Anda di halaman 1dari 79

Creep Performance and Microstructural Characterization

of the Type IV Region in Grade 91 Steel Weldments

Final Report

MSE 489: Senior Capstone Design

Department of Materials Science and Engineering

The University of Tennessee

Submitted by:

Christopher Hobbs, Maverick Echivarre,

Carly Jania and Stephen Whitson

May 4, 2015

Project Advisor: Professor Carl Lundin, University of Tennessee


Industry Contact: John Siefert, Electric Power Research Institute
CONTENTS

Abstract………………………………………………………………………………………... 1
1 Introduction……………………………………………………………………………….. 2
1.1 Energy Production………………………………………………………………….. 2
1.2 Steam Power Plants………………………………………………………………… 2
1.3 Failures in Steam Piping………………………………………………………….... 2
1.4 Grade 91 Steel Alloy………………………………………………………………... 4
1.5 P91 Grade Seam-Welded Steam Piping Failures…………………………………. 9
1.6 Heat Affected Zone (HAZ)…………………………………………………………11
1.6.1 General HAZs……………………………………………………………….. 11
1.6.2 General HAZ Creep Behavior……………………………………………….11
1.6.3 Adjacent Base Metal………………………………………………………… 15
1.6.4 Grain Morphology across the P91 HAZ……………………………………. 17
1.6.5 Subgrain Morphology across the P91 HAZ………………………………... 17
1.6.6 Carbide Morphology across the P91 HAZ…………………………………. 17
1.7 Creep………………………………………………………………………………...22
1.7.1 Creep Mechanism…………………………………………………………… 22
1.7.2 Creep-Rupture Modeling…………………………………………………….24
1.7.3 Reduced Creep Strength in Grade 91 Weldments………………………….. 24
1.7.4 Type IV Cracking in Grade 91 Weldments…………………………………. 26
1.7.5 Longitudinal vs. Transverse Loading………………………………………. 27
1.8 Submerged Arc Welding (SAW)………………………………………………….. 30
1.9 Postweld Heat Treatment (PWHT)………………………………………………. 30
1.10 Carbides……………………………………………………………………………. 33
1.10.1 General Nature/Occurrence in Steels……………………………………... 33
1.10.2 Chemical Composition in P91…………………………………………….. 36
1.10.3 Effects on Creep Performance…………………………………………….. 36
2 Experimental Procedures……………………………………………………………….. 37
2.1 Objectives…………………………………………………………………………... 37
2.2 Materials…………………………………………………………………………….37
2.2.1 Postweld Heat Treatments…………………………………………………... 37
2.3 Metallography……………………………………………………………………… 39
2.4 Creep Testing………………………………………………………………………. 39
2.5 Hardness Testing……………………………………………………………………42
2.5.1 Hardness Traverse…………………………………………………………... 42
2.5.2 Rockwell Hardness………………………………………………………….. 42
3 Results & Discussion…………………………………………………………………….. 44
3.1 Microstructural Characterization………………………………………………... 44
3.1.1 Chemical Composition……………………………………………………… 44
3.1.2 Base Metal Hardness………………………………………………………... 44
3.1.3 Macrographs………………………………………………………………… 44
3.1.4 Hardness Traverse…………………………………………………………... 49
3.1.5 FGHAZ and ICHAZ Microstructure……………………………………….. 49
3.1.6 Carbide Morphology………………………………………………………… 54
3.2 Creep Performance Characterization……………………………………………. 57
3.2.1 Transverse Ductility………………………………………………………….57
3.2.2 Longitudinal Ductility………………………………………………………. 57
3.2.3 Transverse Creep Behavior…………………………………………………. 61
3.2.4 Longitudinal Creep Behavior………………………………………………..61
3.2.5 Transverse vs. Longitudinal Properties…………………………………….. 61
3.2.6 Larson-Miller Plot…………………………………………………………... 63
3.3 Additional Solutions……………………………………………………………….. 65
3.3.1 Weldment Plateaus………………………………………………………….. 65
3.3.2 Weld Strength Reduction Factors…………………………………………... 65
3.3.3 Seamless Steam Pipes……………………………………………………….. 65
4 Conclusions………………………………………………………………………………. 68
4.1 Future Work……………………………………………………………………….. 69
5 Acknowledgments………………………………………………………………………... 70
References……………………………………………………………………………………. 71
Abstract

The work presented in this study was performed with the intent to characterize the
primary creep-failure mechanism of the 9Cr-1Mo-V alloy steel classified as P91, responsible for
pipe rupture and steam escapes that have occurred in steam power plants. Specifically, it aims
utilize microstructural and mechanical characterization of the material to propose a viable
7
method of reducing such creep failure in seam-welded P91 piping. A 1 8” thick P91 plate

containing a submerged arc weld of P91 filler material served as the test specimen. The viability
of a postweld heat treatment (PWHT) in reducing creep deformation in the weldment was
evaluated by treating each sample extracted from the plate with one of three unique PWHTs.
Cross-sectional samples of the weldment were prepared via metallography for optical and
scanning electron microscopy. Microstructural analysis of these focused on grain and carbide
morphology in the heat affected zone (HAZ), an area in the weldment that previous literature
claims to be the most susceptible to creep damage. Creep samples were extracted from the plate
in both longitudinal and transverse orientations: the former served to test creep performance as a
function of stress applied normal to the weldment cross-section, and the latter to test creep
performance as a function of stress applied across the various zones of a weldment. These
samples were machined, then tested under creep conditions designed to mimic the stresses of
P91 in steam piping service, but at higher temperatures and stress loads to induce failure within
the time restraints of the study. An 8-hr sub-critical PWHT at 1425°F and a 64-hr sub-critical
PWHT at 1425°F were determined to generally soften all regions (especially fine-grained
regions) of the weldment, and reduce the creep lifetime of P91 before failing in the HAZ. This
behavior was attributed to the coarsening of carbides concentrated along grain boundaries (but to
a greater degree in the 64-hr samples), thus weakening the material and allowing for a higher
strain rate in creep testing. A normalizing and tempering treatment (N&T; 2-hr normalizing at
1950°F, 3-hr tempering at 1425°F) was determined to offer a higher and more uniform hardness
across all weldment regions. The N&T also increased creep lifetime threefold over the other
PWHTs, and an increase in material ductility allowed for elongation before failure in the base
metal away from the HAZ. This increase in creep strength was attributed to dislocation pile-up
along needle-like grains of the tempered microstructure and the more uniform dispersion of
smaller carbides throughout the P91 matrix than in the 8-hr and 64-hr PWHTed microstructures.

1
1 Introduction

1.1 Energy Production

In 2013, the U.S. generated over 4 trillion kilowatt-hours of energy, with a projected
increase in demand for the following decade. Current energy production primarily utilizes coal
and natural gas, with nuclear power and renewable sources (e.g. wind, water) contributing to a
lesser degree. The combustion of fossil-fuels accounts for 67% of all energy production as
shown in Figure 1. With nearly 7,000 power plants in operation, fossil-fuels remain the dominant
source for energy production in the U.S.[1].

1.2 Steam Power Plants


Power-plants convert the chemical energy stored in the hydrocarbons of fossil fuels into
the thermal energy used to boil water and generate steam. This steam is funneled through a
network of piping and used to rotate turbines to power electric generators. Many of the
components involved in the transportation of steam are subject to elevated temperatures and high
pressures for periods spanning multiple decades, thus increasing their susceptibility to creep
damage. This is most pronounced in the pipes leading from the boiler to the turbines (see Fig. 2)
as steam leaving the boiler can reach over 538°C (1000ºF). Poorly-engineered piping will
succumb to creep damage and rupture under these service conditions, leading to potential loss of
life, capital, and production time.

1.3 Failures in Steam Piping

The materials utilized to create steam piping must resist creep at elevated temperatures.
Traditional materials have been limited to carbon steels containing chromium and molybdenum,
and are valued for their resistance to creep and oxidation. The Mohave power plant of Bullhead
1 1
City, Arizona employed P11 grade (14 Cr - 2 Mo) seam-welded steam piping in its facilities, but

a fatal accident in 1985 caused by the rupture of a P11 grade steel steam pipe resulted in the
deaths of 6 workers and the injury of 10 others. Failure analysis of the ruptured pipe indicated
that creep-damage initiated at the fusion line of a longitudinal weld seam which resulted in
material degradation and subsequent rupture. When the wall thickness could no longer sustain
the pressure created by the steam, the pipe opened longitudinally in a manner resembling a fish

2
Fig. 1. U.S. energy production and consumption in terms of trillion kilowatt-hours, including
future projections[2].

Fig. 2. The Mohave Power Plant with superimposed schematics of the boiler, turbine and control
room to demonstrate their location within the facility[3].

3
mouth (see Fig. 3) and released the steam at high velocity[3].

The accident at Mohave was not an isolated event. In 1986 the Monroe Power Plant of
Detroit Edison, Michigan experienced a similar seam-welded steam pipe rupture due to pipe
degradation[3]. The failed pipe (see Fig. 4) shares the same traits as the Mohave rupture in that
the creep damage initiated at the longitudinal weld seam and then propagated along the seam.
Additional power plants that employed welded Cr-Mo piping experienced the same mode of
steam pipe failure in which piping was found to fail well before the end of its projected life-span.

1.4 Grade 91 Steel Alloy

Seam-welded piping for installation in critical power plant piping systems has been
1 1 1
produced since the early 1940’s. The introduction of the P11 (1 4 Cr - Mo) and P22 (2 4 Cr - 1
2

Mo) ferritic steels in plants in the 1950’s allowed for an increase in steam temperature and a
corresponding increase in thermal efficiency. Grade 91 steel, a modified 9Cr-1Mo-V alloy
originally developed by the Oak Ridge National Laboratory in the late 1970’s, was later
employed to further improve the creep strength of steels used in fast breeder reactors. The
chemical composition limits of Grade 91 steel in accordance with ASME Boiler and Pressure
Vessel Code (BPVC) SA-387[4] are listed in Table 1. Compared to other Cr-Mo steels (e.g. P11
& P22), Grade 91 steel piping (designated as P91) exhibits enhanced creep rupture strength,
yield strength, ultimate tensile strength and toughness. P91’s microstructure is martensitic as
compared to the bainitic structure of P11 and P22 alloy steels, and its superior creep strength
allows for piping to operate at higher temperatures. These properties of P91 offer better lifetime
performance with a thinner wall section. A comparison of the wall section and rupture strength
of P91 to other low alloy steels with the same design specifications is made in Figure 5.

P91 steel is unique due to its high chromium content. It contains 9% Cr and 1% Mo as
compared to the 2.5% Cr found in P22. This additional chromium improves P91’s high-
temperature strength and increases its resistance to oxidation and corrosion. Molybdenum serves
to enhance its resistance to creep. Vanadium and niobium are added to Grade 91 steel to aid in

4
Fig. 3. Image of the Mohave Power Plant accident. The failed pipe section responsible for the
incident is shown to have ruptured in a fish mouth manner along the weld seam[3].

5
Fig. 4. A photograph of the Monroe power plant accident. The crack initiated at the weld seam and ruptured along the seam [3].
6
Table 1. Chemical compositions of various Cr-Mo Steel alloys in terms of weight percent, as
dictated by the ASME Boiler & Pressure Vessel Code[4].

Element P11 (wt%) P22 (wt%) P91 (wt%)

Cr 1.00 - 1.50 1.90 - 2.60 8.00 - 9.50


Mo 0.44 - 0.65 0.87 - 1.13 0.85 - 1.05
C 0.05 - 0.15 0.05 - 0.15 0.08 - 0.12
Mn 0.30 - 0.60 0.30 - 0.60 0.30 - 0.60
P 0.025 max 0.025 max 0.020 max
S 0.025 max 0.025 max 0.010 max
Si 0.50 - 1.00 0.50 max 0.20 - 0.50
V - - 0.18 - 0.25
N - - 0.030 - 0.070
Ni - - 0.40 max
Al - - 0.040 max
Nb - - 0.06 - 0.10
W - - < 0.01
O - - 0.0002

7
Fig. 5. Cr-Mo Alloy comparisons by pipe wall thickness and rupture strength by operating
temperature[5].

8
the formation of carbides in the matrix, thus increasing the hardness and creep resistance of the
steel[6]. The ratio of nitrogen to aluminum in the steel is maintained near the optimal 2.5:1 to
avoid low ductility cracking[7]. As a result of these alloying elements, P91 exhibits a higher
thermal conductivity, better resistance to stress corrosion cracking, a lower thermal expansion
coefficient, and better resistance to thermal fatigue. These properties account for its widespread
use in industry as a replacement for P11 and P22 steels[8].

1.5 P91 Grade Seam-Welded Steam Piping Failures

While P91 received ASME approval in 1984, its implementation in industry was slow at
first. In response to the major catastrophes in Mohave and Monroe, many methods were
developed and implemented to inspect and repair defects in longitudinal seam-welded piping.
For a time, this approach appeared to have solved the problem at hand, yet the early 1990’s
brought a significant increase in seam-welded steam pipe failures. The Mount Storm Power Plant
near Mount Storm Lake, West Virginia experienced four steam pipe ruptures within a four year
period (1992-1996), causing damage and halting power production each time[9]. The fish-mouth
opening behavior seen in Mount Storm pipes (e.g. that of Fig, 6) is reminiscent of the Mohave
and Monroe failures, despite efforts to avoid such an event.

These later failures led to a fairly rapid implementation of P91 grade steel in the power
industry, as it was viewed to be the most viable solution to the problem of creep susceptibility in
steam piping. Unfortunately, P91 is more sensitive to variations in chemistry and heat treatment
than are P11 and P22, thus failures in P91 seam-welded piping have continued to occur over the
last two decades. All these failures were determined to be products of Type IV creep rupture
(discussed in Section 1.7.4), with the majority of having occurred in the U.K. since P91 has been
in service there longer than in the U.S. At least six instances of Type IV failures in P91 are
known to have occurred in a West Burton plant of the U.K. China has experienced three separate
pipe ruptures due to failures of seam welds in P91, resulting in a total of six worker deaths[10].
These failures are often thought to have been caused by either impurities in the material from
welding, incorrect or impure base materials, or improperly applied heat treatments. It has been
demonstrated in a laboratory setting, however, that creep strength performance is similar
between the parent material and the weld material of P91 until the onset of Type IV

9
Fig. 6. A photograph of the ruptured P11 steam pipe involved in the 1992 Mount Storm accident[9].
10
failure in the HAZ of the weldment. Furthermore, though it was determined that neither material
exhibited impurities or suffered from improper chemistry or heat treatment, Type IV failure still
occurred[10].

1.6 Heat Affected Zone (HAZ)

1.6.1 General HAZs

The HAZ is the region in the weldment that does not melt but is altered in microstructure
and material properties due to the cyclic heating and cooling of the welding process. It lies
between the fusion line of the weld deposit and the base metal, as shown in Figure 7. Although
the HAZ pictured here has experienced no phase change during the welding process, it contains a
microstructure dissimilar to that of the base metal and is thus marked by a lighter, noticeably
different etching pattern.

Within the HAZ there are regions of distinct microstructure that form due to the peak
temperatures reached in the welding process. The HAZ contains a coarse-grained region
(CGHAZ) near the fusion line, a fine-grained (FGHAZ) region in the center and a fine-grained
inter-critical region (ICHAZ) near the base metal (see Fig. 8). These regions are created when
the localized heat from welding exceeds the Ac1 critical temperature, transforming the steel
microstructure from martensite to austenite[12]. The degree of microstructural change depends on
the maximum heat exposure and the cooling rate, both of which are dependent on the distance
from the fusion line (see Fig. 9). These variable factors allow for a wide range of microstructure
across the HAZ.

1.6.2 General HAZ Creep Behavior

Certain regions of the HAZ are particularly susceptible to creep deformation due to their
fine-grained microstructure. The FGHAZ and ICHAZ contain smaller grains, thus these regions
have a greater grain boundary area than other regions in the weldment. Since creep functions by
the mechanism of vacancy diffusion, the effect of minute intergranular voids that diffuse to grain
boundaries and cause grain dislocation is more pronounced in the FGHAZ and ICHAZ[12].

11
Fig. 7. Macrograph of a P91 SAW weldment cross-section at 1.4X. The weldment received a PWHT (774°C/1425°F, 8 hrs), and the
microstructure depicts the lighter HAZ adjacent to the fusion line of the weld seam[11].
12
Fig. 8. General weld seam cross-section schematic depicting the location and appearance of the HAZ relative to the weld deposit and
base metal. The HAZ contains a course-grain region (CGHAZ) near the fusion line, a fine-grain region (FGHAZ) at its center, and an
inter-critical region (ICHAZ) near the base metal.
13
Fig. 9. A schematic of the weldment regions as they correspond to the Fe-C binary phase diagram[13].
14
An analysis of creep-rupture in the HAZ requires an understanding of the different types
of cracking that occur therein. These types of weldment cracking are depicted in Figure 10, and
are categorized according to where they initiate, propagate and terminate. Type I damage
involves crack initiation and termination within the weld deposit. Type II features crack initiation
in the weld deposit and termination in the HAZ or the base metal. Type III involves crack
initiation in the CGHAZ, which then typically propagates through the beads of the weld deposit.
The Type IV mode of failure is characterized by cavities and concentrated strain in the FGHAZ
or ICHAZ, which propagates along grain boundaries in these regions to the surfaces of the weld.
Cavities form around precipitates (the majority residing on prior-austenite grain boundaries), and
will grow and align parallel to the direction of strain over time at elevated temperatures, thus
forming cracks. These cracks can then propagate through the weldment and lead to failure of the
weld[10]. Types I and II do not pertain to this investigation since it consists of studying HAZ
damage, and Type III is not of interest as it comprises a coarser-grained microstructure[15].

1.6.3 Adjacent Base Metal

The welding process alters the mechanical properties of the base metal adjacent to the
HAZ by exposure to temperatures above the Ac1 and Ac3 during the welding process. This
increases the diffusion rate for interstitial and substitutional alloying elements, thus allowing
carbon and others to migrate within the matrix and form carbides. The reduced concentration of
carbon in the matrix creates soft regions in the base metal immediately adjacent to the HAZ.
These regions characterized by a lower hardness than the unaffected base metal may cause the
weld to fall beneath the minimum strength requirement dictated by ASTM Standard A387[4]. Soft
regions are therefore undesirable, as sharp reductions in strength typically initiate failures.
Hardness tests traversing the weldment are often performed to determine the location and extent
of a soft region within a weldment. Lengthy subcritical PWHTs have been shown to reduce the
overall material hardness across a weldment, while normalizing and tempering a weldment offers
a more uniform hardness distribution[16].

15
Fig. 10. Weldment schematic, depicting the various modes of cracking relevant to non-
destructive evaluation and creep-performance testing. (a) Top longitudinal view and (b)
transverse cross-section of the weld interface[14].

16
1.6.4 Grain Morphology across the P91 HAZ

The welding process also creates varying grain-sizes across the base metal and three
regions of the HAZ. Figures 11 - 14 show the microstructure across these four regions. The
CGHAZ (Fig. 11) contains the largest grains (ASTM GS No.7), followed by the base metal (Fig.
14, ASTM GS No. 9) and the FGHAZ (Fig. 12) and ICHAZ (Fig. 13) (both ASTM GS No. 11).
During the welding process the CGHAZ is exposed to peak temperatures that are well above the
normalizing temperature [1066°C (1950°F)], allowing for grain growth. The FGHAZ is exposed
to peak temperatures near the normalizing temperature [1066°C (1950°F)] and above the Ac3
[1010°C (1850°F)], allowing for austenite formation and grain-structure refinement[17]. The
ICHAZ is exposed to peak temperatures between the Ac1 [857°C (1575°F)] and the Ac3 [1010°C
(1850°F)]. The rapid air-cooling after welding allows the austenite in the FGHAZ and ICHAZ to
form a fine grain microstructure. Parts of the base metal are exposed to sub-critical temperatures
(<Ac1); this exposure does not cause a phase change, but it promotes diffusion of interstitial and
substitutional elements.

1.6.5 Subgrain Morphology across the P91 HAZ

The welding process also creates varying sub-structures across the HAZ and base metal.
The CGHAZ (Fig. 11) contains martensitic lath substructures. Figures 12 - 14 show a reduction
in the presence of martensitic lath substructures in the FGHAZ and ICHAZ as well as in the base
metal adjacent to the HAZ. This reduction is caused by insufficient temperatures and heating
time during the welding process[15].

1.6.6 Carbide Morphology across the P91 HAZ

The cyclic heating and cooling of the welding process imparts varying carbide
morphologies across the regions of the HAZ and the base metal. Figures 15 - 18 depict the
typical carbide morphology across the HAZ and Figure 18 shows that of the base metal prior to
welding. Carbides in the CGHAZ are dissolved during the welding process due to high peak
temperatures and they maintain dissolution due to infrequent cooling cycles. This accounts for
the absence of carbides in the micrograph of said region (Fig. 15).

17
Fig. 11. Optical light micrograph of the CGHAZ in a P91 grade weldment at 1000X. The sample
received a subcritical PWHT (774°C/1425°F, 8 hrs), and the microstructure depicts grains of
ASTM GS No. 7[15].

Fig. 12. Optical light micrograph of the FGHAZ in a P91 grade weldment at 1000X. The sample
received a subcritical PWHT (774°C/1425°F, 8 hrs), and the microstructure depicts grains of
ASTM GS No. 11[15].

18
Fig. 13. Optical light micrograph of the ICHAZ in a P91 grade weldment at 1000X
magnification. The sample received a subcritical PWHT (774°C/1425°F, 8 hrs), and the
microstructure depicts grains of ASTM GS No. 11[15].

Fig. 14. Optical light micrograph of the base metal for a P91 grade weldment at 1000X
magnification. The sample received a subcritical PWHT (774°C/1425°F, 8 hrs), and the
microstructure depicts grains of ASTM GS No. 9[15].
19
Fig. 15. TEM micrograph of the CGHAZ in a P91 grade weldment[15].

Fig. 16. Typical carbide morphology in the FGHAZ of a P91 grade weldment[15].

20
Fig. 17. Typical carbide morphology in the ICHAZ of a P91 grade weldment[15].

Fig. 18. Typical carbide morphology in the base metal of a P91 grade weldment that was
normalized (1066°C/1950°F, 2 hrs) and tempered (774°C/1425°F, 3 hrs)[15].

21
Carbides in the FGHAZ and ICHAZ, however, are not completely dissolved due to the low peak
temperatures they experience. Partially dissolved carbides coarsen as welding continues, thus
accounting for the larger carbides in Figures 16 and 17 in comparison to those in the base metal
(Fig. 18).

1.7 Creep

1.7.1 Creep Mechanism

Materials in high temperature applications are subject to creep-induced damage and


failure. Creep is defined as plastic deformation of a material under constant load and elevated
temperatures over time[18]. The specific mechanisms behind creep deformation are dislocation
glide, dislocation creep, diffusion creep, and grain boundary sliding[19]. Dislocation glide
describes dislocation movement due to thermal activation, while dislocation creep is movement
due to vacancy and interstitial diffusion. Diffusion creep is caused by the flow of vacancies and
interstitials due to an applied stress, and grain-boundary sliding is the movement of grains due to
vacancy diffusion. The frequency of these creep mechanisms is miniscule at low temperatures
(due to insufficient thermal energy), but creep deformation becomes significant above 40% of
the absolute melting temperature (Tm) of a material[19].

A creep-rupture curve indicates how a material behaves under creep conditions. In a


typical creep curve (see Fig. 19), the primary or transient creep region is characterized by a
continuously decreasing creep rate. This strain rate deceleration indicates increasing creep
resistance and strain hardening. The secondary region indicates steady-state creep and is
characterized by a constant strain rate. Secondary creep is marked by a net equilibrium between
strain-hardening, recovery, and softening. The tertiary region is characterized by an accelerated
strain rate which eventually leads to rupture. This rupture occurs as a result of cavity formation,
grain boundary separation, and finally crack formation[20].

22
Fig 19. Schematic of a typical creep curve of strain versus time at constant stress and temperature. The minimum creep rate is the
slope of the linear segment in the secondary region (𝜀̇ = d𝜀/dt). Rupture lifetime tr is the total time to rupture[20].
23
1.7.2 Creep-Rupture Modeling

Great research effort has been expended towards understanding, predicting and
preventing creep ruptures. Due to the time-dependent nature of such research and the high cost of
long-term testing, many engineers to turn to data projection as a solution. Simple linear
extrapolation of data points proves unreliable in accurately predicting creep behavior. In fact,
research has shown that creep exhibits rate-process behavior, generalized by Equation 1 where
deformation rate 𝜀̇ is dependent upon a constant A, activation energy Q(S), the gas constant R,
and absolute temperature T [18].

𝑄(𝑆)
𝑑𝜀̇ ( )
= 𝐴𝑒 𝑅𝑇 (1)
𝑑𝑡

The creep rate of a material can be predicted if elevated temperatures do not constitute a
phase change, meaning a low carbon steel must not exceed the Ac1 temperature in service or
during testing[21]. Models have been developed to predict the time-to-rupture for a material.
Equation 1 is used to derive the Larson-Miller relation by which creep behavior can be
characterized[18]. The Larson-Miller relation inspired the Larson-Miller Parameter (LMP)
(calculated by Equation 2) which takes into account the time-to-rupture tr, a material constant C
(~30 for P91), and the relevant service temperature (in Kelvin or degrees Rankine)[18].

𝐿𝑀𝑃 = 𝑇(𝐶 + 𝑙𝑜𝑔 (𝑡𝑟 ) ) (2)

Figure 20 presents a typical Larson-Miller plot in which the logarithmic applied stress (log σ) is
plotted as a function of the LMP. The linear relationship created on a Larson-Miller plot can be
used to predict the time-to-rupture for a material by extrapolation from known data.

1.7.3 Reduced Creep Strength in Grade 91 Weldments

Decreased creep resistance in P91 steel is influenced by HAZ grain size, carbide
evolution, carbide morphology, and sub-structures present in the FGHAZ and ICHAZ. The
diffusion of vacancies is one of the principle mechanisms of creep deformation, and is highest

24
Fig. 20. Larson-Miller plot for various P91 grade steel weldments. The creep behavior of the entire weldment (long bold line) is
compared to that of just the HAZ (short bold line)[18].
25
along grain boundaries. Accumulation of vacancies at the grain boundaries leads to grain
boundary sliding, another principle creep deformation process. Microstructures consisting of
small grains will contain a higher grain boundary area and therefore higher susceptibility to grain
boundary sliding[22]. The FGHAZ and ICHAZ exhibit decreased creep resistance due to their fine
grain size.

Carbide evolution in the FGHAZ and ICHAZ reduces the creep resistance of P91 grade
steel. As stated in Section 1.6.6, the welding process dissolves the carbides in the CGHAZ and
coarsens the carbides in the FGHAZ and ICHAZ. Carbide dissolution contributes towards solid-
solution strengthening in the matrix and carbide coarsening depletes the matrix of alloying
elements. The FGHAZ and ICHAZ exhibit decreased creep resistance due to the reduced solid
solution strengthening in the surrounding matrix[15].

The carbide morphology in the FGHAZ and ICHAZ reduces the creep resistance of P91
grade steel. The incoherent matrix-carbide interface exhibits a high affinity for vacancy
diffusion. Coarsened carbides contribute a higher interface area and a higher affinity for vacancy
diffusion[23]. The accumulation of vacancies at the carbide interface creates cavities that can
initiate crack formation and subsequent creep rupture. The FGHAZ and ICHAZ exhibit
decreased creep resistance due to increased cavitation.

The grain substructures in the FG and ICHAZ reduce the creep resistance of P91 grade
steel. As stated in Section 1.6.5, the martensitic microstructure in the HAZ contains lath
substructures. The CGHAZ contains many needle-like lath structures due to the cyclic nature of
the high peak temperatures and cooling segments of the welding process. The FGHAZ and
ICHAZ contain fewer such structures due to their experiencing lower peak temperatures during
welding. Lath structures contribute to high-angle grain boundaries that promote dislocation pile-
up[20]. Vacancy diffusion and dislocation motion are inhibited in the presence of dislocation pile-
up. The FGHAZ and ICHAZ therefore exhibit decreased creep resistance due to a reduced
degree of this dislocation pile-up.

1.7.4 Type IV Cracking in Grade 91 Weldments

Creep rupture occurs in P91 seam-welded steam piping due to reduced creep strength in
the FGHAZ and ICHAZ. Equation 3 is a combination of the power law and the Arrhenius
26
equation that characterizes steady-state creep rate 𝜀̇ as a function of applied stress σ, a constant
B, and the creep-rate exponent n[24].

𝑑𝜀̇
= 𝐵𝜎 𝑛 (3)
𝑑𝑡

Creep rupture tests performed on P91 steel have characterized the weld deposit, HAZ and base
metal regions. Figure 21 shows the creep-rate constants for the HAZ regions and the base metal.
As indicated in this plot, the FGHAZ and ICHAZ are the most susceptible to creep damage
because they have the highest creep-rate exponents[15].

Type IV cracking occurs through the initiation and propagation of cracks in the FGHAZ
and ICHAZ. It initiates due to the proliferation and coalescence of cavitation in these regions
during service. As discussed in Section 1.7.3, the FGHAZ and ICHAZ contain coarsened
carbides that exhibit higher vacancy accumulation due to their increased interface. Cavities
eventually form around the carbides. Microcrack formation may occur if grain boundary cavities
coalesce, and is assisted by grain boundary sliding. Figure 22 shows the proliferation and
coalescence of multiple cavities to form microcracks. The steam pressure in piping systems can
cause formed microcracks to grow and initiate Type IV cracking. These cracks then propagates
through weak microstructure in the P91 weldment. Figure 23 shows crack propagation parallel to
the weld deposit. As also mentioned, the microstructure in the FGHAZ and ICHAZ are
weakened due to grain boundary sliding, reduced solid-solution strengthening, cavitation, and
uninhibited dislocation motion. A Type IV crack will easily initiate and propagate within the
FGHAZ and ICHAZ due to cavitation and weakening of the microstructure.

1.7.5 Longitudinal vs. Transverse Loading

The creep behavior of a material, especially composites, will vary depending on the
orientation of applied stress. As each zone of P91 exhibits different mechanical properties, each
creep sample may be treated as a composite. A weldment loaded in the transverse direction will
behave as a parallel composite, and one loaded in the longitudinal direction will behave as a
series composite (see Fig. 24). All regions of a series composite experience an isostress
condition, yet the region with the fastest creep rate will dominate.

27
Fig. 21. Creep rate exponents representative of the HAZ regions and base metal[15].

Fig. 22. Micrograph of a P91 grade steel HAZ at 1000X. The microstructure is marked by heavy
cavitation[3].

28
Fig. 23. Macrograph (a) and micrograph (b) of Type IV crack initiation and propagation through
the FGHAZ and ICHAZ of an electron beam welded P91 grade steel[10].

Fig. 24. Diagrams representing (a) the transverse (series) loading orientation and (b) the
longitudinal (parallel) loading orientation of composite materials[25].

29
Parallel composites, however, experience an isostrain condition so the slowest creep rate will
dominate. The creep-strain rates of these orientations are presented as Equations 4 and 5, where 𝜀̇
is the strain rate and V the volume fraction of each composite region[20]. Analysis of both
transverse and longitudinal strain is required to characterize the full range of creep behaviors for
a particular weldment and subsequent PWHT.

1 1
= ∑𝑖 (4)
𝜀̇ 𝜀𝑖̇ 𝑉

𝜀̇ = ∑𝑖 𝜀̇𝑖 𝑉 (5)

1.8 Submerged Arc Welding (SAW)

Submerged arc welding (SAW) is an economical joining method characterized by a thick


layer of powdered flux which covers the weld pool during the welding process (see Fig. 25). The
mound of flux protects the weld pool from atmospheric contamination (e.g. by nitrogen and
oxygen) that can form mechanically detrimental oxides. Marked by a high deposition rate, lack
of smoke and ease of automation, SAW is desirable for applications in the joining of steel plate
and tubing such as those required by an industrial steam plant[15]. SAW welds are often further
processed by a PWHT meant to reduce creep susceptibility in the HAZ.

1.9 Postweld Heat Treatment (PWHT)

As the creep behavior data for P91 grade weldments is insufficient to accurately predict
creep life, this study aims to employ a short-term creep test to further characterize the creep
behavior of P91 grade weldments. A PWHT serves to improve the weld’s mechanical integrity
by reducing residual stresses introduced by the welding process. The general thermal processing
of a P91 weld specimen will follow the ramp cycle shown in Figure 26, with a PWHT below the
critical temperature (sub-critical PWHT) after sufficient cooling of the weldment[27].

A lengthy sub-critical PWHT can potentially increase the creep-resistance of P91 grade
weldment by manipulating the carbide size, morphology, and distribution. A successful sub-
critical PWHT is designed to aid in carbide precipitation and dispersion, thus improving the
ductility and toughness of the alloy and reducing Type IV deformation and failure[27]. Without
such a treatment, welded P91 specimens readily experience cavitation and cracking,

30
Fig. 25. Schematic of the submerged arc welding (SAW) process. This welding method utilizes a continuous metal electrode which
doubles as the source of the arc and the filler metal[26].
31
Fig. 26. The thermal ramp-cool processing cycle of a P91 weldment as prescribed by industrial member Vallourec[27].
32
potentially leading to the same failure modes prevalent in welded P11 and P22 steam piping (see
Fig 27). As dictated by the 2013 ASME BPVC, a P91 PWHT must be performed within the
temperature range of 732 – 774°C (1350 - 1425°F)[28]. Because P91 is a hypoeutectoid alloy by
composition, the sub-critical PWHT selected for the present study will be performed below the
lower critical temperature 𝐴𝐶 1 (857°C/1575°F), at 774°C (1425°F). This will allow for adequate
diffusion and carbide coarsening without forming any austenite phase[17, 29]. The duration of the
PWHT is dictated by the thickness of the weld piece (generally 1 hour per 25 mm)[17].

Normalizing and tempering is a common postweld thermal processing method used to


normalize varying grain sizes created by welding and temper the normalized microstructure.
ASME Boiler and Pressure Vessel Code SA-335 recommends normalizing at a temperature
between 1038°C and 1079°C (1900 - 1975°F) for 2 hours, allowing the weldment to air-cool to
room temperature, then tempering between 732°C and 799°C (1350 - 1470°F) for 3 hours[28].
These steps allow the microstructure of the FGHAZ to fully austenitize above the Ac3 before it is
air-cooled to form martensite and then tempered below the Ac1[17]. Figure 28 illustrates the
change this heat treatment has on the rupture behavior of a P91 weldment sample, where (a) has
received a sub-critical PWHT only, and (b) has received an additional N&T modification. It is of
note that the HAZ is barely visible in the N&T sample (“Ghost-Like” HAZ), and the fracture
behavior takes place well away from the HAZ. Slight elongation of the creep sample suggests
that this secondary thermal processing increases ductility across the weld. The N&T-modified
sample (Fig. 28b) was reported to experience a tripling of its creep lifetime[11].

1.10 Carbides

1.10.1 General Nature/Occurrence in Steels

The formation of martensite in alloy steels requires a quenching of the metal. This
imparts an altered microstructure resulting in improved mechanical properties. This rapid cooling
transforms the face-centered cubic crystal structure of austenite to body-centered tetragonal[20].
While the slow cooling of austenite allows for cementite (Fe3C) or ferrite (α-Fe) to precipitate
out of the austenite matrix, the rapid change in crystal structure during quenching inhibits the
diffusion of carbon atoms.

33
Fig. 27. Mt. Storm U1 1996 failed P11 main steam pipe weld, no PWHT. This P11 failure
displayed is representative of the same Type IV failure mode observed in P91 weldments[9].

34
Fig. 28. Ruptured creep samples of P91 steel welds, (a) after a sub-critical PWHT
(774°C/1425°F, 8 hrs) and (b) after normalizing (1066°C/1950°F, 2 hrs) and tempering
(774°C/1425°F, 3 hrs) treatment. The creep test evaluated both samples at 621°C (1150°F) and
103.4 MPa (15 ksi). The former sample ruptured in the HAZ after 610.3 hours while the latter
ruptured in the base metal after 1818.8 hours[11].

35
The cooling rate here may in fact be fast enough so as to retain part of the austenite matrix [20].
Martensite is most readily identified by its plate or needle-like grain morphology. It is also
known for its severe lack of ductility, thus any applications seeking to profit from the great
strength of martensite must utilize it in its tempered form[30]. The tempering process heats the
metal to increase ductility and toughness, but also causes carbon-containing particles or carbides
to gradually precipitate out of the supersaturated solid solution of the matrix, especially at
temperatures exceeding 99°C (210°F)[31].

1.10.2 Chemical Composition in P91

The carbides present in P91 steel are primarily those containing the alloying elements of
the steel--Cr, Mo, and V. Treatment of the steel at elevated temperatures (above half of the
absolute melting temperature) aids in the dispersion of these elements throughout the material
bulk, but it is also at these temperatures that carbides precipitate out of the matrix. As treatment
temperatures increase, Cr, Mo and V atoms are incorporated into the carbides, which then
coarsen and segregate along grain boundaries with extended time at elevated temperature (e.g.
upwards of 499°C/930°F)[32, 33].

1.10.3 Effects on Creep Performance

A comprehensive study of the creep performance of a P91 grade weldment must relate
the creep performance to the microstructure, while considering the processing conditions. To this
end, an in-depth analysis of the present carbides must be implemented. In a 1991 study, it was
demonstrated in unserviced steam piping of 1Cr-0.3Mo-0.25V that vanadium carbides are plate-
like and abundant along dislocation lines, leaving 300-500 nm wide precipitate-free pockets. In
a sample of the same alloy held at 538°C (1000°F) for 100,000 hours under a hoop stress of 50
MPa, microstructural analysis revealed the inclusion of fine cuboid molybdenum-rich carbides
within those formerly carbide-free pockets. The existing vanadium carbides did indeed coarsen
but maintained their crystallographic orientation. Furthermore, creep rupture strength was
observed to decrease by 15%[34]. While an appreciably higher Cr content in P91 is certain to
produce different results, it is clear that specific morphological and compositional changes in the
carbides must be considered to assess the creep performance of the alloy.

36
2 Experimental Procedures

2.1 Objectives

The experiments of this study aim to determine the effects of three different heat
treatments on: (1) the microstructure and grain size in the HAZ, (2) creep behavior and time-to-
rupture of samples normal to the weldment cross-section (longitudinal) and spanning (transverse)
the weldment, (3) the hardness variation across the weld deposit, HAZ, and base metal, (4) the
hardness of the base metal, and (5) the morphology and evolution of cavities and carbides within
the microstructure. Results of the creep and hardness testing, as well as an analysis of
microstructure, were used to recommend an optimal PWHT condition that increases the creep-
resistance of P91 weldments.

2.2 Materials

The weld qualification plate, fabricated by the Babcock & Wilcox Company (B&W P91
7
plate), is a 18 inch thick P91 plate joined by 44 SAW welding passes. The B&W weldment (see

Fig. 29) received an 8-hr sub-critical PWHT at 774ºC (1425ºF) after fabrication.

2.2.1 Postweld Heat Treatments

The samples extracted from the P91 weldment for purposes of metallography and creep
testing (see Sections 2.3 and 2.4, respectively) were placed into three experimental groups.
Samples in the first group received no additional PWHT beyond that of the original 8-hr
treatment at 774ºC (1425ºF). Samples in the second group received an additional 56-hr sub-
critical PWHT at 774ºC (1425ºF) for 64 hours total of post-weld thermal processing. Samples in
the third group were normalized (2-hr austenization at 1066°C /1950°F) and then tempered for 3
hours at 774ºC (1425ºF).

37
Fig. 29. Sectioning plan for extraction of metallography samples and longitudinal creep specimens from the B&W P91 weldment.
38
2.3 Metallography

Metallographic techniques were used to reveal the microstructure of each sample for
further analysis and an examination of creep-induced changes to grain morphology, cavitation
and carbide evolution. Samples were ground on a grinding wheel using silicon carbide abrasive
discs and polished using a diamond compound and alumina slurry. The samples were etched
using Vilella’s Reagent (1 g picric acid, 5 mL HCl, 100 mL ethanol) for 20-50 seconds to reveal
the microstructure. This polishing-etching procedure was repeated 3-4 additional times to ensure
the samples’ true microstructure had been revealed[35]. Optical microscopy was used to examine
the microstructure, while scanning electron microscopy was used to analyze carbide morphology
and evolution.

2.4 Creep Testing

Six cylindrical rod-like blanks (approx. 3 ½” long, ⅞” Ø) were extracted longitudinally


from the B&W P91 plate along the fusion line (see Fig. 30) using electrical discharge machining
(EDM). In this way, each extracted sample was a composite containing base metal, the HAZ and
weld deposit in the gauge diameter, allowing for isostress or isostrain testing of all three primary
weldment regions. The blanks were divided into three categories: one to be tested in the as-
received condition (8-hr PWHT at 774°C/1425°F), one to be tested after receiving an additional
PWHT for 56 hours at 774°C (1425°F) (for a total of 64 hour PWHT), and one to be tested after
receiving a normalizing and tempering treatment (1066°C/1950°F for 2 hours and 774°C/1425°F
for 3 hours, respectively). In addition to the longitudinal samples, two transverse creep sample
blanks were extracted from the B&W P91 plate (Fig. 31). One transverse sample was heat
treated for an additional 56 hours (for a total of 64) at 774°C (1425°F) and the second was saved
for future studies.
The blanks were machined to their final shape (3.025” long; gauge 1.40” long with
0.357” Ø) after heat treatment, so as to remove any surface contaminants (e.g. an oxide layer or
brass deposit from the EDM wire) that may have been introduced in the extraction and heat
treatment processes. Figure 32 shows the creep samples (a) before (cylindrical blanks) and (b)
after machining.

39
Fig. 30. Photograph of the P91 sample depicting the layout of the longitudinal creep sample
extraction locations with respect to the zones of the weldment cross-section.

Fig. 31. Photograph of the P91 sample depicting the layout of the transverse creep sample
extraction locations with respect to the zones of the weldment cross-section.

40
Fig. 32. Photograph of (a) a cylindrical creep sample blank removed by EDM prior to heat treatment or final machining, and (b) the
creep sample after heat treatment and final machining, ready for testing.
41
Creep performance testing of all samples was performed at Materials Testing Corporation in Ann
Arbor, Michigan. Longitudinal samples were tested at 621°C (1150°F) and 18 ksi. The 64-hr
PWHT transverse sample was tested at 621°C (1150°F) and 15 ksi to allow results to be
compared to those of an 8-hr PWHT sample and an N&T P91 sample tested in a previous
study[11].

2.5 Hardness Testing

2.5.1 Hardness Traverse


In order to identify the soft regions in the weldment according to heat treatment,
microhardness tests which spanned the weld deposit, HAZ and base metal were performed
perpendicular to the fusion line of the polished (but unetched) cross-sectional samples. Figure 33
shows an example of the location of the microhardness indentations with respect to the weldment
in a cross-sectional sample.

2.5.2 Rockwell Hardness

ASTM Standard A387 dictates that P91 steel have a tensile strength between 85 and 110
ksi[36]. To verify that the PWHTs studied did not cause the material samples to fall outside this
range, hardness was tested in the base metal samples of each of the three experimental sample
groups (8-hr PWHT at 774°C/1425°F; 64-hr PWHT at 774°C/1425°F; 2-hr normalizing at
1066°C/1950°F and 3-hr tempering at 774°C/1425°F). Hardness was evaluated using a 1/16” Ø
tungsten carbide ball indenter on the B scale for Rockwell Hardness (HRBW) per the procedure
outlined in ASTM Standard E18[37]. The obtained HRBW values were used to approximate the
tensile strength of each sample according to conversions listed in ASTM A370[38].

42
Fig. 33. Schematic of the hardness traverse tests performed in relation to the weldment,
perpendicular to the fusion line.

43
3 Results & Discussion

3.1 Microstructural Characterization

3.1.1 Chemical Composition

Prior to mechanical testing of the as-received weldment, verification of chemical


composition was required to determine if the sample conformed to the elemental limits specified
in ASME SA-387 for grade 91 alloys. Table 2 summarizes these results. The elemental weight
percentages fell within the allowable ASME ranges, though slight variation in compositions from
the 1999 chemical analysis report provided with the plate. This may be attributed to differences
in instrumentation or localized chemical variation among different areas of the plate, thus
chemical analysis determined the plate to be P91 steel.

3.1.2 Base Metal Hardness

Table 3 summarizes the results of the base metal hardness tests of each heat treatment
under investigation. The material hardness and converted tensile strength were found to decrease
as the duration of the sub-critical PWHT increases (64-hr vs. 8-hr). Decreased tensile strength is
not ideal in materials for application in steam pipes. This HRBW variation is slight, however,
and all three heat treatments yielded material compliant with ASTM A387 as all strength values
fell within the prescribed ultimate tensile strength (UTS) range. A lower UTS would require
lower steam pressures in power plants, thus decreasing energy efficiency. The N&T sample
exhibited the highest HRBW and tensile strength values and is most appropriate for use in
enhancing the hardness and strength in P91 for steam pipe application.

3.1.3 Macrographs

The applied heat treatments altered both the macrostructure and subsequent
microstructure of each sample. Macrographs of the P91 weldment cross sections are included as
Figures 34 - 36. The 8-hr PWHT (Fig. 34) and 64-hr PWHT (35b) samples exhibit very similar
macrostructures. Both feature HAZs 2-3 mm wide, distinct weld beads and a distinguishable
fusion line. This suggests that applying a longer sub-critical PWHT (64 hours) does not
significantly alter the microstructure and subsequent mechanical

44
Table 2. Chemical composition of the weldment compared to the ASME SA-387 T/P91
specification[4].

Base Metal
ASME SA-387
EPRI, 2015 Weld Deposit
Element Composition
(wt% unless (wt%)
Limits (wt%)
indicated)
C 0.105 0.11 0.08 - 0.12
Mn 0.43 0.57 0.30 - 0.60
Si 0.323 0.15 0.20 - 0.50
Cr 8.70 9.39 8.00 - 9.50
Mo 0.91 0.90 0.85 - 1.05
V 0.211 0.22 0.18 - 0.25
Nb 0.069 0.034 0.06 - 0.10
Ni 0.14 0.74 0.40 max
Ti 0.002 0.002 -
Co 0.014 0.018 -
Cu 0.115 0.04 -
Al 0.028 0.015 0.040 max
B 0.0003 < 0.001 -
W 0.004 < 0.01 < 0.01
N 0.0488 0.051 0.030 - 0.070
S 0.002 0.01 0.010 max
P 0.009 0.012 0.020 max
Sb 0.001 (10 ppm) < 0.002 -
As 0.0066 (66 ppm) 0.01 -
Sn 0.0116 (116 ppm) 0.004 -
Zr <0.001 < 0.001 -
Pb 0.4 ppm - -
Bi <0.1 ppm - -
O 0.0034 0.045 0.0002

Table 3. Base metal hardness for P91 grade weldment samples subjected to varying heat
treatments. *Strength values converted from hardness values using ASTM A370[38].

N&T 8 hr 64 hr

HRBW 94.63 92.00 90.31

SD (±) 0.46 0.24 0.33

Strength (ksi)* 100 92 89

Suitable for Service? ✓ ✓ ✓

45
Fig. 34. Macrograph of a P91-grade weldment cross-section sample which received an 8-hr PWHT (774°C/1425°F).
46
Fig. 35. Macrographs of P91-grade weldment cross-section samples which received (a) a N&T treatment (2 hrs at 1066°C/1950°F and
3 hrs at 774°C/1425°F), and (b) a 64-hr PWHT (774°C/1425°F).
47
a

c
Fig. 36. Panoramic optical macrograph of each heat treated P91 cross-section sample at 9X. The microstructure of (a) the 8-hr PWHT
sample is similar to that of (b) the 64-hr PWHT sample, while the microstructure of (C) the N&T sample is distinct from the PWHT
samples, especially in the weld deposit. The N&T microstructure is also more uniform across the base metal, HAZ and weld deposit.
48
properties from those of the 8-hr PWHT. The N&T sample experienced the most significant
change by comparison: the HAZ is indistinguishable from the base metal and the weld beads less
distinct from one another. The more uniform macrostructure across the weld deposit, HAZ and
base metal (Fig. 36) implies a uniform microstructure. Because significantly weak regions
adjacent to stronger regions initiate failure, uniform mechanical properties across a weldment are
ideal. The N&T sample is the most apt of the heat treatments to mitigate the negative effects of
creep-prone microstructure as it imparts a more uniform microstructure across the weldment.

3.1.4 Hardness Traverse

The hardness distributions in Figure 37 exhibit changes in hardness that correspond to


microstructures imparted by the three heat treatments. The 64-hr PWHT does not significantly
increase the hardness distribution of a weldment over that of the 8-hr PWHT, despite the longer
(and more expensive) treatment time. Hardness and tensile strength are ideal for resisting creep
damage and the N&T sample exhibited a higher and more uniform hardness distribution overall.
This uniformity is ideal because local weak regions rupture faster than strong ones. In this
respect, N&T is therefore the most ideal of the three heat treatments for increasing creep-
resistance in P91 weldments.

3.1.5 FGHAZ and ICHAZ Microstructure

The most creep-susceptible HAZ regions must be identified in order to characterize and
attempt to improve the creep-strength of the entire weldment. Figures 38 - 41 catalog the
microstructure across the weldment zones of the 8-hr sub-critical PWHT samples. It is apparent
that the largest grains were found in the CGHAZ, followed by the base metal and then the
FGHAZ and ICHAZ. The latter two regions contain the smallest grains, and are therefore the
most susceptible to creep damage due to grain boundary sliding. Any efforts to improve the
creep-resistance of the weldment must be focused on the FGHAZ and ICHAZ.

The PWHTs affect the microstructure in the HAZ, and can potentially improve creep-
resistance. Figures 42 - 44 demonstrate the variety of microstructures created by the three
PWHTs. Both sub-critical PWHT microstructures are similar despite a treatment time difference
of 56 hours. The N&T sample features a more defined substructure containing martensitic laths.
These lath substructures impart a high-angle grain boundary that promotes dislocation pile-up.

49
350

8-hr PWHT
64-hr PWHT
300
N&T

250
Hardness (HV 500 gf)

200

150

FL
100
WD HAZ BM

50
-1.0 -0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Distance from Fusion Line (mm)

Fig. 37. Hardness profile across all zones of the P91 grade weldment utilizing a 500 gram load on the Vickers Scale. The 8-hr PWHT
& 64-hr PWHT samples were found to be softer and less uniform across the weld deposit, HAZ and base metal, than the N&T.
50
Fig. 38. SEM micrograph of the base metal in a P91 grade weldment given a subcritical PWHT
(774°C /1425°F, 8 hrs). Etched with Vilella’s reagent. 1000X.

Fig. 39. SEM micrograph of the FGHAZ in a P91 grade weldment given a subcritical PWHT
(774°C /1425°F, 8 hrs). Etched with Vilella’s reagent. 1000X.

51
Fig. 40. SEM micrograph of the CGHAZ in a P91 grade weldment given a sub-critical PWHT
(774°C /1425°F, 8 hrs). Etched with Vilella’s reagent. 1000X.

Fig. 41. SEM micrograph of the weld deposit in a P91 grade weldment given a sub-critical
PWHT (774°C /1425°F, 8 hrs). Etched with Vilella’s reagent. 1000X.

52
Fig. 42. Optical micrograph of the FGHAZ in a P91-grade weldment given a sub-critical PWHT
(774°C/1425°F, 8hrs). Etched with Vilella’s reagent. 400X.

Fig. 43. Optical micrograph of the FGHAZ in a P91-grade weldment given a sub-critical PWHT
(774°C/1425°F, 64 hrs). Etched with Vilella’s reagent. 400X.

53
Fig. 44. Optical micrograph of the fine-grain HAZ for a P91-grade weldment that was
normalized for 2 hours at 1066°C (1950°F) air-cooled and tempered for 3 hours at 774°C
(1425°F). Etched with Vilella’s reagent. 400X.

Thermally activated dislocation motion and vacancy diffusion are restricted among dislocation
pile-ups and creep is hindered. The heating cycle of welding eliminates these martensitic laths in
the HAZ. Subcritical PWHT does little to create a strengthened microstructure whereas the N&T
heat treatment reestablishes these critical martensitic laths. The N&T postweld processing
imparts a more creep-resistant microstructure in comparison to the other heat treatments.

3.1.6 Carbide Morphology

The different PWHTs have been shown to affect the creep-resistance of the samples.
Figures 45 - 47 demonstrate the carbide morphology as a function of PWHT. Increasing the
length of the sub-critical PWHT coarsens the carbide sizes in the FGHAZ. Higher carbide-matrix
interface area is detrimental to the creep-resistance of the material due to its higher vacancy
diffusion rates and subsequent cavity proliferation. The N&T heat treatment creates fine, well-
dispersed carbides which increase the creep resistance of the material. The dissolution of
carbides disperses alloying elements into the surrounding matrix, providing solid-solution

54
Fig. 45. SEM micrograph of the FGHAZ for a P91-grade weldment given a subcritical PWHT
(774°C/1425°F, 8hrs). Etched with Vilella’s reagent. 10,000X.

Fig. 46. SEM micrograph of the FGHAZ for a P91-grade weldment given a subcritical PWHT
(774°C/1425°F, 64hrs). Etched with Vilella’s reagent. 10,000X.
55
Fig. 47. SEM micrograph of the FGHAZ for a P91-grade weldment that was normalized
(1066°C/1950°F, 2hrs) and tempered (774°C/1425°F, 3hrs). Etched with Vilella’s reagent.
10,000X

50

45 Base Metal
Weld Deposit
40
Reduction in Area (%)

35

30

25

20

15

10

0 2 4 6 8 10 12
Distance from Rupture (mm)

Fig. 48. A distribution of ductility in the 64-hr sub-critical PWHT transverse P91 sample.

56
strengthening for the P91 sample. The strengthening mechanism in the N&T sample is in sharp
contrast to the that of the sub-critical PWHT samples, as carbide coarsening weakens a material
by depleting the alloying elements from the surrounding matrix. The N&T sample exhibits more
creep-resistant properties as a result of this carbide morphology.

3.2 Creep Performance Characterization

3.2.1 Transverse Ductility

Proper PWHT can improve the creep strength of a weldment. Figure 48 shows the base
metal is more creep resistant than the weldment in the 64-hr sub-critical PWHT sample. The
ductility of all transverse samples was characterized by comparing the creep behavior of this
sample to that of two samples characterized in a previous study[11]. Figure 49 catalogs the
ruptured transverse creep samples. The micrographs therein show that the weldment is more
creep resistant than the base metal. The sub-critical PWHT samples exhibited little reduction in
gauge cross-sectional area prior to failing in the FGHAZ/ICHAZ. The N&T PWHT sample
exhibited significant reduction in area prior to failing in the base metal away from the HAZ.
Significant ductility is therefore ideal, because it offers an early warning mechanism of
elongation prior to rupture. The ductility imparted by the N&T treatment--and by extension the
improved creep properties of the material--is ideal for more adequately processing P91 in creep-
intensive applications.

3.2.2 Longitudinal Ductility

The ductility of the longitudinal samples was characterized by visual analysis of the
ruptured creep samples. Figure 50 collects macrographs of these three samples. The reduction in
gauge cross-sectional area varies with the applied PWHT and the ductility of each individual
weld region is difficult to quantify. A qualitative characterization is offered by analyzing the
degree of deformation at the fracture edge. Macrographs of the sub-critical PWHT samples (Fig.
51 a, b) show a marked reduction in area for the weld deposit and base metal (indicated by the
flared or tapered outer edges), but no such reduction in area for the HAZ. This demonstrates non-
uniform deformation among the weld zones, where the HAZ behaves less like a ductile material
and more like a brittle one. The N&T creep sample experiences more uniform deformation

57
a

c
Fig. 49. A comparison of the rupture location in the (a) 8-hr PWHT[11], (b) 64-hr PWHT and (c)
N&T[11] transverse creep samples.

58
a b c

WD HAZ BM WD HAZ BM WD HAZ BM

Fig. 50. Macrographs of the ruptured P91 grade longitudinal creep samples. The (a) 8-hr PWHT and (b) 64-hr PWHT samples
experienced more reduction in gauge cross-sectional area than the (c) N&T sample.
59
a

Fig. 51. Micrographs of the ruptured longitudinal creep sample ends. The more ductile failure
mode of the weld deposit and base metal than the HAZ in (a) the 8-hr PWHT sample and (b) the
64-hr PWHT sample, is contrasted by the more uniform failure across (c) the N&T sample.

Table 4. Rupture lifetime of transverse samples tested at 621°C (1150°F) and a stress of 15 ksi.

64 hr 8 hr[11] N&T[11]

Time to Rupture (hrs) 401.3 610.3 1818.8

Reduced Area at Rupture (%) 38 38 72.4

60
across the fractured end, and this behavior can extend a material’s time-to-rupture. Based on
these results, the N&T sample will likely exhibit the longest creep life.

3.2.3 Transverse Creep Behavior

The creep behavior of the three transverse samples varied according to PWHT. Table 4
lists the creep lives of these samples, along with the percent area reduction (%RA) as the site of
rupture. Compared to the original 8-hr PWHT condition, the 64-hr PWHT treatment shortened
the creep life of the alloy to 401.3 hrs. The N&T treatment nearly tripled the creep life over the
8-hr PWHT (1818.8 hrs vs. 610.3 hrs in the 8-hr PWHT), and more than quadrupled the creep
life over the 64-hr PWHT. Both sub-critical PWHTs experienced the same %RA while the N&T
sample endured twice the %RA before failure. The N&T treatment imparts more favorable creep
properties as shortened creep life is detrimental to the integrity of seam-welded steam piping.

3.2.4 Longitudinal Creep Behavior

The creep behavior of the longitudinal samples as indicated by microstructural


characterization is explained by analyzing their creep curves. Figure 52 plots the creep behavior
of these samples as a function of time. Both sub-critical PWHT samples experienced high strain
rates and their data curves appear to lack any region of steady-state creep. The 64-hr PWHT
sample exhibited decreased creep life when compared to the 8-hr PWHT sample. The N&T
curve features a prominent steady-state creep-strain region and the longest creep life of the three
(3 times that of the 8-hr PWHT and 6 times that of the 64-hr PWHT). These very creep
behaviors were predicted by the previous analyses of hardness and microstructure. Creep
resistance is improved in the N&T samples due to a higher and more uniform hardness
distribution, finer and more dispersed carbides, and a greater concentration of martensitic lath
substructures.

3.2.5 Transverse vs. Longitudinal Properties

It was confirmed that the mechanical properties of a weld depend on the orientation of
loading due to the nature of composite materials. It was also determined that the results of
longitudinal orientation testing in this study followed the same trend found in the transverse
orientation testing of the previous 1999 study by Lundin et al.[11].

61
25
Creep test conditions: 1150°F, 18 ksi

Ruptured (82.9 hrs) 64 h PWHT @ 1425°F

8 h PWHT @ 1425°F

20 Normalized and tempered

Ruptured (143.1 hrs)


Ruptured (481.5 hrs)
Creep Strain (%)

15

10

0
0 100 200 300 400 500
Test Duration (h)

Fig. 52. Comparison of creep behavior in the longitudinal samples. The samples include one
given an 8-hr PWHT at 774°C (1425°F), one given a 64-hr PWHT at 774°C (1425°F), and one
normalized for 2 hours at 1066°C (1950°F) and tempered for 3 hours at 774°C
(1425°F). Samples were tested at 621°C (1150°F) and a stress of 18 ksi.

62
Because the transverse samples served as series composites, all regions were exposed to the
same stress. The fine-grained microstructure of the FGHAZ and ICHAZ regions allowed for the
fastest creep rates. Thus, the overall failure of the transverse creep samples was caused by the
Type IV deformation which initiated in these regions. The longitudinal samples serve as parallel
composites, with all regions exposed to the same strain but different degrees of stress. The base
metal creep rate was slowest due to a microstructure least modified by welding. This behavior
dominated in the creep performance of the longitudinal samples and extended creep life.
Therefore, a longitudinal weldment (or weld seam which lies parallel to steam flow in power
plants) will experience the slowest creep rate and the offer the longest creep life of the pipe.

3.2.6 Larson-Miller Plot

The Larson-Miller parameter of each sample was calculated by Equation 2 and compared
to a database reported in the same 1999 UT-ORNL study characterizing the effect of heat
treatment on the creep strength of P91 weldments[11]. Figure 53 plots the experimental LMP
values of this study to the ORNL database. The longest black line represents the creep strength
for the 8-hr PWHT weldment, thus the 8-hr sub-critical PWHT LMP of this study falls on said
line. The 64-hr PWHT data point in either stress orientation (longitudinal or transverse) reduces
the creep strength of the material since their LMPs fall below the 8-hr PWHT line. By contrast,
the LMP of the N&T sample indicates an increase in creep strength as the red star lies above and
to the right of the 8-hr PWHT line. This behavior represents the same trend apparent in the
characterizations of hardness and creep performance. These results serve to further validate the
N&T treatment’s potential as an effective tool to enhance the creep performance of P91
weldments. The ability of the N&T to (1) create a microstructure more resistant to Type IV creep
deformation, (2) extend P91 creep life, (3) improve hardness and creep strength throughout the
weldment and (4) offer a creep-rupture warning system, make it the most promising of the three
treatments tested.

63
Fig. 53. Larson-Miller plot for P91 creep sample data. The calculated LMPs for the 8-hr PWHT, 64-hr PWHT and N&T longitudinal
samples as well as that of the 64-hr PWHT transverse sample were plotted with data from a previous study to assess the effect of each
heat treatment and applied stress orientation on material creep strength[11].
64
3.3 Additional Solutions

3.3.1 Weldment Plateaus

A HAZ oriented horizontally (parallel to the direction of primary stress) will be more
creep-resistant than one oriented vertically (normal to this direction). Figure 54 depicts a
weldment schematic containing HAZ plateaus. Such a weldment design may experience a longer
creep-life due to the horizontal segments of the HAZ that are introduced when the piping is
joined by welding. HAZ plateaus only serve to extend the creep life a weldment since creep
damage in the FGHAZ and ICHAZ will still occur and eventual pipe replacement is still
necessary[39].

3.3.2 Weld Strength Reduction Factors

Regulations have been set by the ASME BPVC to reduce the likelihood of steam pipe
rupture. This specifications prescribes weld strength reduction factors that may be employed to
reduce the service pressure for P91 grade seam-welded steam pipes. Table 5 shows the
recommended reduction factors for P91 grade steels as a function of PWHT and service
temperature. The reduction factors prescribed for a sub-critical PWHT do not take account for
the decrease in creep-strength with increasing sub-critical PWHT time. More research must be
conducted with PWHT to determine more accurate reduction factors[40].

3.3.3 Seamless Steam Pipes

Extruded P91 grade steam pipes will be tend to be more resistant to creep than a seam-
welded pipe. Being that the HAZ is the most creep-susceptible region in a seam-welded steam
pipe, Type IV cracking, deformation and rupture can be avoided entirely if the a pipe does not
have a longitudinal seam weldment. An extruded steam pipe would have no such longitudinal
seam, and therefore no HAZ prone to Type IV cracking and failure. Unfortunately, extrusion is
not currently an economical viable option in the construction or maintenance of a steam power
plant, as facilities cable of extruding large diameter steam pipes are few in number and exist only
outside the U.S.

65
Fig. 54. Weld seam cross-section schematic depicting HAZ plateaus designed to offer longer creep life by placing segments of the
HAZ parallel to the direction of primary steam flow stress in application.
66
Table 5. Weld strength reduction factors as supplied in the 2013 ASME BPVC[40].

67
4 Conclusions

1. Sub-critically PWHTed samples (both 8-hr and 64-hr at 1425°F) show no significant
difference in optical microstructure. The subcritical temperature affects the evolution of
carbides but does not affect the grain size.
2. Normalizing 2 hours at 1950°F followed by tempering 3 hours at 1425°F creates a more
uniform microstructure across the weld (eliminates CG, FG, and IC HAZ).
3. An additional sub-critical PWHT (64-hr total) does not significantly affect base metal
hardness compared to an 8-hr PWHT. According to ASTM A387, the material will
therefore be suitable for service regardless of PWHT duration.
4. Hardness is generally more uniform across the weld in the N&T sample, compared to the
8-hr and 64-hr PWHTed samples.
5. Based on results from microstructural and creep performance characterization, PWHT
contributes greatly to creep deformation and shortens rupture lifetime. A longer sub-
critical PWHT (64-hr) reduces the creep resistance compared to the 8-hr PWHT.
6. A N&T treatment significantly improves the rupture lifetime of the weld.
7. The 64-hr PWHT transverse sample exhibits a shorter rupture life than the 8-hr PWHT or
N&T transverse samples (401.3, 612.3, and 1818.8 respectively). The N&T exhibits the
longest rupture life, and the 64-hr PWHT exhibits the shortest rupture life.
8. The LMPs for both the transverse and longitudinal 64-hr PWHT samples fall below the
acceptable trendline and therefore exhibit unsatisfactory creep resistance.
9. The N&T sample also demonstrates significant improvement in creep resistance in the
longitudinal orientation. The 64-hr longitudinal sample ruptured in 81.6 hrs, the 8-hr
longitudinal sample ruptured in 141.6 hrs, and the N&T longitudinal sample ruptured in
477.1 hrs. The rupture life of the N&T sample was 3 times greater than the 8 hr sample
and 6 times greater than the 64 hr sample. The longitudinal samples follow the same
trend as observed in the transverse samples (N&T > 8-hr > 64-hr).
10. Short-term tests yield similar results as long-term transverse tests, which validates their
utility in identifying promising directions toward an optimal heat treatment.
11. When compared to the ORNL weldment/weld stress rupture database, the LMPs of this
study indicate that an N&T treatment exhibits superior long-term creep strength
compared to a sub-critical PWHT in both the longitudinal and transverse orientations.

68
12. Longer exposure to a sub-critical PWHT increases total strain at elevated temperature and
stress. The transverse samples expose the susceptibility of the HAZ to creep.
13. The 64-hr PWHT contained the largest carbides and these were concentrated along grain
boundaries. This contributes to the lower creep resistance of the 64-hr PWHT. The N&T
sample contained smaller carbides that were more evenly distributed throughout the
matrix, contributing to the higher creep strength of this sample.
14. Based on the results from creep testing and microstructural analysis, an N&T treatment is
the optimal heat treatment to increase the creep strength and rupture lifetime of a P91
weldment.

4.1 Future Work

1. Current standards should be changed to reflect the low cross-weld creep strength in P91,
especially for those used in steam service.
2. Long-term testing at conditions similar to steam piping service should be conducted on
N&T samples to more accurately determine expected rupture life.
3. Further study of N&T treatments should be conducted to determine the effects of the
normalizing temperature on the long-term creep strength of P91 welds to determine the
most effective or most practical normalizing temperature.
4. Further study of sub-critical PWHT should be conducted in order to determine the most
effective combination, by analyzing the effects of the temperature and time of PWHT.
5. Carbide analysis may also be used for in-depth analysis of the changes in morphology
and chemical composition carbides can experience as a function of PWHT and how this
affects the creep strength of the weldment.

69
5 Acknowledgments

The authors would like to thank the following individuals for their invaluable
contributions to the success of this study:

Dr. Carl Lundin for his mentorship, direction, vast experience in welding, constructive criticism
and always pressing for higher quality to inspire critical thinking and a dedication to excellence.

John Siefert of the Electric Power Research Institute for his additional characterization and for
addressing industry-related concerns of the project.

Teaching assistants John Bohling and Maneel Bharadwaj for their attention to detail and tireless
assistance with sample preparation and characterization, as well as for their constructive
criticism in numerous drafts and presentations pertaining to this work.

Doug Fielden, Larry Smith and Danny Hackworth of the College of Engineering’s Mechanical
Systems Group for their assistance in sample extraction and machining.

Randy Stooksbury of the College of Engineering’s Facilities Group for his assistance with
obtaining supplies for sample preparation and characterization.

Will Hoskins of the Materials Joining Group for his assistance with sample preparation.

Dr. Josh Burgess for his guidance in sample preparation and the characterization of carbides.

Greg Jones for his patience and experience in characterization by scanning electron microscopy
and energy dispersive spectroscopy.

Drs. Maulik Patel and Jason Fowlkes for their presentation reviews and constructive criticism of
intermediate report drafts.

Carla Lawrence, Martha Gale, Gena Wilson, Tonya Brewer and Tracy Lee of the Materials
Science and Engineering Department Main Office, for their assistance with printing and
obtaining office supplies.

All other professors and faculty of UT’s Materials Science and Engineering Department for their
contribution to teaching and training the next generation of Materials Science Engineers.
70
References

[1] U.S. Energy Information Administration, Monthly Energy Review and Annual Energy
Outlook 2014. <http://www.eig.gov/forecasts/aeo>.
[2] U.S. Energy Information Administration, What is U.S. Electricity Generation by Energy
Source? <http://www.eia.gov/tools/faqs>.
[3] C.D. Lundin. “Steam Piping Failure, a Utility Industry Problem.” Presentation. The
University of Tennessee, Knoxville, Materials Joining Group.
[4] ASME SA-387. “Specification for Pressure Vessel Plates, Alloy Steel, Chromium-
Molybdenum.” ASME Boiler and Pressure Vessel Code, 2013: II, Part A, ASME: New
York, NY.
[5] S.J. Brett. “Identification of Weak Thick Section Modified 9Cr Forgings in Service.”
Swansea Creep Conference Proceedings, April 2001. The University of Swansea, UK.
[6] S. Wilmes, G. Zwick. “Effect of Niobium and Vanadium as an Alloying Element in Tool
Steels with High Chromium Content.” Proceedings of the 6th International Tooling
Conference, 10-13 September 2002, Karlstad, Sweden.
[7] K. Coleman. “Fabrication, Construction, and Operation Issues with Grade 91 Piping and
Tubing.” Electric Power Research Institute Presentation to the Association of Edison
Illuminating Companies, 8 February 2007.
[8] B. Silwal, L. Li, A. Deceuster, B. Griffiths. “Effect of Postweld Heat Treatment on the
Toughness of Heat-Affected Zone for Grade 91 Steel.” AWS Welding Journal,Vol. 92,
March 2013, p. 80-87.
[9] C.D. Lundin. “Failure Modes of Long Seam Welded Cr-Mo High Energy Piping.”
Materials Joining Group. Presentation to the Materials Properties Council, 1992.
[10] D.J. Abson, J.S. Rothwell “Review of Type IV Cracking of Weldments in 9– 12%Cr
Creep Strength Enhanced Ferritic Steels.” International Materials Reviews, Vol. 58,
2013, pp. 437-473.
[11] C.D. Lundin, P. Liu, Y. Cui. “Stress Rupture Behavior of Modified 9Cr-Mo (P91)
Weldment in Sub-Critically PWHT and N&T Condition.” Presentation. The University of
Tennessee, Knoxville, Materials Joining Group. 1999.
[12] J. Parker. “Factors Affecting Type IV Creep Damage in Grade 91 Steel Welds.”
Materials Science and Engineering A, Vol. 578, 2013, pp. 430-437.
[13] J. Indacochea, G. Wang, R. Seshadri, Y. Oh. “Creep Rupture Properties of High-
Temperature Bainitic Steels after weld repair.” Journal of Engineering Materials and
Technology, Transactions of the ASME, Vol. 122, No. 3, 2000, pp. 259-263.

71
[14] H.J. Schuller, L. Haigh, A. Woitscheck. “Cracking in the Weld Region of Shaped
Components in Hot Steam Lines – Materials Investigations’, Der Maschinenschaden,
Vol. 47, No. 1, 1974, pp. 1–13.
[15] C.D. Lundin, P. Liu. “Determination of Creep Behavior of Singular HAZ Regions to
Model the Behavior of the Entire HAZ of Cr-Mo Steels (ASME T/P11, T/P91, T/P92,
T/P122) in Elevated Temperature Service.” WRC Bulletin 475, Welding Research
Council, Inc., New York, NY, USA, 2000.
[16] C.D. Lundin, J. Henning, R. Menon, K. Khan. “Transformation, Metallurgical Response
and Behavior of the Weld Fusion Zone and Heat Affected Zone in Cr-Mo Steels for
Fossil Energy Application.” DOE. ORNL Report. UT Welding Research Society, 1987.
[17] T. Ruglic. “Normalizing of Steel.” Hinderliter Heat Treating, Inc. ASM Metals
Handbook, 2nd Ed., Vol. 6: Welding, Brazing and Soldering, ASM International, 1998.
[18] F.R. Larson, J. Miller. “A Time-Temperature Relationship for Rupture and Creep
Stresses.” Transactions of the ASME, Vol 74, July 1952, pp. 765-775.
[19] G.E. Dieter. Mechanical Metallurgy. 3rd ed. New York, NY: McGraw-Hill, 1988.
[20] W.D. Callister, D.G. Rethwisch. Materials Science and Engineering: An Introduction. 8th
ed. Hoboken, NJ: Wiley, 2010.
[21] A. Mendelson, S.S. Manson. “The Extrapolation of Families of Curves by Recurrence
Relations, With Application to Creep Rupture Data.” Transactions of the ASME, Ser. D:
Journal of Basic Engineering, Vol. 82, 1960, pp. 839-847.
[22] I. Finnie, W.R. Heller. Creep of Engineering Materials. New York, NY: McGraw-Hill,
1959.
[23] C.R. Brooks. “Principles of the Heat Treatment of Plain Carbon and Low Alloy Steels.”
ASM International, 1996, pp. 217-234.
[24] M.A. Meyers, K.K. Chawla. Mechanical Behavior of Materials, 2nd ed. UK: Cambridge
University Press, 1999, p. 573.
[25] T.G. Nieh. “Elastic Constants.” Presentation. The University of Tennessee, Knoxville. 6
Feb. 2014.
[26] J.S. Ogborn. “The Fundamentals of Welding: Submerged Arc Welding.” The Lincoln
Electric Company. ASM Metals Handbook, 2nd Ed., Vol. 6: Welding, Brazing and
Soldering, ASM International, 1998.
[27] G. Guntz, M. Julien, G. Kottmann, F. Pellicani, A. Pouilly, J.C. Vaillant. “5.3. Welding.”
The T 91 Book: Ferritic Tubes and Pipe for High Temperature Use in Boilers, 1990,
Vallourec Industries, France.

72
[28] ASME SA-335. “Specification for Seamless Ferritic Alloy-Steel Pipe for High
Temperature Service.” ASME Boiler and Pressure Vessel Code, 2013: VIII, Div. 2.
ASME: New York, NY.
[29] T. Vurherer, M. Dunđer, L.J. Milović, M. Zrilic, I. Samardžić. “Microstructural
Investigation of the Heat-Affected Zone of Simulated Welded Joint of P91 Steel.”
Metalurgija, Vol. 52, No. 3, 2013, pp. 317-320.
[30] J.D. Verhoeven. “Steel Metallurgy for the Non-Metallurgist.” American Society for
Metals, 2007.
[31] M. Wisti, M. Hingwe. “Tempering of Steel.” ASM Metals Handbook, Vol. 4: Heat
Treating, pp. 291-324. Materials Park, OH: ASM International, 2004.
[32] D.J. Burgess. “Characterization of the Fine-Scale Weld Deposit Microstructure and its
Influence on the Elevated Temperature Properties of 2.25Cr-1Mo-0.25V Weldments in
Heavy Wall Pressure Vessels.” Doctoral Dissertation, The University of Tennessee,
Knoxville, Dec. 2014.
[33] H.-O. Andrén, S. Karagöz, C. Guangjun, L. Lundin, H. Fischmeister. “Carbide
Precipitation in Chromium Steels.” Surface Science, Vol. 246, 1991, pp. 246-251.
[34] R. Singh, S. Banerjee. “Morphological and Compositional Changes of the Carbides in Cr-
Mo-V Ferritic Steel.” Materials Science and Engineering, Vol. A132, 1991, pp. 203-211.
[35] B.L. Bramfitt, S.J. Lawrence. “Metallography and Microstructures of Carbon and Low-
Alloy Steels.” ASM Metals Handbook, Vol. 9: Metallography and Microstructure, pp.
624-625. Materials Park, OH: ASM International, 2004.
[36] ASTM A387. “Specification for Pressure Vessel Plates, Alloy Steel, Chromium-
Molybdenum.” ASTM International, 2015. <http://www.astm.org/Standards/A387.htm>.
[37] ASTM E18. “Standard Test Methods for Rockwell Hardness of Metallic Materials."
ASTM International, 2015. <http://www.astm.org/Standards/E18.htm>.
[38] ASTM A370. “Standard Test Methods and Definitions for Mechanical Testing of Steel
Products.” ASTM International, 2015. <http://www.astm.org/Standards/A370.htm>.
[39] J. Siefert, J. Parker. “Creep Strength Enhanced Ferritic Steels: An Examination of the
Challenges Associated with Grade 91 Steel.” Electric Power Research Institute.
Presentation at ASME Code Week, Bellevue, WA, May 11, 2014.
[40] ASME Table PG-26. “Weld Strength Reduction Factors to be Applied When Calculating
Maximum Allowable Pressure or Minimum Required Thickness of Components
Fabricated With a Longitudinal Seam Weld.” ASME Boiler and Pressure Vessel Code,
2013: Section I. ASME: New York, NY.

73
74

Anda mungkin juga menyukai