-scale Z-phase precipitates, it is anticipated
that this new type of steel will show less degradation of creep strength in the type IV region that
other 912%Cr steels, but this appears not yet to have been demonstrated.

Weld metal development


Weld metal compositions usually follow parent steel compositions. However, often the Mn and Ni
contents are raised to improve strength and toughness. 124 These elements can depress the
eutectoid temperature. It is necessary for the PWHT temperature to be below the Ac 1 temperature
for the weld metal, and therefore certain elements that affect the Ac 1 temperature need to be
controlled. As Mn and Ni depress the lower transformation temperature, the sum of Mn+Ni should
not exceed 15%, or the PWHT temperature range would need to be restricted.125 These
elements also depress the Ms temperature, and this increases the risk of generating retained
austenite.
According to Swanekamp,62 the ASME II Task Group has proposed new PWHT limits on grade 91
components based on Ni plus Mn content, specifically:
(i) PWHT temperature range must be 13501425F (732774 C) if the chemical composition of
the filler metal is not known precisely
(ii) if the chemical composition of matching filler metal is known precisely, the maximum PWHT
temperature can be increased to 1470 F (800 C) (if Ni+Mn,10%) or 1450 F (788 C) (if Ni+Mn
is between 10 and 15%).
In view of the low creep strength of the HAZ compared with the parent steel in most creepresistant steels, the creep strength of the weld metal is of secondary importance. As noted above,
weld metal compositions usually follow parent steel compositions; however, weld metals do not
display type IV cracking. In the light of the promising results displayed by the new generation of
boron-containing steels that appear to have a reduced susceptibility to type IV cracking, the creep
strength of the weld metal could become the limiting factor for cross-weld creep strength. Further
development of welding consumables must await the commercial realisation of newer steel
compositions, but hopefully will not lag too far behind.
Although titanium is potentially a useful addition for the formation of fine MX precipitates, which
enhance creep strength, high solution temperatures are required during steel processing, leading
to extra costs for the steelmaker, making it unattractive for parent steels. Taneike et al. and Abe et
al. employed a solution treatment temperature of 1300C to take titanium carbo-nitrides into
solution and observed a substantial reduction in the minimum creep rate at 650C in 84%Cr
steels containing 0047%Ti, 013%C and 0006%N.126,127
While it may be impractical to employ such a high solution treatment temperature for parent
steels, due to cost implications, welds are effectively small-scale castings that undergo rapid
solidification, so the incorporation of small additions of Ti into 9%Cr weld metals is more
straightforward. Sperko suggested that Ti is an effective substitute for Nb. 125 While the recovery of
Ti in weld metal is difficult to control, partial substitution for Nb may be beneficial. Ti has been
shown to be beneficial for the creep strength of weld metal; see Fig. 32. However, the potential
detrimental effect of Ti on toughness must be recognised and the problem addressed. 83

Repair of type IV damage and avoidance of PWHT


On a basic level, repairing a type IV crack in a weldment can be dealt with in a similar way to any
other weld

33 Effect of PWHT on weld metal toughness demonstrated with Chromet 9-B9 129

repair, i.e. the damaged area is generally removed and either a fill weld is used or a new piece of
material is welded in. Matching fillers and PWHT are standard practice, and procedures used during
fabrication are used for repair of service-aged plant. However, there are a number of complications
and issues with creepresistant steels that require consideration. The introduction of a weld and the
associated HAZ represents a disturbed part of the microstructure where temperatures above the
Ac1 and Ac3 have resulted in a redistribution of elements, including full or partial dissolution of
precipitates and the formation of new martensite. One purpose of PWHT is to temper martensite in
the weld metal and HAZ. However, if a PWHT is imposed, the microstructure developed in the
parent steel during the tempering stage of manufacture can be modified and softened by
overheating or prolonged PWHT, and this is a common cause for concern.
The effect on cross-weld creep properties of an (inadvertent) incorrect PWHT during fabrication is
neither well understood, nor well documented. Most attention has been given to the weld metal
toughness, which is important during start-up. The effect of PWHT on HAZ toughness and crossweld creep strength has not been thoroughly investigated, although Shiga et al. 119 presented
limited data for grade 91 steel showing that the creep rupture strength after PWHT is lower than
that in the as-welded condition.128
The improved toughness imparted by a higher PWHT temperature or longer PWHT time for grade
91 weld metal is shown in Fig. 33.129 While this information relates to the behaviour of one
particular weld metal, it is particularly notable that the toughness is approximately doubled by
imposing a 760C/2 h PWHT rather than a PWHT in the AWS range. The improved absorbed
energy levels imply that the dislocation density is reducing, carbon is coming out of solution
(tempering the martensite) and precipitates are forming and coarsening. While PWHT is commonly
considered necessary to precipitate fine-scaleMX, this might not always be the best approach for
optimising the long-term creep properties, as further tempering occurs at service temperatures.
After nucleation, a prolonged heat treatment may reduce the interparticle spacing, which is
fundamental in influencing creep strength.100 In parts of the HAZ where the temperature has not
been high enough to dissolve precipitates (more particularly the fine VN precipitates, which
dissolve at temperatures above 1050C), the effect of PWHT on creep life may be considered to be
a negative one. However, cross-weld creep testing (where the fracturemode was type IV cracking)
of weldments subject to several different subcritical heat treatments (< Ac 1), considered to cover
most scenarios arising in practice) with a total hold time of no more than 9 h, has shown negligible
difference in terms of rupture life.130,131 It should, however, be recognised that somewhat longer
hold times at similar subcritical heat treatment temperatures have been shown to affect the

precipitates and creep performance significantly in parent materials and are therefore likely to also
affect weldment strength.132134 The same studies detailed the deleterious effects of treatments
above the Ac1 temperature.
In aged material where precipitates have already undergone some coarsening, PWHT will not help
to improve the creep strength of the HAZ, and is more likely to degrade it, thereby tending to
increase the risk of type IV cracking. Since the creep life of an aged material may be somewhat
degraded, it might be possible to relax the creep properties of the filler material used for repair, so
it is essential to have an understanding of the creep properties of the aged material to ensure
adequate strength in the weld metal. While it is the type IV region of the HAZ and not the weld
metal that is the weak link, at least for the current generation of steels, the weld metal selected
must have sufficient creep strength to support any offloading that may occur during service.
Since the toughness of a component is likely to degrade during elevated temperature service, it is
not only during start-up after fabrication that there is a risk of brittle fracture. For repairs during
fabrication or early service, where the toughness properties of the joint before and after PWHT are
known, a decision can be taken as to whether it is safe to omit PWHT. Residual stresses and the
changes in toughness during prolonged service and due to repair are generally not known, and so
PWHT after a repair in aged material is generally considered necessary.
In view of the difficulties associated with PWHT in the field and the likely detrimental effect on
cross-weld creep properties, a weld repair method for 912%Cr steels that does not require PWHT
would be welcomed, as it would save valuable hours during maintenance shutdown periods. Such
techniques are well established for lower alloy steels, as discussed by Friedman, 126 who presented
a four-layer MMA technique, and by Mitchell and Tolaini, 127 who described the use of a low carbon
CrMo flux-cored wire for the repair of type IV cracking. 135,136 For low alloy steels, care is required
to ensure that stress corrosion cracking does not occur before some tempering of the hard weld
metal and HAZ has been effected by exposure to service temperatures. 137139 The use of a
consumable that gives good ductility and toughness in the as-welded weld metal will be necessary
for repairs in aged material without PWHT. Hence, the 9%Cr nickel-based consumable EPRI P87,
which is discussed in the next section, is a natural candidate.
Controlled deposition techniques developed for lower alloy creep-resistant steels to avoid the need
for PWHT, e.g. those discussed in the various articles in Welding Research Council Bulletin 412,
particularly Friedman,126 have not been widely used for the newer, higher Cr steels, even though
there are likely to be advantages if suitable techniques can be developed. 135,140 Maximising the
refinement and in particular the tempering of the CGHAZ and weld metal microstructure through
the application of controlled deposition procedures may be a way to avoid the use of PWHT in
some situations. However, devising such procedures presents a considerable challenge, since
martensite of high hardness (> 400 HV) forms in both the weld metal and the HAZ at all normal
cooling rates, and grade 91, having good creep strength, is very resistant to tempering, A fine
grain size in the HAZ is required (for good toughness), and thus a low heat input must be used;
the temperature must be reduced from 200C (the usual preheat temperature) to 100uC (to allow
transformation to martensite) in between each layer. Further layers need to introduce sufficient
heat to effect tempering, and yet the steel is resistant to tempering.
Furthermore, since stress corrosion cracking has been known to occur in weldments awaiting
PWHT, the hardness must be reduced to a sufficiently low level (which is yet to be defined) to
avoid its occurrence. One benefit does arise from the low martensite transformation temperature
range, namely that the accompanying expansion partially offsets the build-up of residual stress,
resulting in lower levels of residual stress than those arising in 225%Cr1%Mo
steel.141,142 Extensive weld procedure qualification testing will be required before this desirable goal
is achieved. This is an active area of research in several research centres, including TWI.
The necessary procedures can be time-consuming, difficult to carry out and rely heavily on the skill
of the welder. Vekeman and Huysmans used a carefully constructed, controlled deposition
procedure for cold weld repair of new T91, using a combination of the half and temper bead
techniques, with a 22%Cr1%Mo filler.138 Brett commented on weld repair without PWHT of a
grade 91 header: Prior to the 2006 outage, a weld repair procedure to be used without PWHT,
capable of maintaining the header in operation between the 2006 and 2008 outages, was
approved.15 Details of this procedure, which involved the use of 9%Cr1%Mo welding electrodes
(with carbon at the lower end of the available range), have now been revealed. 143 They carried out
CTOD tests on the as-welded repair, and creep tests following a heat treatment that simulated

service, and concluded that a repair had adequate toughness, and adequate creep strength to
serve for the 4-year interval between planed outages.
A novel approach to the improvement of the crossweld creep strength of seam welds in low alloy
steel pipe has been described by Coleman and Coleman and Gandy.144,145 Using flux-cored arc
welding, they deposited a wide three-layer controlled deposition band straddling the whole of the
seam weld, thereby strengthening the weldment and putting the potentially creep-weak HAZ into
compression, and thus extending the creep life considerably. A demonstration repair to a
225%Cr1%Mo elbow was welded with a 232 C preheat and no PWHT.144 In their studies of
repairs in 225%Cr1%Mo and 9%Cr1%Mo steels, Bhaduri et al. used half bead and temper
bead repairs separately, preferring the latter for grade 91 steel. 146 TWI is currently extending the
approach for service-aged tube material. As noted above, a most important consideration is that
the interpass temperature must be low enough to allow complete transformation to martensite in
the weld metal and HAZ (~100 C) between passes (or between layers) compared with the
normally high interpass temperature used (~200 C), so that the martensite transformation
proceeds to near completion, and subsequent passes are able to temper the underlying
microstructure effectively, and effect some relief of residual stresses. The difference in creep
properties between controlled deposition welds and heat-treated welds appears not to have been
investigated. Other methods for the avoidance of PWHT may be worth exploring, including
autogenous remelting to temper the weld beads and the HAZ.
There appears to be little published information on the creep properties of as-deposited weld metal
and asformed HAZ for grade 91 steels. In indentation creep tests, the creep strength of asdeposited grade 91 weld metal was shown to be much higher than that of the parent steel. 147
The service performance of several weld repairs for high temperature power plant applications has
been discussed by Klenk et al.148 Existing methods for the weld repair of low alloy heat-resisting
ferritic steels have been reviewed by Issler et al.149 The comment is made that PWHT is the most
problematic step in practice, and that Ni-based consumables can be used to repair without PWHT,
at least for temporary repairs. The advantages and disadvantages of using Ni-based weld metal
have been reviewed by Brett et al. and are listed in Table 4. 150 One difficulty is that dissimilar welds
created in such a way are difficult to inspect nondestructively, and so are typically replaced with
matching consumables at the next available opportunity. Ultrasonic inspection of dissimilar joints is
receiving much attention at TWI. If successful methods are developed, this could lead to more
widespread use of Ni-based filler materials in creep-resistant steel fabrications for the power
industry.
Clearly, a large proportion of repairs are likely to be carried out following the occurrence of some
type IV cracking. Any repair welding is likely to introduce a type IV region of lower creep strength
than the service-aged parent steel. Hence, careful consideration must be given to the
consequences of carrying out such a repair. The issues that should be addressed in a method
statement relating to the repair welding of 9%Cr steels have been described by Henry and
Bezzant, who stated that such repairs can be successfully performed if:
(i) there is uncompromising attention to detail
(ii) the unique metallurgy of the material is fully understood
(iii) the current condition of the material is fully understood
(iv) the reason for the repair is clearly understood (v) a detailed repair plan is developed by a
competent technical specialist and all aspects of the plan are successfully executed by the end user
or by the mechanical contractor acting on the end users behalf.151
Weld repairs carried out in Swedish power plant have been reviewed by Storesund and
Samuelson.152 Most of the repairs were to 05Cr05Mo025V steel, but some involved repairs
related to service-induced damage in dissimilar welds between low alloy and 12%Cr steels. No
repairs were found in pipework of 912%Cr steels. They gave the following recommendations
relating to repair welding:
(i) avoid welding procedures that may cause a strongly creep soft HAZ. The HAZ will always have a
creep-soft part (the fine-grained and intercritical parts). That part, as well as the whole HAZ, will
be wider with increasing heat input. A wider HAZ decreases the creep strength of the weld.
Coarsening of carbides may be more pronounced during the weld thermal cycle with higher heat
input, particularly in the HAZ. The coarsening of carbides results in lower creep strength, so the
use of a heat input as low as practically possible is recommended. To avoid wide HAZs, it is also

recommended to use interpass temperatures of #300uC. (Higher than 250uC would be more
appropriate for 9%Cr steels.)
(ii) select design solutions that minimise system stresses as (even) small system stresses reduce
the creep life of weldments significantly. If the repair was not dictated by the presence of
appreciable system stresses, it is very important to reduce or eliminate these
(iii) select materials for weld repair that are somewhat overmatched (creep strong) in relation to
the remaining service-exposedmaterial. Hardness testing, replica testing or the extraction of boat
samples, service data and chemical analysis give understanding of the creep strain resistance in
the aged material. New weld metal should be tested corresponding to an ordinary weld test. These
investigations will help selection of weld repair material and procedure. Typically a new weld metal
of the same material as the service-exposed ones will give a strongly overmatched repair
(iv) wide and medium deep geometry of the excavation is optimal for the lifetime of a weld repair.
Deeper or full repairs are, however, necessary if cracks, microcracks or creep cavities are present
deep in the material. Repair of at least the whole width of the original weld, including the HAZs is
always recommended, even if the damage is local, such as cracks in the HAZ only at one side of
the weld.
Table 4 Summary of the advantages and disadvantages of the use of ferritic and Nibased consumables for repair
welding150

Ferritic
Nickel-based
Advantages
Advantages
Minimal material discontinuity Lower residual stress
Intrinsic resistance to hydrogen-assisted
Ultrasonic inspectability
cracking
Better fracture toughness No specialised
Long-term integrity
weld deposition techniques required
Disadvantages
Disadvantages
Higher residual stress
Transition joint limitations
Risk of hydrogen-assisted
Problems of inspectability
cracking
Poor fracture toughness
Specialised weld deposition
techniques essential

Technology gaps and future trends


The use of creep-resistant steels in existing plant and their popular application in current builds
mean that there will be a risk of continued recurrence of type IV cracking for the foreseeable
future, and therefore, in addition to the usual monitoring strategies, detection and repair strategies
will be required for at least the next 30 years. In China and India, where (in the light of the rapid
expansion in electric power generation in these countries) hundreds of new power plants have
recently been built using creep-resistant steels, there is likely to be considerable scope for the
detection, monitoring and repair of type IV cracking. Worldwide standards need to reflect the
knowledge that now exists. There remains a real danger arising from fabrications built to old code,
including seam-welded elbows, with apparently no allowance for the diminished cross-weld creep
strength. While the cross-weld stress to which girth welds are subjected is half that to which seam
welds are subjected, failure to limit system stresses can, as discussed earlier, make girth weld
vulnerable to failure by type IV cracking. Clearly, there is an urgent need to increase awareness of
the severe degradation in creep strength that welding introduces, and to reflect this information
in standards. When improvements are effected in parent steels through boron and possibly other
alloying additions, and cross-weld strength in the type IV region increases significantly, improved
weld metals will be required, to avoid the weld metal becoming the creepweak region. It is
therefore now time to undertake an intensive programme of weld metal development, involving
long-term creep rupture testing.

There is growing awareness of type IV issues in Europe, the USA and Japan, where measures are
being taken to mitigate the risks, and research to develop resistant steels is under way. Following
further development, boron-containing steels are likely to be put into widespread use in fossil-fuel
plant at temperatures up to 650C. Long-term cross-weld data are therefore required for these
newer steels.
In weldments in the 9%Cr steels, martensite forms both in the weld metal and in the CG and FG
HAZs. The transformation from austenite to martensite occurs at a sufficiently low temperature
that the associated expansion will partially offset the contraction occurring during weldment
cooling. The effect of this expansion is to reduce the level of residual stress in this class of steels.
While this is a topic that has not received extensive attention, it is unlikely to be an area of
concern in fabrications subjected to a PWHT. However, if strategies are ever devised for carrying
out repair welding without a PWHT, residual stresses will assume greater significance.
In the near future, N+T treatments to minimise the reduction in cross-weld strength could become
more commonplace, and may also come to be used for seamwelded pipe. This may bring a
different set of problems associated with temperature control and the effect of such treatments on
weld metals in particular. Electron beam welding (which removes the complication of filler metals)
may offer productivity improvements for large repetitive manufacture of some parts. One area
where EB welding may be of use, especially when coupled with an N+T treatment, may be for the
manufacture of large dished ends of pressure vessels by joining forged petals together. However,
further research to prove such technology is required.
An interesting recent development is the revised normalising and (double) tempering treatment for
grade 92 steel that was first modelled and then carried out by Yin and co-workers and Morris and
co-workers.153,154 These later authors showed that a new heat treatment schedule (normalising at
1150C and double tempering at 660C for 3 h+3 h) extended the (parent steel) creep life by
more than a factor of 3, with a further substantial increase arising from a reduction in the C/N
ratio. (This demonstration of the strong influence of normalising temperature provides
experimental support for the neural network prediction of Francis et al., mentioned in the section
on Neural network analyses.102) It will clearly be of interest to establish if the improved creep
performance is sustained for times appropriate to power plant service. So far, no cross-weld data
have been reported, but it will also be of interest to establish the weld strength factor appropriate
to 105 h tests.
To meet future energy needs, base-load power is most likely to be met by nuclear stations, with
renewable energies contributing. In countries where coal is abundant, coal-fired stations will
continue to make a major contribution, particularly as carbon capture and storage becomes viable.
For coal-fired stations and combined cycle gas turbine units, the degradation of alloy steels in high
temperature service as a result of combined creep and fatigue is likely to become increasingly
important. Continued effort is required with data needed not only on grades 91 and 92 steels, but
also on the emerging alloys.
If the pilot plant operating at > 700C shows good progress, then more dissimilar welds between
ferritic and nickel-based alloys will become commonplace, and this may bring new problems,
including the cross-weld creep strength and non-destructive testing difficulties. The welding of
dissimilar metal joints poses special problems in the selection of filler metals and PWHT
temperatures. During PWHT and early service, carbon will diffuse to the higher alloy steel from the
lower alloy steel, thereby creating a carbon-depleted zone in the latter, which is creep-weak. The
issue has been researched by several authors, including Roman et al., who obtained the lowest
cross-weld creep strength of a variety of dissimilar grade 22 to grade 91 joints where the grade 22
was buttered with grade 91 weld metal.155 In his extensive tests, which included elevated
temperature exposure simulating service, Allen concluded that grade 22 weld metal was the most
appropriate consumable for such joints.156 A table giving recommended weld metal compositions
for dissimilar metal joints is given in AWS D108-96.157 A newly developed versatile welding
consumable, EPRI P87, a nickel-based consumable alloyed with 9%Cr in order to minimise the
carbon diffusion that occurs during welding, PWHT and service of dissimilar welds, has been
reported.104,158160 When joining CrMo steel components with buttered joint faces that have been
subjected to PWHT to other similarly prepared CrMo steels or to stainless steels, the completed
joint can be left as-welded.
In the light of the comments made above, important areas for future activity, including research,
appear likely to be the following:

(i) further changes in standards to reflect the low cross-weld creep strength of creep-resistant
steels, and the implications for fabrications, particularly those in steam service
(ii) scaling up the use of boron-containing steels, in the light of research findings of the apparent
absence of the type IV degradation in crossweld creep strength
(iii) new steels strengthened by fine Z-phase dispersions should be further investigated, especially
the effect of welding them
(iv) the consequences of NzT treatments following welding merit further study, as does the EB
welding of thick section units that can be fabricated offsite
(v) further investigation of the use of higher normalising temperatures than those currently in use
should be carried out, together with revised tempering treatments to investigate whether
improvements can be achieved in long-term creep strength
(vi) the consequences for creep life of thermal cycling merit further study, particularly for grades
91 and 92 steels and emerging steels, to Abson and Rothwell Review of type IV cracking of
weldments in 912%Cr creep strength enhanced ferritic steels International Materials Reviews
2013 VOL 58 NO 8 465 ensure that any detrimental effects are understood and allowed for in
design
(vii) dissimilar welds between martensitic steels and Ni alloys are coming into increasing focus in
light of the target 700uC operating temperatures. Increasing use appears likely to be made of the
new Ni-based consumable EPRI P87. Continued research is therefore required to gain a more
complete understanding of the benefits that it provides. The prospect of using it for the repair of
creep-resistant steels without PWHT merits investigation.

Conclusions
From a review of the literature particularly with respect to type IV cracking in 912%Cr creepresistant steels, the following conclusions have been drawn.
1.

Type IV cracking has become a significant problem for steel in creep service. Laboratory
tests have shown that it has generally increased in severity as the creep strength of parent
steel grades has increased over the last 40 years.

2.

The mechanism of type IV failure is strongly linked to the relative overageing of


precipitates in the intercritical and fine grain regions of the HAZ. The welding thermal cycle
causes a degree of premature aging before service and some softening of this HAZ region.
These fine grain regions of HAZ microstructure promote localised diffusion and expedite
precipitate growth during service. The resulting creep weak region accumulates strain in a
continuous narrow band of the HAZ, approximately parallel to the fusion boundary, leading to
void nucleation, cracking and finally to low ductility fracture.

3.

Despite the occurrence of type IV failures in lower alloy steels, grade 91 steel and other
creepresistant steels were employed with inadequate attention paid to cross-weld properties.
Hence, many instances of type IV cracking have been detected in high temperature plant.
Failures have occurred, and continue to occur, as a result of inadequate allowance for the
presence of a creep-weak HAZ region and the influence of service stresses.

4.

The extent of the disparity in creep strength between the parent steel and the type IV
region of the HAZ increases as the creep life increases, i.e. as the applied stress decreases.
Hence, the extrapolation of results from laboratory tests to likely service lifetimes must be
carried out with due care and attention. The inclusion of data collected from long term crossweld creep specimens that fail by the type IV mechanism are essential before any
extrapolation is carried out for the purposes of lifetime predictions.

5.

The introduction of new ASME rules in 2008 addressing the poor creep strength of welded
boiler components should help to eliminate the premature occurrence of type IV damage in
new fabrications.34 However, the recent code changes require the use of thicker components
that are more expensive to manufacture and they may suffer from higher residual, system
and thermally induced stress levels. The increased thickness to mitigate type IV cracking over
component lifetimes and to comply with code requirements, negates many of the advantages
that these steels were designed to offer. As a consequence, at the time of writing, the current
grades of creep-resistant steels available are no longer perceived to offer such an advantage
over older, low alloy counterparts.

6.

The introduction of new alloys with controlled boron and nitrogen levels appears to be the
most promising way forward in terms of mitigating type IV failure in the long term. In the
short term, normalising and tempering treatment of welded components will become more
widespread, with final site welds either in areas of low system stress or of greater section
size.

7.

The most desirable goal in the repair of creepresistant steels and type IV damage in them
is the removal of the need for PWHT. Solutions to this problem may include controlled
deposition techniques derived from those used for lower alloy steels, and also the use of Nibased consumables. The criterion for the need for PWHT and the effect on creep properties is
not clear.

Appendix 1
Digest from EN 13445-2:2002 (E) Issue 35 (2009-01)
Unfired pressure vessels. Materials
Weld creep strength reduction factor (WCSRF): factor to account for possible creep strength
reduction at the weldment
4242 Creep properties of weldments: Creep properties of weld joints subjected to stresses
normal to the weld may differ significantly from those of the base material.
For the design of vessels in the creep range, this is taken into account in EN 13445-3 by making
use of a weld creep strength reduction factor z c obtained from tests on weldments. If no data are
available, a default value of zc is used.
An acceptable method to determine zc by cross-weld tests is given in Annex C (see also Ref. 161).

Annex C (informative). Procedure for determination of the


weld creep strength reduction factor (WCSRF)
The WCSRF will be taken as 1 when all the following conditions are fulfilled by the steel
manufacturer:
1.

Stress rupture tests on weldments made on specimens of the same steel products as used
in the vessel and which are comparable as regards consumable shall be carried out according
to the European Creep Collaborative Committee (ECCC) Recommendations. 162

2.

Two test temperatures shall be selected within a range of 30C about the mean design
temperature. At each of these temperatures, creep tests shall be carried out at stresses
selected to give durations up to one-third of the creep design life (typically 1000, 3000, 10
000, 30 000, 60 000, 100 000 h, etc.). It has to be shown that the lower limit of the achieved
creep values of the welded joint are not lower than the lower accepted scatter band (-20%)
of specified mean values of the creep strength of the base material according to the materials
standard. However if the failure is located in the heat affected zone (HAZ), extrapolation is
not allowed without further testing at longer times showing no further apparent decrease. In
this case extrapolation may be made by a factor equivalent to the factor Abson and Rothwell
Review of type IV cracking of weldments in 912%Cr creep strength enhanced ferritic steels
466 International Materials Reviews 2013 VOL 58 NO 8 showing stabilised conditions used in
these longer tests.

3.

When no cracking in the HAZ has been found in the tests prescribed above, an additional
set of tests at a higher temperature shall be made with the value of the LarsonMiller
parameter (LMP) equal to or greater than that at the extrapolation point This testing shall be
made to confirm that the location of the failure does not change from the base material to
HAZ. The temperature shall ideally be no more than 50C greater than the higher
temperature test in C.2 (in order to avoid an unacceptable modification of the
microstructure). The stress shall lead to a minimum testing time of 10kh. The temperature
and testing time shall be selected so that the creep timetemperature parameter (TTP), e.g.

LMP in these tests is at least the value at the extrapolation point (time and temperature). A
minimum of three samples shall be tested. The fracture location of the creep specimens shall
be checked by microscopic examination.
4.

If fracture location of the creep specimens in C.3 is within the base material, the WCSRF
may be taken as unity for a time equal to the time achieved in the tests in C.2 multiplied by a
maximum of three.

5.

When the creep strength properties of cross weld specimens fall below the minimum value
given in the scatter band a specific weld reduction factor can be used based on the ratio of
the average value of the creep strength compared to 80% of the mean value of the base
material.

Appendix 2
Estimation of remaining life
Introduction
Similar to other engineering assessment work, the first step in high temperature component life
assessment is to determine the likely damage mechanisms involved in the deterioration of the
material. The creep of most components is characterised by continuum damage (creep rupture or
creep strain). However, in some cases fracture mechanics dominates the total life of the
component; examples of this scenario could be thick components such as headers and high
pressure/temperature steam piping. Once the applicable damage mechanisms are identified for
each part of the system (pure creep rupture, creep fatigue, creep crack growth, etc.), an
appropriate assessment procedure, including nondestructive testing, continuous monitoring,
fitness-forservice, etc. can be adopted to ensure the integrity of component in a given operating
interval. Welds that are subjected to cross-weld service stress or to high system stresses are
potentially at risk of failure by type IV cracking. Hence, consideration or the potential to type IV
cracking will figure in any life assessment of systems in creep service.
The following initial screening inspections are recommended, and these can be used in addition to
and before other techniques. (Note, however, that each method may be usable only over a
particular temperature range.)
(i) visual survey for bulging, sagging and general deformation, particularly in weld region
(ii) dimensional checks ultrasonic thickness measurements, tube strapping (for diametric growth
of tubes)
(iii) crack detection for surface-breaking flaws at accessible welds, header ligaments, etc., using
magnetic particle inspection, dye penetrant and/ or electromagnetic techniques to determine
defect size. Ultrasonic inspection can be used for suspected internal cracks
(iv) thermographic inspection through peep holes to check for hot spots on tube banks/boiler
sections, to focus inspection sample locations
(v) examination of thermocouple data from process operators to check for hot regions, to focus
inspection and sampling.
A variety of techniques are available for further examination, including the preparation and
examination of acetate replicas to establish the extent of any creep cavitation; hardness
measurement, small-scale punch creep testing on extracted samples; creep rupture testing on
extracted samples; the continuous monitoring of creep strain; intermittent measurement of creep
strain and microfocus X-ray. The comparatively recent development of equipment for the removal
by electrodischarge machining of a small slab of material facilitates damage assessment and
remaining life assessment, as the sample can be used for acetate replica preparation and
hardness, small punch creep, conventional creep (albeit with a miniature sample) and ultrasonic
testing.163 A critical review of the assessment of creep damage in steels employed in the power
generation industry has been carried out by Sposito et al. 66 In the light of shortcomings of the use
of replicas for damage assessment in the secondary and tertiary stages of creep, they concluded
that ultrasonic and potential drop techniques appear to be most promising, but that further
research is needed before they are fully mature for deployment in the field.

Several techniques are available for monitoring creep strain, and thus estimating remaining life.
Hardness measurements, discussed below, are not sufficiently sensitive to allow prediction of the
creep behaviour of weldments, and thus the risk of type IV cracking. The remaining techniques, if
deployed in weldment regions or across the HAZ, have the potential to do so.

Remaining life assessment from hardness measurements


With the caveat noted in the section on Hardness, namely that hardness minima across the HAZ
do not always represent the most creep-weak region, the remaining creep life can be estimated
from hardness readings. However, it should be stated that using hardness values to determine
remaining creep life is extremely specific to each individual heat of steel, and should therefore be
used as an continuing monitoring technique from one inspection to the next. If the material
hardness is determined at the start of its service life, the approach discussed in the section on
Hardness (equation (8)) can be applied to steel in elevated temperature service, with hardness
being measured at room temperature during a shutdown. Allen and Fenton and Allen have
developed an approach for estimating remaining life from hardness measurements. 164,165 The model
requires the determination of creep rupture strength data that is normalised by dividing by the
high temperature flow stress.

Damage assessment using acetate replicas


In view of the need to have an acetate sheet in contact with the metal surface, this technique can
only be used on high temperature plant during a shutdown. The preparation of acetate replicas
and their examination (usually in an optical microscope), is a well-established technique. However,
care must be exercised to ensure that the surface region being sampled is representative of the
underlying bulk material and subsurface damage. This principle is most likely to be violated where
it is desired to establish the extent of creep cavitation in a highly localised region, such as the
HAZ. Particularly for a double V weld, the presence of creep cavitation near midthickness may not
be monitored readily from the surface, as there will not normally be a similar extent of creep
cavitation at the outer surface.
According to Jaske and Viswanathan, the guidance given by the classification of Neubauer and
Wedel listed below, which was derived from extensive observations on steam pipes in German
power plants, is used by many plant operators in their steam pipe assessment programmes. 166,167
(i) class A: isolated cavities, requires no remedial action
(ii) class B: oriented cavities, requires reinspection within 153 years
(iii) class C: microcracks, requires repair or replacement within 6 months
(iv) class D: macrocracks, requires immediate repair or replacement.

Remaining life assessment from creep measurements


There are two primary methods by which creep measurements can be made, both of which require
extracted samples. Indentation creep measurements can be made, typically on a small slab of the
test material.168 Data generated are compared, for example on a LarsonMiller plot, with
established data for the test material to determine a converted stress. A similar approach is taken
if sufficient material is extracted to permit the machining and testing of uniaxial creep rupture
specimens.169

Continuous monitoring of creep strain


Continuous monitoring requires the attachment of strain gauges. Online displacement
measurements at elevated temperature are typically undertaken using capacitance devices, which
usually monitor the strain in the parent steel.170 The trend in strain accumulation under
multidirectional conditions is similar to uniaxial, i.e. the rate decreases during primary creep before
accelerating in a tertiary stage.

Intermittent measurement of creep strain


An initial image of the surface of the material of interest is taken, and a detailed comparison made
with later images, permitting the microstrain to be determined. Image correlation methods are
more suitable for monitoring the HAZ in a weld.171 Morris et al. discussed the autoreference creep
management and control (ARCMAC) system used by E.ON UK. 172 They report that using a light-

emitting diode light source and appropriate software, the equipment can capture strains as low as
500 microstrain which, they state, is more than twice the expected strain on steam pipe after 2
years service. A similar device, speckle image correlation analysis (SPICA), which requires a light
source and a video camera, has been used extensively by KEMA over several years. The method
involves making an optical fingerprint of a given surface in order to compare it with another image
recorded later. Local and integral strain can be calculated. Evaluation criteria are based on the
results of tests conducted on test specimens; for HAZ, the criteria are based on the elongation of
this particular area. By using electron speckle pattern interferometry, flaws are detected by
examining interference patterns. 173 This technique is reported to be able to detect deformation and
displacement, and allow strain measurement down to 1025. 171
Laser profilometry is used for measuring the inner profile of piping and accurate measurements of
tube diameters are possible, allowing the detection of incipient creep damage by measuring a 2%
diameter increase.174

Ultrasonic testing
For life prediction of super heater/reheater coils, the metal service temperature can be estimated
by in situ measurement of the thickness of steam-side oxide scale, which is effected by measuring
time-of-flight of high frequency ultrasonic waves. 175 The data are used in the software program
developed by BHEL for life prediction. 176Bhattacharya mentioned the use of A-scan for a similar
purpose in boiler components in Indian industry.177
Ultrasonic or radiographic techniques may be able to detect the physical damage due to creep
(cavities or cracks), and thus detect advanced stages of type IV cracking. For improved imaging,
phased array ultrasonics or time-of-flight deflection may be preferred or microfocus X-ray
equipment with digital screens may be used. Acknowledgements The authors are grateful to Dr A.
K. Bhaduri and Dr K. Laha, for supplying data used in plotting Fig. 17 and to colleagues at TWI,
including B. J. Cane, C. J. Ablitt and J. R. Rudlin for their help and advice. The contributions of the
referees, in suggesting many improvements to the manuscript, are also gratefully acknowledged.

Acknowledgements
The authors are grateful to Dr A. K. Bhaduri and Dr K. Laha, for supplying data used in plotting
Fig. 17 and to colleagues at TWI, including B. J. Cane, C. J. Ablitt and J. R. Rudlin for their help
and advice. The contributions of the referees, in suggesting many improvements to the
manuscript, are also gratefully acknowledged.

References
1.

J. Wheeldon, J. Parkes and D. Dillon: UltraGen: a proposed initiative by EPRI to advance


deployment of ultra-supercritical pulverized coal power plant technology with near-zero
emissions and CO2 capture and storage, Proc. 5th Int. EPRI Conf. on Advances in material
technology for fossil power plants, Marco Island, FL, USA, October 2007, EPRI, 82.

2.

M. Fukuda, H. Sone, E. Saito, Y. Tanaka, T. Takahashi, A. Shiibashi, J. Iwasaki, S. Takano


and S. Izumi: Refurbishment of aged PC power plants with advanced USC technology, Proc.
5th Int. EPRI Conf. on Advances in material technology for fossil power plants, Marco Island,
FL, USA, October 2007, EPRI, 29.

3.

R. Viswanathan, R. Purget, S. Goodstine, J. Tanzosh, G. Stanko, J. Shingledecker and B.


Vitalis: US program on materials technology for ultrasupercritical coal-fired boilers, Proc. 5th
Int. EPRI Conf. on Advances in material technology for fossil power plants, Marco Island, FL,
USA, October 2007, EPRI, 1.

4.

P. Peel, B. Scarlin and R. Vanstone: From materials development to advanced steam


turbines, Proc. 7th Int. Charles Parsons Turbine Conf. on Power generation in an era of
climate change, Glasgow, UK, September 2007, University of Strathclyde, 389 402.

5.

F. Lin, S. Cheng and X. Xie: The development of electric power and high temperature
materials application in China an overview, Proc. 5th Int. EPRI Conf. on Advances in
material Abson and Rothwell Review of type IV cracking of weldments in 912%Cr creep
strength enhanced ferritic steels 468 International Materials Reviews 2013 VOL 58 NO 8
technology for fossil power plants, Marco Island, FL, USA, October 2007, EPRI, 46.

6.

H. J. Schu ller, L. Haigh and A. Woitscheck: Cracking in the weld region of shaped
components in hot steam lines materials investigations, Der Maschinenschaden, 1974, 47,
(1), 113.

7.

C. J. Middleton, J. M. Brear, R. Munson and R. Viswanathan: An assessment of the risk of


type IV cracking in welds to header, pipework and turbine components constructed from the
advanced ferritic 9% and 12% chromium steels, Proc. 3rd EPRI Conf. on Advances in
materials technology for fossil power plants, (ed. R. Viswanathan et al.), 6978; 2001,
London, Institute of Materials.

8.

C. H. A. Townley: A life of welds at high temperature, High Temperature Structural Design.


Proceedings, International Conference, Venice, Italy, 24-26 Oct 1990. Ed: L.H.Larsson. Publ:
Bury St Edmunds IP32 6BW, UK; Mechanical Engineering Publications Ltd for European
Structural Integrity Society; 1992. ISBN 0-85298-771-4. ESIS Publication 12. pp.401-413. 1
fig., 1 tab., 4 ref. ISBN: 0-85298-771-4.

9.

C. F. Etienne and J. H. Heerings: Evaluation of the influence of welding on creep


resistance, Steel Res., 1994, 65, (5), 187196.

10.

F. Bruhl, H. Cerjak, H. Mu sch and K. Niederhoff: Erweichungsverhalten der Wa


rmeeinfluzone des hochwarmfesten Chromstahls X10CrMoVNb91 in Schweiverbindungen
und nach Simulationsbehandlung, Schweien Schneiden, 1990, 42, (10), 515520.

11.

F. Bru hl, K. Haarmann, G. Kalwa, H. Weber, H. Cerjak, H. Mu sch, K. Niederhoff, G.


Wellnitz and M. Zschau: Behaviour of the 9% chromium steel P91 and its weldments in short
and long term creep tests, Proc. ASME/IEEE Power Generation Conf., Boston, MA, USA,
October 1990, ASME/IEEE, Paper 90.JPGC/ Pwr-62 Jt.

12.

C. J. Middleton and E. Metcalfe: A review of laboratory type IV cracking data in high


chromium ferritic steels, Proc. Int. Conf. on Steam plant for the 1990s, London, UK, April
1990, Mechanical Engineering Publications Limited for Institution of Mechanical Engineers,
275282.

13.

T.-U. Kern, M. Staubli, K. H. Mayer, B. Donth, G. Zeiler and A. di Gianfrancesco: The


European effort in development of new high temperature rotor materials COST 536, Proc.
8th Liege Conf.: Materials for advanced power engineering, (ed. J. Lecomte-Beckers et al.),
Vol. 53, Part 2, 843854; 2006, Ju lich, Forschungszentrum Ju lich GmbH.

14.

D. J. Allen and S. J. Brett: Premature failure of a P91 header end cap weld: minimising the
risks of additional failures, Proc. Int. Symp. on Case histories on integrity and failures in
industry, (ed. V. Bicego et al.), 133143; 1999, Warley, Engineering Materials Advisory
Services Limited.

15.

S. J. Brett: Service experience with a retrofit modified 9Cr (grade 91) steel header, Proc.
EPRI 5th Int. Conf. on Advances in materials technology for fossil power plants, Marco
Island, FL, USA, October 2007, EPRI, 590.

16.

J. M. Brear and A. Fleming: Prediction of P91 life under plant operating conditions, Proc.
ETD Int. Conf. on High temperature plant integrity and life extension, Cambridge, UK, April
2004, Robinson College.

17.

H. Cerjak and P. Mayr: Creep strength of welded joints of ferritic steels, in Creep resistant
steels, (ed. F. Abe et al.), 472503; 2008, Great Abington, Abington Publishing.

18.

D. J. Abson and J. S. Rothwell: Mechanical properties of high productivity weldments in


the advanced 9% chromium Steel FB2, Mater. High Temp., 2010, 27, (3), 253264. [See
confidential TWI Research Report for Industrial Members, 248/ 2010, March 2010.]

19.

M. Yaguchi, T. Ogata and T. Sakai: Creep strength of high chromium steels welded parts
under multi-actual stress conditions, Proc. 2nd ECCC Creep Conf. on Creep &amp; fracture in
hightemperature component design &amp; life assessment issues, (ed. I. A. Shibli and S.
R. Holdsworth), 215226; 2009, Zu rich, DEStech Publications, Inc.

20.

S. J. Brett: Type IIIa cracking in KCrMoV steam pipework systems, Sci. Technol. Weld.
Join., 2004, 9, (1), 4145.

21.

V. Vodarek, Z. Kubon, R. Foret and S. V. Hainsworth: Microstructural evolution in P23/P91


Heterogenous welds during creep at 500600C, Proc. IIW Int. Conf. on Safety and reliability
of welded components in energy and processing industry, Gratz, Austria, July 2008, Graz,
Verlag der Technischen Universitaet Graz, 233238.

22.

M. Steen, M. de Witte and C. Coussement: Creep properties of 12%Cr weldments. Part 2:


Components tests and design implications, Proc. 4th Int. Conf. on Fracture of engineering
materials and structures, (ed. B. Wilshire and R. W. Evans), 679690; 1990, London,
Institute of Metals.

23.

C. Coussement and M. de Witte: Behaviour of 12% Cr and improved 9% Cr weldments


impact on the occurrence of type IV cracking in service, Proc. Int. Conf. on Joining and
welding 2000, The Hague, The Netherlands, July 1991, International Institute of Welding,
189204.

24.

D. J. Smith, N. S. Walker and S. T. Kimmins: Type IV creep cavity accumulation and failure
in steel welds, Int. J. Press. Vessels Pip., 2003, 80, (9), 617627.

25.

F. Masuyama: Effect of specimen size and shape on creep rupture behaviour of creep
strength enhanced ferritic steel welds, Int. J. Press. Vessels Pip., 2010, 87, 617623.

26.

S. T. Kimmins and D. J. Smith: On the relaxation of interface stresses during creep of


ferritic steel weldments, J. Strain Anal. Eng. Des., 1998, 33, (3), 195206.

27.

D. W. J. Tanner ,W. Sun and T. H. Hyde: The effect of weld fusion zone in a cross-weld
specimen under creep, 12th Int. Conf. on Creep and fracture of engineering materials and
structures, Paper B40 (CD-ROM), 2012, JIP, Japan.

28.

J. A. Francis, G. M. D. Cantin, W. Mazur and H. K. D. H. Bhadeshia: Effects of weld preheat


temperature and heat input on type IV failure, Sci. Technol. Weld. Join., 2009, 14, (5), 736
742.

29.

T. H. Hyde, W. Sun, A. A. Becker and J. A. Williams: Effect of weld angle and axial load on
the creep failure behaviour of an internally pressurised thick walled CrMoV pipe weld, Int. J.
Press. Vessels Pip., 2001, 78, 365372.

30.

T. H. Hyde: Finite element modelling of the welding and subsequent creep damage
analyses of P91 pipe welds, Proc. 3rd Int. Conf. on Integrity of high temperature welds,
London, UK, April 2007, IOM Communications, 227243.

31.

P. F. Morris: Terms and terminology for creep testing, ECCC Recomm., 2001, 2, (IIb).

32.

D. J. Allen, B. Harvey and S. J. Brett: FOURCRACK an investigation of the creep


performance of advanced high alloy steel welds, Int. J. Press. Vessels Pip., 2007, 84, 104
113.

33.

K. Kimura, M. Tabuchi, Y. Takahashi, K. Yoshida and K. Yagi: Long term creep strength and
strength reduction factor for welded Joints of ASME grades 91, 92 and 122 type steels, Proc.
IIW Int. Conf. on Safety and reliability of welded components in energy and processing
industry, Gratz, Austria, July 2008, Verlag der Technischen Universitaet Graz, 5157.

34.

Boiler and pressure vessel code I rules for construction of power boilers, American Society
of Mechanical Engineers, Table PG 26, 2008.

35.

S. B. Bhoje and P. Chellapandi: Creep fatigue design of FBR components, in Creep-fatigue


damage rules for advanced fast reactor design, Proc. International Atomic Energy Agency
Technical Committee Meet., Manchester, UK, June 1996, IAEA, IAEA-TECDOC-933, 97115.

36.

R. A. Ainsworth and P. J. Budden: Design and assessment of components subjected to


creep, J. Strain Anal., 1994, 29, (3), 201 208.

37.

Technische Regeln fur Dampfkessel (TRD) [TRD210-TRD 301] Carl Heymanns Verley;
1988, Berlin, Ko ln/Berlin Beuth Verlag.

38.

M. A. Gomes, A. Thomas, P. Seliger, S. Concart, B. Pathiraj, P. Veron,A. A. Becker and


T.H.Hyde: Overview of theWELDON EU project, Proc. 5th Int. Conf. on Mechanics and
materials in design, Porto, Portugal, July 2006, University of Porto, 229238.

39.

C. Becht IV: New weld joint strength reduction factors in the creep regime in ASME B31?3
piping, J Press. Vessel Technol., 2006, 128, 4648.

40.

K.-H. Mayer and F. Masuyama: The development of creepresistant steels, in Creepresistant steels, (ed. F. Abe et al.), 2008, Cambridge, Woodhead Publishing, 155173.

41.

J. Schubert, A. Klenk and K. Maile: Determination of weld strength factors for the creep
rupture strength of welded joints, Proc. Int. Conf. on Creep and fracture in high temperature
components design &amp; life assessment issues, London, UK, September 2005, DEStech
Publications, 792805.

42.

S. Holmstrom and P. Auerkari: Predicting weld creep strength reduction for 9%Cr steels,
Int. J. Press. Vessels Pip., 2006, 83, (1112), 803808.

43.

K. Laha, K. S. Chandravathi, P. Parameswaran and K. Bhanu Sankara Rao: Type IV


cracking susceptibility in weld joints of different grades of CrMo steel, Metall. Mater. Trans.
A, 2009, 40A, 386397.

44.

EN 13445-2:2002 (E) Issue 35 (2009-01) Unfired pressure vessels. Materials. Abson and
Rothwell Review of type IV cracking of weldments in 912%Cr creep strength enhanced
ferritic steels International Materials Reviews 2013 VOL 58 NO 8 469

45.

VdTU" V-Merkbla tt Schweisstechnik1153, Schweisstechnik Richtlinien fuer die


eignungspruefung von schweisszusaetzen, 2009.

46.

R. A. Ainsworth: R5 procedures for assessing structural integrity of components under


creep and creep-fatigue conditions, Int. Mater. Rev., 2006, 51, (2), 107126.

47.

F. W. Brust, G. M. Wilkowski, P. Krishnaswamy and K. Wichman: Creep and creep-fatigue


crack growth at structural discontinuities and welds, STP-NU-039, ASTM Standards
Technology LLC, West Conshohocken, PA, USA, 2011.

48.

T. H. Hyde, M. Saber and W. Sun: Creep crack growth data and prediction for a P91 weld
at 650uC, Int. J. Press. Vessels Pip., 2010, 87, 721729.

49.

R. Klueh: Elevated-temperature ferritic and martensitic steels and their application to


future nuclear reactors, ORNL report ORNL/TM2004/176, ORNL, Oak Ridge, TN, USA, 2004.

50.

C. R. Chen, H. Ghassemi Armaki, K. Maruyama, Y. Minami and M. Igarashi: Microstructural


degradation during high temperature exposure up to 105h and its effect on creep of gr 91
steel, Proc. 6th Int. Conf. on Advancesin materials technology for fossil power plants, Santa
Fe, NM, USA, September 2010, EPRI, S2B- 03_33, 15pp. Conference on CD.

51.

H. Danielsen: Z in phase 912%Cr steels, PhD thesis, The Technical University of


Denmark, Copenhagen, Denmark.

52.

H. K. D. H. Bhadeshia: Design of ferritic creep-resistant steels, ISIJ Int., 2001, 41, (6),
626640.

53.

M. E. Abd El-Azim, M. E. Nasreldin, G. Zies and A. Klenk: Microstructural instability of a


welded joint in P91 steel during creep at 600C, Mater. Sci. Technol., 2005, 21, (7), 779
790.

54.

F. V. Ellis and R. Viswanathan: Review of type IV cracking in piping welds, Proc. Int. Conf.
on Integrity of high-temperature welds, Nottingham, UK, 1998, Institution of Mechanical
Engineers, 125134.

55.

L. Milovic: Microstructural investigations of the simulated heat affected zone of the creep
resistant steel P91, Mater. High Temp., 2010, 27, (3), 233242.

56.

F. Abe: Creep deformation behaviour and alloy design philosophy of creep resistant
tempered martensitic 9Cr steel, Proc. 6th EPRI Int. Conf. on Advances in materials
technology for fossil power plants, S2B-01_2, 21pp, conference on CD. Santa Fe, NN, USA,
AugustSeptember 2010, Electric Power Institute.

57.

J. S. Lee, K. Maruyama, I. Nonaka and T. Ito: Mechanism of type IV failure in weldment of


a mod 9Cr-1Mo steel, in Creep deformation and fracture, design and life extension
(materials science & technology), (ed. R. S. Mishra et al.), 139148; 2005, Warrendale, PA,
TMS.

58.

J. A. Francis, W. Mazur and H. K. D. H. Bhadeshia: Estimation of type IV cracking tendency


in power plant steels, ISIJ Int., 2004, 44, (11), 19661968.

59.

K. Kimura, K. Sawada, H. Kushima and Y. Toda: Influence of composition partitioning on


creep strength of high chromium ferritic creep resistant steels, Proc. IOM Conf. on Integrity
of high temperature welds, London, UK, April 2007, IOM Communications, 497506.

60.

K. Kimura, K. Sawada, Y. Toda and H. Kushima: Creep strength assessment of high


chromium ferritic creep resistant steels, Mater. Sci. Forum, 2007, 539543, 31123117.

61.

S. J. Brett, D. L. Oates and C. Johnston: In-service type IV cracking in a modified 9Cr


(Grade 91) header, Proc. ECCC Creep Conf., (ed. I. A. Shibli et al.), 563572; 2005, London,
DEStech Publications Incorporated.

62.

R. Swanekamp: Handling nine-chrome steels in HRSGs, Power Eng., 2006, 110, (2),
available at: http://pepei.pennnet.com/
display_article/249004/6/ARTCL/none/none/1/Handling-Nine- Chrome-Steel-in-HRSGs/

63.

M. J. Cohn, J. F. Henry and D. Nass: Fabrication, construction and operation problems for
grade 91 fossil power components, Trans. ASME J. Press. Vessel Technol., 2005, 127, (2),
197203.

64.

F. Masuyama: Creep degradation in welds of Mod 9Cr-1Mo steel, Int. J. Press. Vessels Pip.,
2006, 83, (1112), 819825.

65.

T. Endo, K. S. Park and F. Masuyama: Proc. Conf. on Case histories in failure investigation,
(ed. V. Bicego et al.), 831840; 1999, Cradley Heath, Materials Advisory Service.

66.

G. Sposito, C. Ward, P. Cawley, P. B. Nagy and C. Scruby: A review of non-destructive


techniques for the detection of creep damage in power plant steels, NDT & E Int., 2010, 43,
(7), 555567.

67.

F. Masuyama: Advance power plant developments and materials experience in Japan,


Proc. 8th Liege Conf., (ed. J. Lecompte- Beckers et al.), Vol. 53, Part I, 175187; 2006, Ju
lich, Forschungszentrum Ju lich GmbH.

68.

K. Laha, K. S. Chandravathi, P. Parameswaran, K. B. S. Rao and S. L. Mannan:


Characterisation of microstructures across the heat-affected zone of the modified 9Cr-1Mo
weld joint to understand its role in promoting type IV cracking, Metall. Mater. Trans A, 2007,
38A, 5868.

69.

F. Abe:: R&D of advanced ferritic steels for 650uC USC boilers, Proc. Int. Workshop on
The innovative structural materials for infrastructure in 21st century, Tsukuba, Japan,
January 2000, Ultra-Steel 2000. National Research Institute for Metals, 119 129.

70.

M. Tabuchi, T. Watanabe, K. Kubo, M. Matsui, J. Kinugawa and F. Abe: Creep crack growth
behaviour in the HAZ of weldments of W containing high Cr steel, Int. J. Press. Vessels Pip.,
2001, 78, (1112), 779784.

71.

S. K. Albert, M. Matsui, T. Watanabe, H. Hongo, K. Kubo and M. Tabuchi: Microstructural


investigations on type IV cracking in a high Cr steel, ISIJ Int., 2002, 42, (12), 14971504.

72.

A. Shibli and F. Starr: Some aspects of plant and research experience in the use of new
high strength martensitic steel P91, Int. J. Press. Vessels Pip., 2007, 84, (12), 114122.

73.

A. Shibli and D. Robertson: Experience with the use of P/T91 steel in power plant boilers
and development of tools for the component integrity/life assessment, Proc. 5th EPRI Int.
Conf. on Advances in material technology for fossil power plants, Marco Island, FL, USA,
October 2007, EPRI, 140.

74.

D. Robertson and A. Shibli: High temperature weld performance of 912Cr martensitic


steels: plant and research experience, Proc. 3rd Int. Conf. on Integrity of high temperature
welds, London, UK, April 2007, IOM Communications, 2534.

75.

A. I. Shibli and K. Coleman: Failures of P91 steel at the West Burton plant in England raise
concerns about the long term behaviour of the advanced steel, Combined Cycle J., 2006,
available at: www.ommi.co.uk/etd/ETD-EPRI-%20P91%20Failures. pdf

76.

S. J. Brett, D. J. Allen and J. Pacey: Failure of a modified 9Cr header end plate, Proc. Conf.
on Case histories in failure investigation, (ed. V. Bicego et al.), 873884; 1999, Cradley
Heath, Materials Advisory Service.

77.

S. J. Brett: Identification of weak thick section modified 9Cr forgings in service, Proc. 3rd
Conf. on Advances in material technology for fossil power plants, Swansea, UK, April 2001,
University of Swansea and EPRI, 343351.

78.

S. J. Brett: The creep strength of weak thick section modified 9Cr forgings, Proc. Baltica V
Int. Conf. on Condition and life management for power plants, Porvoo, Finland, June 2001,
VTT, 3544.

79.

S. J. Brett: Early type IV cracking on retrofit grade 91 steel hardness, Safety and
reliability of welded components in energy and processing industry, International Institute of
Welding (IIW) 62041, 15-77, Proc. IIW Int. Conf., Graz, Austria, 2008.

80.

F. V. Ellis and R. Viswanathan: Review of type IV cracking in piping welds, Proc. 3rd Int.
EPRI Conf. on Welding and repair technology for power plants, Scottsddale, AZ, USA, June
1998, EPRI, 125134.

81.

J. R. Faulds, R. Viswanathan, J. L. Landrum and S. M. Walker: Guidelines for the


evaluation of seam-welded high energy piping, EPRI reports RP8046-04 and RP2819-24,
Electric Power Institute, Palo Alto, CA, USA, 1996.

82.

J. M. Rogers and R. M. Tilley: Standardization of acoustic emission testing of fossil power


plant seam-welded high energy piping, Proc. ASME/JSME Conf. on Pressure vessels and
piping, San Diego, CA, USA, July 2004, ASME/JSME, Paper no. PVP2004-2248, 113-131.

83.

D. J. Abson, J. S. Rothwell and B. J. Cane: Advances in welded creep resistant 912%Cr


steels, Proc. 5th Int. EPRI Conf. on Advances in materials technology for fossil power plants,
October 2007, Marco Island, FL, USA, October 2007, EPRI, 790.

84.

R. M. Goldhoff: Comparison of parameter methods for extrapolating high temperature


data, J. Basic Eng., Dec. 1959, 629644.

85.

J. B. Conway: Stress-rupture parameters: origin, calculation and use; 1969, London, New
York, Paris, Gordon and Breach.

86.

H. E. Evans: Mechanism of creep fracture; 1984, London, Elsevier.

87.

K. Kimura: A method of long-term creep rupture data analysis for high Cr ferritic creep
resistant steel, Proc. Symp. on Ultra super critical steels for fossil power plants, Beijing,
China, 1620 May 2005, 145156.

88.

K. Kimura: Assessment of long-term creep strength and review of allowable stress of high
Cr ferritic creep resistant steels, Proc. Abson and Rothwell Review of type IV cracking of
weldments in 912%Cr creep strength enhanced ferritic steels 470 International Materials
Reviews 2013 VOL 58 NO 8 PVP2005 Conf., Denver, CO, USA, July 2005, PVP2005-71039,
8pp, Pressure Vessels and Piping Division.

89.

B. Wilshire and P. J. Scharning: Long-term creep life prediction for a high chromium steel,
Scr. Mater., 2007, 56, (8), 701704.

90.

B. Wilshire and P. J. Scharning: A new methodology for analysis of creep and creep
fracture data for 912% chromium steels, Int. Mater. Rev., 2008, 53, (2), 91104.

91.

B. Wilshire and M. R. Bache: Cost effective prediction of creep design data for power plant
steels, Proc. 2nd ECCC Creep Conf. on Creep & fracture in high temperature components
design & life assessment issues, (ed. I. A. Shibli and S. R. Houldsworth), 4455; 2009,
Zurich, DEStech Publications Incorporated.

92.

S. J. Williams, M. R. Bache and B. Wilshire: Recent developments in analysis of high


temperature creep and creep fracture behaviour, Mater. Sci. Technol., 2010, 26, (11), 1332
1337.

93.

J. M. Brear, A. Fairman, C. J. Middleton and L. Polding: Predicting the creep life and failure
locations of weldments, Key Eng. Mater., 2000, 171174, 3542.

94.

K. Bell: An analysis of published creep rupture data for modified 9%Cr steel weldments,
TWI confidential research report for Industrial Members 598/1997, TWI, Quincy, IL, USA,
March 1997.

95.

B. Nath and F. Masuyama: Materials comparisons between NF616 HCM12A and TB12M - 1:
dissimilar metal welds, Proc. EPRI/National Power Conf. on New steels for advanced plant up
to 620C, (ed. E. Metcalfe), 114134; 1995, London, Society of Chemical Industry.

96.

T. H. Hyde and W. Sun: Determining creep properties for columnar and equiaxed regions
of a 9CrMoNbV weld metal at 650 deg. C using bulk uniaxial creep test data, Proc. 3rd Int.

Conf. on Integrity of high temperature welds, IOM Communications Ltd (Institute of


Materials), London, UK, April 2007, 139148.
97.

I. J. Perrin and D. R. Hayhurst: Continuum damage mechanics analyses of type IV creep


failure in ferritic steel crossweld specimens, Int. J. Press. Vessels Pip., 1999, 76, (9), 599
617.

98.

R. J. Hayhurst, F. Valkili-Tahami and D. R. Hayhurst: Verification of 3-D parallel CDM


[continuum damage mechanics] software for the analysis of creep failure in the HAZ [heat
affected zone] region of Cr-Mo-V crosswelds, Int. J. Press. Vessels Pip., 2009, 86, (8), 475
485.

99.

M. Bauer, A. Klenk, K. Maile and E. Roos: Numerical investigations on optimisation of weld


creep performance in martensitic steels, in Mathematical modelling of weld phenomena 8,
(ed. H. Cerjak et al.), 409423; 2007, Graz, Technische Universitat Graz.

100. R. Faulkner and Y. Yin: Materials modelling, Presentation given to the European COST
Meeting in Bergamo, Italy, 2008.
101. E. Holzer and H. Cerjak: Prediction of the loss of precipitation strengthening in modern 9
12%Cr steels a numerical approach, Proc. 5th EPRI Conf. on Advances in material
technology for fossil power plants, Marco Island, FL, USA, October 2007, EPRI, 3A-04, 12pp.
102. J. A. Francis, W. Mazur and H. K. D. H. Bhadeshia: Welding procedures and type IV
phenomena, Proc. 7th Int. Conf. on Trends in welding research, (ed. S. A. David et al.),
737742; 2006, Materials Park, OH, ASM International.
103. T. Sato, K. Tamura, K. Mitsuhata and R. Ikura: Improvement of creep rupture strength of
9Cr1MoNbV welded joints by post weld normalising and tempering, Proc. 5th Int. EPRI Conf.
on Advances in material technology for fossil power plants, Marco Island, FL, USA, October
2007, EPRI, 874.
104. K. Coleman and J. Hainsworth: Improved filler metal enables higher-temperature
dissimilar metal welds, Power, 2009, 153, (7), 5861.
105. V. K. Sikka: Method for welding chromium molybdenum steels, US Patent 4, 612, 070,
September 1986.
106. C. Coussement, M. de Witte, A. Dhooge, R. Dobbelaere and E. van der Donckt: High
temperature properties of improved 9%Cr steel weldments, Rev. Soudure, 1990, 46, (1), 58
63.
107. F. Abe and M. Tabuchi: Microstructure and creep strength of welds in advanced ferritic
power plant steels, Sci. Technol. Weld., 2004, 9, (1), 2230.
108. S. K. Albert, M. Tabuchi, H. Hongo, T. Watanabe, K. Kubo and M. Matsui: Effect of welding
process and groove angle on type IV cracking behaviour of weld joints of a ferritic steel, Sci.
Technol. Weld. Join., 2005, 10, (2), 149156.
109. P. Mayr: Weldability of modern 9%C steels for application in USC power plants, PhD
thesis, Graz University of Technology, Austria, 2007.
110. M. Santella: Improving the performance of creep strengthenhanced ferritic steels, Proc.
23rd Annual Conf. on Fossil energy materials, Pittsburgh, PA, USA, May 2009, US
Department of Energy, Office of Fossil Energy, Online proceedings:
http://www.netl.doe.gov/publications/proceedings/09/fem/ index.html, accessed on
28/1/2012.

111. F. Abe, H. Okada, S. Wanikawa, M. Tabuchi, T. Itagaki, K. Kimura, K. Yamaguchi and M.


Igarashi: Guiding principles for development of advanced ferritic steels for 650uC USC
boilers, Proc. 7th Int. Conf. on Materials for advanced power engineering,
Forschungszentrum, Ju lich, Liege, Belgium, September 2002, 13971406.
112. T. Fujita: Current progress in advanced high Cr ferritic steels for high temperature
applications, ISIJ Int., 1992, 3, (2), 175181.
113. T. Fujita: Advances in 912%Cr heat resistant steels for power plants, Proc. 3rd EPRI
Conf. on Advance in materials technology for fossil power plants, (ed. R. Viswanathan et al.),
3365; 2001, London, Institute of Materials.
114. A. Zielinska-Lipiec, K. Spiradek-Hahn, K. Bryla and A. Czyrska- Filemonowicz: Boron in
high chromium steels for USC steam power plant, Proc. 7th Int. Charles Parsons Turbine
Conf. on Power generation in an era of climate change, Glasgow, UK, September 2007,
University of Strathclyde, 489498.
115. F. Abe: High performance creep resistant steels for 21st century power plants, Proc. 1st
Conf. on Super high strength steels, Rome, Italy, November 2005, Associazione Italiana di
Metallurgia, Paper 84, 12pp.
116. F. Abe, M. Tabuchi, M. Kondo and S. Tsukamoto: Suppression of type IV fracture and
improvement of creep strength of 9Cr steel welded joints by boron addition, Int. J. Press.
Vessels Pip., 2007, 84, (12), 4452.
117. C. R. Das, S. K. Albert, J. Swaminathan, A. K. Bhaduri, B. Raj and B. S. Murty:
Improvement in creep resistance of modified 9Cr-1Mo steel weldment by boron addition, IIW
Doc. IX-2336- 10 (formerly IX-C-1015-10).
118. M. Kondo, M. Tabuchi, S. Tsukamoto, F. Yin and F. Abe: Suppressing type IV failure via
modification of heat affected zone microstructures using high boron content in 9Cr heat
resistant steel welded joints, Sci. Technol. Weld. Join., 2006, 11, (2), 216 223.
119. P. Mayr, I. Holzer, F. Mendez-Martin, A. Agu ero, V. Gonza lez, S. Baumgartner and H.
Cerjak: Experience with 9Cr3W3CoVNbBN base material and crosswelds at 650C for
implementation in USC power plants, Proc. NIMS 3rd Symp. on Heat resistant steels and
alloys for high efficiency USC power plants, Tsukuba, Japan, June 2009, National Institute for
Materials Science, available at: http://www.nims.go.jp/hrdg/ USC/programme_June3.html.
120. R. W. Fountain and J. Chipman: Solubility and Precipitation of boron nitride in iron-boron
alloys, Trans. Metall. Soc. AIME, 1962, 224, 599605.
121. K. Sakuraya, H. Okada and F. Abe: BN type inclusions formed in ferritic heat resistant
steel, Energy Mater., 2006, 1, (3), 158166.
122. P. Mayr, I. Holzer, F. Mendez-Martin, M. Albu, S. Mitsche, V. Gonzalez and A. Aguero:
Improved creep and oxidation behaviour of martensitic 9Cr steel by the controlled addition of
boron and and nitrogen, Proc. 9th Liege Conf.: Materials for advanced power engineering,
(ed. J. Lecomte-Beckers et al.), Forschungszentrum, Ju lich, 350361; 2010.
123. H. Danielsen and J. Hald: Martensitic creep resistant steel strengthened by Z-phase,
European Patent WO2008106978 (A1), August 2008.
124. P. de Smet and H. van Wortel: Controlling heat treatment of welded P91, Weld. J., 2006,
85, (6), 4244.
125. Sperko Engineering Services: Welding Grade 91 alloy steel, May 2007, available at:
http://www.sperkoengineering.com/html/ Grade%2091%20R%2010-05.pdf

126. M. Taneike, N. Fujitsuna and F. Abe: Improvements of creep strength by fine distribution
of TiC in 9Cr ferritic heat resistant steel, Mater. Sci. Technol., 2004, 20, (11), 14551461.
127. F. Abe, M. Taneika and K. Sawada: Alloy design of creep resistant 9Cr steel using a
dispersion of nano-sized carbonitrides, Int. J. Press. Vessels Pip., 2007, 84, (12), 312.
128. C. Shiga, A. Gotoh, T. Kojima, Y. Horii, Y. Fukada, K. Ikeuti and F. Matsuda: State of the
art review on the effect of PWHT (postweld heat treatment) on properties of steel weld
metal, Weld. World/Soudage dans le Monde, 1996, 37, (4), 163176.
129. Metrode Products Limited: Welding consumables for P91 steels for the power generation
industry, 2006, Issue 6. Abson and Rothwell Review of type IV cracking of weldments in 9
12%Cr creep strength enhanced ferritic steels International Materials Reviews 2013 VOL 58
NO 8 471
130. J. S. Rothwell: The effect of thermal history during fabrication on the mechanical
properties of weldments in grade 91 creep resistant steel, TWI industrial members report ref:
976/2010, TWI, December 2010.
131. J. S. Rothwell and D. J. Abson: The effect of thermal history during fabrication on the
mechanical properties of weldments in grade 91 creep resistant steel, Proc. 8th Int. Charles
Parsons Turbine Conf., Portsmouth, UK, September 2011, Paper 3.3a, IOM Communications.
132. L. Li, P. Zhu, G. West and R. C. Thomson: The effect of duration of stress relief heat
treatments on microstructural evolution and mechanical properties in Grade 91 and 92 power
plant steels, Proc. 6th Int. Conf. on Advances in materials technology for fossil power plants,
EPRI. 2B-05, 14pp, La Fonda, NM, USA, 2010.
133. R. C. Maclachlan, J. J. Sanchez-Hanton and R. C. Thomson: The effect of simulated post
weld heat treatment temperature overshoot on microstructural evolution in P91 and P92
power plant steels, Proc. 6th Int. Conf. on Advances in materials technology for fossil power
plants, EPRI. 3A-05, 13pp, La Fonda, NM, USA, 2010.
134. T. C. Totemeier and I. J. Perrin: Effect of tempering on microstructure and properties of
grade 91 steel, Proc. 6th Int. Conf. on Advances in materials technology for fossil power
plants, EPRI. 3A-06, 1p abstract, La Fonda, NM, USA, 2010.
135. L. M. Friedman: EWI [Edison Welding Institute]/TWI controlled deposition repair welding
procedure for 125%Cr-0?5%Mo and 225%Cr-1%Mo steels, Welding Research Council
Bulletin 412, 1996, 2734.
136. K. C. Mitchell and J. Tolaini: Development and applications of ferritic cold weld repairs in
the UK power industry, Proc. 6th Int. EPRI [Electric Power Research Institute]-RRAC [Repair
and Replacement Applications Center] Conf. on Welding and repair technology for power
plants, Sandestin, FL, USA, June 2004, EPRI, 840848.
137. J. F. Henry: Growing experience with P91/T91 forcing essential code changes, Combined
Cycle J., 2005, 8, 1017.
138. J. Vekeman and S. Huysmans: Cold weld repair of T91, Proc. Conf. on Safety and
reliability of welded components in energy and processing industry, Graz, Austria,
International Institute of Welding (IIW), July 2008, 353358.
139. K. Natesan, S. Majumdar, P. S. Shankar and V. N. Shah: Preliminary materials selection
issues for the next generation nuclear plant reactor pressure vessel, Argonne National
Laboratory Report ANL/EXT-06-45, Argonne National Laboratory, Lemont, IL, USA, September
2006.

140. Challenges and solutions in repair welding for power and processing plants proceedings
of a workshop, Welding Research Council Bulletin 412, 1996.
141. W. K. C. Jones and P. J. Alberry: A model for stress accumulation in steels during welding,
Proc. Int. Conf. on Residual stresses in welded construction and their effects, November
1977, London, The Welding Institute, 1526.
142. J. A. Francis, H. K. D. H. Bhadeshia and P. J. Withers: Welding residual stresses in ferritic
power plant steels, Mater. Sci. Technol., 2007, 23, (9), 10091020.
143. K. C. Mitchell and S. J. Brett: Development of modified 9Cr cold weld repair techniques,
Proc. 20th EPRI Conf. on Welding and repair technology for power plants, Marco Island, FL,
USA, June 2012, EPRI, F12, 107-120.
144. K. Coleman: Development and use of weld repair methodologies in the USA, OMMI, 2003,
2, (1).
145. K. Coleman and D. Gandy: Alternative filler materials for DMWs [dissimilar metal welds]
involving P91 materials, Proc. 5th EPRI Int. Conf. on Advances in materials technology for
fossil power plants, Marco Island, FL, USA, October 2007, EPRI, Paper 5A-03.
146. A. K. Bhaduri, S. K. Rai, T. P. S. Gill, S. Sujith and T. Jayakumar: Evaluation of repair
welding procedures for 225Cr-1Mo and 9Cr-1Mo steel welds, Sci. Technol. Weld. Join., 2001,
6, (2), 8992.
147. R. Sturm, M. Jenko and B. Ule: Accelerated test determination of creep properties for
9%Cr-1%Mo steel weldments at different PWHT (postweld heat treatments) conditions, Proc.
3rd Int. HTDA (High Temperature Defect Assessment) and Integrity Conf. on Integrity of high
temperature repair welds Lisbon, Portugal, September 2002, Instituto de Soldadura e
Qualidade (ISQ), 259265.
148. A. Klenk, S. Issler, I. A. Shibli and J. A. Williams: Some characteristics of weld repair for
creep applications, OMMI, April 2003, 2, (1).
149. S. Issler, A. Klenk, A. A. Shibli and J. A. Williams: Weld repair of ferritic welded materials
for high temperature application, Int. Mater. Rev., 2004, 49, (5), 299324.
150. S. J. Brett, D. J. Abson and R. L. Jones: The repair welding of power plant without postweld heat treatment, Proc. Int. Conf. on Integrity of high-temperature welds, Nottingham,
UK, November 1998, Professional Engineering Publishing Limited, 5363.
151. J. F. Henry and K. Bezzant: Successful weld repair of the CSEF steels, Proc. 10th Int. EPRI
Conf. on Welding and repair technology for power plants, Marco Island, FL, USA, June 2012,
EPRI, F8, 7589.
152. J. Storesund and L. Samuelson: Creep life assessment of pipe girth weld repairs with
recommendations, in Creep life of pipe girth weld repairs, OMMI, December 2002, 1, (3).
153. Y. F. Yin, R. G. Faulkner, P. F. Morris and P. D. Clarke: Modelling and experimental studies
of alternative heat treatments in Steel 92 to optimise long term stress rupture properties,
Energy Mater.: Mater. Sci. Eng. Energy Syst., 2008, 3, (4), 232242.
154. P. F. Morris, P. D. Clarke and U. A. Sachadel: Design of heat treatments for 912%Cr
steels to optimise creep resistance for power plant applications, Proc. 9th Liege Conf. on
Materials for advanced power engineering, Forschungszentrum, Ju lich, Liege, Belgium,
September 2010.
155. M. A. Roman, V. Bicego, G. Fedeli, B. Brown, H. van Wortel, N. Taylor and E. Ayala:
Optimisation of dissimilar metal weld procedures for high temperature design, Proc. Conf. on

Advanced heat resistant steels for power generation, (ed. R. Viswanathan and J. Nutting),
644654; 1999, London, IOM Communications Ltd.
156. D. Allen: High temperature performance of dissimilar P91 to P22 [Cr-Mo] steel weldments,
Proc. 5th Int. EPRI [Electric Power Research Institute]-RRAC [Repair and Replacement
Applications Center] Conf. on Welding and repair technology for power plants, Point Clear,
AL, USA, June 2002, EPRI Paper P2.
157. Recommended practices for welding of chromium-molybdenum steel piping and tubing,
AWS D10?8-96, American Welding Society, Doral, FL, USA, 1996.
158. J. Siefert and J. Shingledecker: Temperbead repair of T91 using EPRI P87 filler metal,
Proc. 10th EPRI Conf. on Welding and repair technology for power plants, Paper F13, Marco
Island, FL, USA, June 2012, EPRI.
159. Metrode Products Limited: P87 Technical profile, available at:
http://www.metrode.com/en-gb/home.htm
160. J. A. Siefert, J. M. Sanders, J. M. Tanzosh, W. F. Newell Jr, and J. P. Shingledecker:
Development of EPRI P87 solid wire, Mater. High Temp., 2010, 27, (3), 243252.
161.

VdTU V-Merkblatt 1153.

162.

European Creep Collaborative Committee (ECCC) Recommendations, 2003.

163. K. Okamoto, H. Kitagawa, J. Kusumoto, A. Kanaya and T. Kobayashi: Development of


electric discharge equipment for small specimen sampling, Int. J. Press. Vessels Pip., 2009,
86, (9), 633636.
164. D. J. Allen and S. T. Fenton: A hardness-based creep rupture model for new and service
aged P91 steel, Proc. Int. Conf. on Life management and maintenance for power plants
BALTICA VII. Helsinki-Stockholm-Helsinki, Finland, June 2007. VTT, 156 170.
165. D. J. Allen: A hardness-normalised model of creep rupture for P91 steel, Proc. 2nd Int.
ECCC Conf., Dubendorf, Switzerland, April 2009, ECCC, 659668.
166. C. E. Jaske and R. Viswanathan: Remaining-life prediction for equipment in high
temperature/pressure service, Mater. Perform., 1991, 30, (4), 6167.
167. B. Neubauer and U. Wedel: Restlife estimation of creeping components by means of
replicas, in Advances in life prediction methods, (ed. D. A. Woodford and J. R. Whitehead),
307; 1983, New York, ASME.
168. T. Izaki, T. Kobayashi, J. Kusumoto and A. Kanaya: Creep life assessment method for
boiler pipes using small punch creep test, Int. J. Press. Vessels Pip., 2009, 86, (9), 637642.
169. V. Mentl and V. Liska: Remaining life assessment of steam turbine boilers, Proc. Baltica VII
Conf. on Life management and maintenance for power plants, (ed. J. Veivo and P. Auekari),
106112; 2007, Helsinki, VTT.
170. J. Parker: Life management of creep strength in ferritic steel in boilers and piping, Proc.
Baltica VII Conf. on Life management Abson and Rothwell Review of type IV cracking of
weldments in 912%Cr creep strength enhanced ferritic steels 472 International Materials
Reviews 2013 VOL 58 NO 8 and maintenance for power plants, (ed. J. Veivo and P. Auekari),
2036; 2007, Helsinki, VTT.
171. C. Delle Site, C. de Petris and C. Mennuti: NDT tools for life assessment of high
temperature pressure components, Proc. 9th European NDT (ECNDT) Conf., Berlin, Germany,

September 2006, Deutsche Gesellschaft fuer Zerstoerungsfreie Pruefung e. V.2529


September 2006, paper 7.2.4.
172. A. Morris, A. Puri, C. Maharaj, M. Kourmpetis, M. Sjodahl and J. P. Dear: Predicting
lifetimes of components in power station engineering plant, Proc. Int. Baltica VII Conf. on
Life management and maintenance for power plants, (ed. J. Veivo and P. Auekari), 7081;
2007, Helsinki, VTT.
173. K. Penny and W. Kohlho fer: Some neglected NDE tools, Proc. ECCC Conf. on Creep and
fracture in high temperature components, London, Uk, September 2005, ECCC, 326335.
174. R. D. Roberts: Managing creep strain in steam reformer tubes using laser-based NDT,
Proc. 60th Annual Conf. and Exposit., Houston, TX, USA, April 2005, NACE International,
CORROSION 2005. 05404, 43pp.
175. R. J. Viswanathan, J. R. Foulds and D. A. Roberts: Methods for estimating the temperature
of reheater and superheater tubes in fossil boilers, Proc. Int. Conf. on Life extension and
assessment, The Hague, The Netherland, June 1988, Nederlands Instituut voor Lastechniek
paper No 1.6.2.
176. R. J. Pardikar, N. Ayodhi and P. S. Subrahmanyam: Regaining lost capacity of aging boiler
components by performance evaluation and residual life assessment: an exciting challenge
for NDE, Proc. Conf. on Trends in NDE science and technology, Proc. 14th World Conf. on
Nondestructive testing (14th WCNDT), New Delhi, India, December 1996, Paper 369.
177. D. K. Bhattacharya: NDT for residual life assessment in Indian industry, Insight, 2003, 45,
(1), 3140.
178. R. Viswanathan, R. Purget and U. Rao: Materials technology for advanced coal power
plants, 2005, available at: www.msm.com. ac.uk/phase-trans/2005/link/188.pdf
179. J. Hald: Metallography and alloy design in the COST 536 action, Proc. 8th Liege Conf.:
Materials for advanced power engineering, (ed. J. Lecomte-Beckers et al.), Vol. 53, Part 2,
917930; 2006, Ju lich, Forschungszentrum Ju lich GmbH.
180. D. G. Robertson and S. R. Holdsworth: ECCC data sheets, OMMI, 2005, available at:
http://www.ommi.co.uk/etd/eccc/ open.htm.
181. Rothwell J S and Abson D J 2012: The effect of thermal history during fabrication on the
creep performance of weldments in grade 91 creep-resistant steel, Creep 2012, Proc. 12th
International Conference on Creep and Fracture of Engineering Materials and Structures, 27
31 May 2012, Kyoto Terrsa, Kyoto, Japan, Japan Institute of Metals, 2012. Abson and
Rothwell Review of type IV cracking of weldments in 912%Cr creep strength enhanced
ferritic steels International

You might also like