Вы находитесь на странице: 1из 10

archives of civil and mechanical engineering 18 (2018) 713–722

Available online at www.sciencedirect.com

ScienceDirect

journal homepage: http://www.elsevier.com/locate/acme

Original Research Article

A brief study on d-ferrite evolution in dissimilar P91


and P92 steel weld joint and their effect on
mechanical properties

Chandan Pandey a,*, Manas Mohan Mahapatra b, Pradeep Kumar a,


Nitin Saini a, Jayant Gopal Thakre a, R.S. Vidyarthy a, H.K. Narang c
a
Department of Mechanical and Industrial Engineering, Indian Institute of Technology Roorkee, Uttrakhand 247667,
India
b
School of Mechanical Sciences, Indian Institute of Technology, Bhubaneswar, Odisha 751013, India
c
Department of Mechanical Engineering, National Institute of Technology Raipur, Chhattisgarh 492010, India

article info abstract

Article history: Ferritic/martensitic 9Cr-1Mo-V-Nb steel also designated as ASTM A335 used in construction
Received 26 September 2017 of several components of power plants operating in temperature range of 600–650 8C. In
Accepted 2 December 2017 present investigation, dissimilar weld joint of P91 and P92 steel were prepared using the
Available online 21 December 2017 autogenous tungsten inert gas (A-TIG) welding and multi-pass gas tungsten arc welding
(GTAW) process. A comparative study was performed on evolution of d-ferrite patches in
Keywords: weld fusion zone and heat affected zones (HAZs) of welded joints. The evolution of d-ferrite
A-TIG patches was studied in as-welded and post-weld heat treatment (PWHT) condition. PWHT
GTAW was carried out at 760 8C for tempering time of 2 h and 6 h, for both A-TIG and GTA weld
Hardness joints. It was observed that presence of higher content of ferrite stabilizer in P92 steel
Charpy toughness promote the formation of d-ferrite patches in weld fusion zone as well as HAZs. To study the
d-Ferrite effect of welding process and PWHT, Charpy V impact energy and microhardness tests were
performed. For microstructure characterization, field-emission scanning electron micro-
scope (FESEM) and optical microscope were utilized.
© 2017 Politechnika Wrocławska. Published by Elsevier Sp. z o.o. All rights reserved.

the weldability of P91 steel is a serious issue. It has been


1. Introduction
reported that, compared to that of weld, a higher creep rupture
life of P91 base metal is observed regardless of the arc welding
Modified 9Cr-1Mo (P91) steels are developed as first wall and process and heat treatment used [2]. It occurs due to presence
blankets structure of ITER test modules [1]. In present years, of non-equilibrium structure in heat affected zone (HAZ). The

* Corresponding author.
E-mail addresses: chandanpy.1989@gmail.com (C. Pandey), manasfme@gmail.com (M.M. Mahapatra),
kumarfme@gmail.com (P. Kumar), nit030078@gmail.com (N. Saini).
https://doi.org/10.1016/j.acme.2017.12.002
1644-9665/© 2017 Politechnika Wrocławska. Published by Elsevier Sp. z o.o. All rights reserved.
714 archives of civil and mechanical engineering 18 (2018) 713–722

non-equilibrium structure formed because of high heating and rate, filler composition, heat input and number of passes play an
cooling rate in the weld thermal cycle. P91 steel can be welded important role in deciding the formation of d-ferrite.
by any of the commonly used arc welding processes such as The d-ferrite formation in weld fusion zone have a
manual metal arc welding (MMA), gas tungsten arc welding noticeable effect on mechanical properties. d-Ferrite formation
(GTAW), shielded metal arc welding (SMAW), electron beam in P91 steel weldments are generally reported at highest peak
welding (EBW), and submerged arc welding (SAW) [3–5]. A lot of temperature [10,11]. The region exposed at temperature more
efforts have been made to develop P91 and P92 steel electrode than 1200 8C leads the formation fine prior austenite grains
that provide optimum strength and toughness to weldments. (PAGs) because of d-ferrite nucleation on preferential sites of
The initial problem faced during welding of P91 steel is prior austenite grain boundaries (PAGBs). The d-ferrite
selection of matching filler wire/rod that provide optimum nucleation limits the growth of austenite and resulting in
composition in weld fusion zone. Composition of Ni and Mn in fine PAGs formation. Chandravathi et al. [11] reported that
filler wire required more attention due to their strong influence formation of soft d-ferrite in microstructure led to reduction in
on temperature Ac1, Mf, and Ms. In P91 base metal, maximum hardness. Schafer [9] studied the effect of d-ferrite formation
1%(Ni + Mn) is allowed. For 1.4% (Ni + Mn), Mf temperature on tensile properties and observed that d-ferrite upto 8% have
decreased lower than 200 8C (preheat temperature) that led to no any significant effect on the strength. However, a different
formation of retained austenite in filler metal. Santella et al. [6] opinion was reported related to effect of d-ferrite on toughness
reported a numerical equation to calculate the Ac1 temperature and ductility. Some author(s) [10,11] have reported positive
based on Mn and Ni contents. The equation is given below: effect of retained d-ferrite on toughness while others [12]
emphasize the negative influence. Schafer [9] reported that
combined effect of d-ferrite and brittle 'dendritic' carbides of
Ac1 ð CÞ ¼ 854:50:643:91ðMn þ NiÞ90:4ðMn þ NiÞ2 (1) M23C6 types (M  65%Cr + 30%Fe + rest) reduced the toughness
by encapsulate the d-ferrite. However, pure and soft d-ferrite
Increase in Ni and Mn content in filler metal decrease the increased the toughness and ductility. Anderko et al. [13] have
Ac1 for weld metal. The V and Nb are the other elements in filler also reported the negative influence of d-ferrite on impact
wire that affects the mechanical properties of weld fusion toughness of fusion zone. Presence of d-ferrite in base metal,
zone. Arivazhaga et al. [7] studied the effect of V + Nb content weld zone or HAZ have found a negative influence on creep
in weld fusion zone on Charpy toughness of weld fusion zone. strength of P91 steel [14,15]. Presence of d-ferrite accelerated
Higher Charpy toughness was observed for the low V + Nb the recovery rate of martensite and growth rate of intermetal-
content in weld zone. Fusion welding process introduces the lic phase like laves phase (Fe2W or Fe2Mo) and ultimately
heterogeneity in microstructure by its nature. Single phase reduced the creep and fatigue strength of steel.
tempered martensitic structure in P91 weld zone provides In present case, effect of A-TIG and GTAW welding process
optimum strength and toughness compared to two phase of on d-ferrite evolution have been studied for as-welded and
tempered martensite and d-ferrite. In P91 weldments, forma- PWHT condition. PWHT was performed at 760 8C for varying
tion of d-ferrite in weld fusion zone lead to reduction in tempering time of 2–6 h. The welded samples were character-
mechanical properties of weldments. A schematic evolution of ized using the Charpy V impact energy test, hardness test and
sub-zone in multi-pass welded P91 joint is shown in Fig. 1. field emission scanning electron microscope (FESEM). For
As per Schaeffler diagram, high chromium and low nickel phase analysis X-ray diffraction analysis was performed.
equivalent in the composition of base metal and weld zone
might be also promote the formation of d-ferrite [8,9]. The d-ferrite
2. Experimental methodology and material
formation occurs during the unconventional mechanical and
thermal treatment. Formation of d-ferrite are reported commonly
in P91 and P92 steel welds joint. In autogenous tungsten inert gas Creep strength ferritic/martensitic P91 and P92 steel were
(TIG) welding process, composition of base metal, heat input and supplied in normalized and tempered condition, as per
cooling rate plays an important role in d-ferrite formation while in manufacturer. Normalizing was performed at 1040 8C for
multi-pass gas tungsten arc welding (GTAW) process, cooling 40 min, followed by air cooling and then tempered at 760 8C

Fig. 1 – Schematic evolution of sub-zones in P91 weld joints (CGHAZ: coarse heat affected zone; FGHZA: fine grain heat
affected zone; ICHAZ: inter critical heat affected zone).
archives of civil and mechanical engineering 18 (2018) 713–722 715

for 2 h and finally air cooled. Schematic microstructure evolution Composition of GTAW filler wire and weld zone for both A-TIG
during the normalizing and tempering process is shown in Fig. 2. and GTA welds are given in Table 1. The process parameters
Normalizing was performed to develop the martensitic for A-TIG process and GTAW root and filling pass are depicted
microstructure with high dislocation density. Tempering in Table 3. Table 4 shows the welding process details for A-TIG
process leads to tempering of martensite with evolution of and GTA welding process.
precipitate that enhance the stability of microstructure. During After the completion of welding post weld heating was
the normalizing, fine Nb-rich MX precipitates are reported to be performed at 250 8C for 60 min to remove the diffusible
undissolved [16], as shown in Fig. 2. Evolution of coarse Fe, Mo hydrogen content. After post weld heating, post weld heat
and Cr-rich M23C6 type precipitates and fine V and Nb-rich MX treatment (PWHT) was performed at 760 8C for varying time of
precipitates occur during the tempering process [17]. Other 2–6 h. For the microstructure characterization, polished
precipitates that observed in P91 and P92 steels are M7C3 [M:Fe, sample etched in Vilella's reagent (1 g picric acid, 5 ml
Mn, Cr], M3C,Cr2N and Fe7W6 [18,19]. Chemical composition of hydrochloric acid, and 100 ml ethanol). For microstructure
as-received P91 and P92 steel is shown in Table 1 [5]. Mechanical characterization, field emission scanning electron microscope
properties of P91 and P92 steel are given in Table 2. (FESEM) with energy dispersive spectroscopy (EDS) and optical
For autogenous tungsten inert gas (A-TIG) welding, plates microscope were utilized. Microhardness tests were performed
of dimension 150 mm  100 mm  5 mm were prepared with- on microhardness tester at load of 500 g and dwell time of 10 s.
out edge preparation. Plates are surface ground on faces to Charpy toughness tests were performed on standard sub-size
make the intimate contact during the butt welding of plates. specimen (55 mm  10 mm  5 mm) as per ASTM E-23. Notch
The A-TIG welding was performed on fully automated was made at the center of the weld fusion zone.
machine in flat (1 G) position. For A-TIG welding process, heat
input was 1.75 kJ/mm.
For GTAW plate of thickness 8 mm with conventional 3. Results and discussion
groove weld design were utilized. Preheating of plates were
performed at 250 8C by using the flame heating while inter- 3.1. As-received material and mechanical properties
pass temperature during the gas tungsten arc welding (GTAW)
process was maintained in range of 200–250 8C. For GTAW Typical micrographs of 'as-received' P91 and P92 steel are
process, heat input varies in range of 0.504–0.655 kJ/mm. shown in Fig. 3(a) and (b), respectively. Microstructure consists

Fig. 2 – Schematic microstructure evolution during normalizing and tempering process.

Table 1 – Chemical composition of P91 pipe, AWSER90S-B9 (9CrMoV-N) filler wire, and weld metal (wt.%) [5].
Element Chemical composition (wt.%)

C Mn Cr Si Mo V Nb Ni S Ti W B Cu Fe
P91 steel 0.12 0.54 8.48 0.28 0.95 0.18 0.05 0.35 0.011 0.012 <0.001 – 0.06 Rest
P92 steel 0.11 0.58 9.09 0.48 0.42 0.24 0.073 0.30 0.003 – 1.87 0.03 Rest
Filler metal 0.12 0.50 8.83 0.30 0.90 0.20 0.06 0.50 0.019 0.001 – – – Rest
A-TIG weld 0.16 0.45 8.90 0.29 0.95 0.23 0.06 0.30 0.010 <0.002 <0.01 – 0.05 Rest
GTA weld 0.11 0.52 8.28 0.28 0.87 0.20 0.04 0.44 0.019 <0.002 <0.001 0.05 Rest

Table 2 – Mechanical properties of P91 and P92 steel in as-received condition [18,20].
Material Yield strength (MPa) Tensile strength (MPa) Elongation (%) Hardness (HV) Charpy V impact energy (J)
P91 steel 475  25 715  15 20  2 247  5 200  8
P92 steel 530  10 678  10 23  2 227  4 198  8
716 archives of civil and mechanical engineering 18 (2018) 713–722

Table 3 – Parameters used for making P91 weldment by GTAW process [5].
Process No. of layerCurrent (I) ampsVoltage (V) voltsTravel speed (T) (mm/min)Arc efficiencyAvg. heat input (kJ/mm)
GTAW root pass 1 110 12 135 0.70 0.410
GTAW filler pass 1–5 120–130 14–18 140–150 0.70 0.504–0.655
A-TIG – 230 14–15 80 0.70 1.75

Table 4 – Summary of welding process used in weld preparation [5].


Parameters A-TIG GTAW
Electrode EW-2%Th EW-2%Th
Electrode tip angle 608 608
Polarity DCEN, water cooled torch DCEN, water cooled torch
Shielding gas Commercially pure argon (99.9%) Commercially pure argon (99.9%)
Weld groove No Groove angle: 758, root height: 1.5 mm, root gap: 1.5 mm
Arc gap 3 mm 3 mm

Fig. 3 – Microstructure of P91 steel and P92 steel in 'as-received' condition (a) P91 steel, (b) P92 steel, (c) X-ray diffraction
analysis.

of martensite equiaxed laths, lath blocks, prior austenite grain observed from the micrographs. M23C6 precipitates derive
boundaries (PAGBs), and lath boundaries. The stability of during the tempering process, as shown in Fig. 2. Nb-rich fine
tempered microstructure is derived from the precipitates that MX precipitates remain undissolved during the austenitizing
decorated along the PAGBs, lath blocks, lath packets and inside process and other V and Nb-rich MX precipitates evolve during
the intra-lath region. Precipitates located along the boundaries the tempering process [23]. X-ray diffraction analysis also
and lath packets having size in range of 50–140 nm are confirm the presence of M23C6 and MX precipitates in P91 and
confirmed as Cr, Fe and Mo-rich M23C6 carbide precipitates P92 steel. Cr, Fe and Mn-rich M7C3, and Cr2N phase are also
while spherical fine precipitates having size in range of 20– observed in diffraction analysis.
45 nm are confirmed V and Nb-rich MX type precipitates [3,21].
Coarse M23C6 precipitates enhance the creep strength of 3.2. Microstructure of weld fusion zone and FGHAZ in as-
material by pinning the movement of subgrain boundaries and welded condition
PAGBs while fine MX precipitates impede the movement of
dislocations and provide the precipitation hardening [22]. Macrostructure of A-TIG and GTAW welds joint is shown in
Higher density of precipitates along the PAGBs are clearly Fig. 4 [5]. Micrographs of dissimilar weld joints for A-TIG and
archives of civil and mechanical engineering 18 (2018) 713–722 717

GTAW process are shown in Fig. 5. In A-TIG process, weld fusion shows the untempered columnar laths with different
fusion zone exhibited the d-ferrite patches beside the area of spatial orientation, as shown in Fig. 6(c).
untempered lath martensite. The d-ferrite have observed
different size and shape. In P91 side fine grain heat affected 3.3. Microstructure evolution during PWHT
zone (FGHAZ), typical columnar lath morphology is observed
inside the lath packets of prior austenite grain boundaries Micrographs of weld fusion zone and FGHAZ for A-TIG and
(PAGBs). The d-ferrite patches of small size is also observed in GTA welds joint after the PWHT of 2 h are shown in Fig. 7. The
FGHAZ. Higher heat input in A-TIG process results in ferrite content was still observed in weld zone. Hence, it can be
dissolution of precipitates but some coarse M23C6 precipitates stated that PWHT did not affect the ferrite content. The PWHT
and fine Nb-rich MX precipitates remain undissolved [24]. In have not observed a significant effect on the weld fusion zone
P92 side FGHAZ, microstructure shows the similar behavior as microstructure except the evolution of precipitates along the
P91 side FGHAZ except the PAGBs in P92 side are observed grain boundaries and inside intra-lath region. The ferrite
more clearly. Lath blocks and packets are also clearly seen region are observed to be free from any type of precipitates.
inside the PAGBs. For GTAW process, weld fusion zone is The grain size of weld fusion zone showed a minute change
characterized with typical columnar lath morphology with and it decreased from 64  25 mm to 62  8 mm. Evolution of
small area of d-ferrite patches. In FGHAZ of P91 and P92, coarse precipitates were also occurred un FGHAZ but at the same time
undissolved M23C6 precipitates are observed. The presence of grain coarsening and undissolved precipitates coarsening
precipitates in FGHAZ is attributed to lower heat input in were also occurred. In as-welded condition, grain size for
GTAW process that is not sufficient to dissolve the precipitates FGHAZ P91 and FGHAZ P92 were measured to be 36  13 mm to
completely. In GTAW, FGHAZ also exhibited the equiaxed 28  7 mm, respectively. After the PWHT of 2 h, reformation of
laths due to auto-tempering of previous pass from the grains resulted a negligible reduction in grain size and it was
subsequent pass. The d-ferrite patches are observed to be measured 35  15 mm to 22  6 mm for FGHAZ P91 and FGHAZ
missing in both P91 and P92 FGHAZ. The optical micrograph of P92, respectively. For GTA welds joint grain size was measured
A-TIG weld fusion zone also confirms the presence of ferrite to be very less. In as-welded condition, it was 16  6 mm to
patches, as shown in Fig. 6(a) and (b) [5]. In GTA welds, weld 12  7 mm for FGHAZ P91 and FGHAZ P92, respectively. However,
grain coarsening was noticed in FGHAZ after the PWHT of 2 h
and grain size was measured 26  6 mm to 24  7 mm for FGHAZ
P91 and FGHAZ P92, respectively. Uniform grain size in FGHAZ of
P91 and P92 side shows the uniformity in microstructure along
the dissimilar GTA welds joint. In A-TIG welds coarse grains of
weld and FGHAZ in as-welded condition are attributed to higher
heat input as compared to GTA welds.
Secondary electron micrographs of weld zone and FGHAZ
for A-TIG and GTA welds are shown in Fig. 8. The d-ferrite
patches are clearly observed in A-TIG welds zone, as shown in
Fig. 4 – Macrostructure of weld joints for (a) A-TIG, (b)
Fig. 8(b). The ferrite boundaries and PAGBs are observed to be
GTAW.
decorated with coarse precipitates while inside the ferrite
region, no precipitates are observed. AT higher magnification,

Fig. 5 – Microstructure in A-TIG weld joint (a) FGHAZ P91 side, (b) weld zone, (c) FGHAZ P92 side; microstructure in GTAW joint
(d) FGHAZ P91 side, (e) weld zone, (f) FGHAZ P92 side.
718 archives of civil and mechanical engineering 18 (2018) 713–722

Fig. 6 – (a) Microstructure of A-TIG weld fusion zone in low and high magnification (a) and (b), respectively; (c) weld fusion
zone of GTA welds.

Fig. 7 – Optical micrographs for A-TIG welds after the PWHT of 2 h (a) FGHAZ P91 side, (b) weld zone, (c) FGHAZ P92 side; GTA
welds after the PWHT of 2 h (d) FGHAZ P91 side, (e) weld zone, (f) FGHAZ P92 side.

distribution and morphology of precipitates are shown. It precipitates are observed along the PAGBS and lath blocks. The
confirmed the coarse globular shape precipitates along the ferrite patches are observed to be absence due to lesser weight
lath blocks and PAGBs while fine MX precipitates in size order percentage of ferrite stabilizer in P91 steel. In P92 side FGHAZ,
of 20–35 nm, inside the matrix region. In GTA welds, weld zone small areas of d-ferrite patches besides the lath martensite are
shows the columnar lath morphology without any presence of observed, as shown in Fig. 8(c). In GTA welds, both P91 and P92
ferrite patches. The precipitates are observed to be along the side FGHAZ show the clear absence of ferrite patches. The
lath blocks or lath boundaries. In FGHAZ, for A-TIG welds joint columnar lath morphology was also observed to be missing.
both columnar and equiaxed lath are observed. In P91 side, The equiaxed lath with PAGBs and precipitates are observed,
archives of civil and mechanical engineering 18 (2018) 713–722 719

Fig. 8 – Micrographs of weld zone and FGHAZ for A-TIG welds joint (a) FGHAZ P91 side, (b) weld zone, (c) FGHAZ P92 side;
micrographs of weld zone and FGHAZ for GTA welds joint (d) FGHAZ P91 side, (e) weld zone, (f) FGHAZ P92 side.

Fig. 9 – Schematic change in microstructure of weld zone and FGHAZ during welding, after the welding, and after the PWHT.

as shown in Fig. 8(d) and (f). The change in microstructure of The grain size of FGHAZ does not show much significant
FGHAZ and weld fusion zone during the weld thermal cycle difference after the PWHT of 6 h. It was measured to be 27  9 mm
and after the PWHT are shown in schematic Fig. 9 [25]. to 25  8 mm for FGHAZ P91 and FGHAZ P92, respectively. Hence,
After the PWHT of 6 h, micrographs of A-TIG and GTAW it can be stated that for A-TIG weld joint uniformity in
joints are presented in Fig. 10. It was observed that d-ferrite did microstructure was obtained after the 6 h of PWHT. However,
not respond to PWHT and it remained in the microstructure, as in GTA weld joint it was observed after the 2 h of PWHT.
shown in Fig. 10(b) and (c). The grain size of fusion zone also
remain constant after the PWHT of 6 h (62  31 mm). However, 3.4. d-Ferrite evolution and microhardness
heterogeneity in grain size was observed to be increased.
PWHT at 6 h resulted in uniform grain formation in FGHAZ for The dissimilar weld joint of P91 and P92 steel has been
both P91 (42  18 mm) and P92 side (41  19 mm). In GTA weld prepared in two different conditions. The composition of weld
joint, columnar lath morphology retained in the microstruc- fusion zone are predicted in Table 1. The ferrite forming
ture. However, near the fusion boundary along the P92 side, tendency is calculated form the Schneider formula which is
few patches of d-ferrite were observed, as shown in Fig. 10(g). most commonly used for the 9–12% Cr ferritic steels and other
720 archives of civil and mechanical engineering 18 (2018) 713–722

Fig. 10 – Optical micrographs for A-TIG welds after the PWHT of 6 h (a) FGHAZ P91 side, (b) weld zone at 200T, (c) weld zone at
500T (d) FGHAZ P92 side; GTA welds after the PWHT of 6 h (e) FGHAZ P91 side, (f) weld zone, (g) Fusion boundary of P92 side,
(h) FGHAZ P92 side.

high Cr steels [26]. It is calculated by taking the difference of


chromium equivalent (Creq) and nickel equivalent (Nieq). It is
given by Eq. (1) [27]:

Ferrite factor ðFFÞ ¼ Cr þ 2Si þ 1:5Mo þ 5V þ 1:75Nb


þ 0:75WNi0:5Mn30C25N0:3Cu
(2)

FF was calculated to be 6.18 and 7.08 for A-TIG and GTA


welds. Oñoro [27] had reported that for such low FF (<8.5), it is
difficult to form the ferrite parches. Although, in some studies
having FF lower than proposed limit, showed the formation of
d-ferrite [13].
In present study, d-ferrite patches are still observed for such
low FF in A-TIG welds. In A-TIG welds, volume fraction of
d-ferrite contents was measured 0.60% while in FGHAZ of P92 Fig. 11 – Variation in hardness of A-TIG and GTA weld in as-
side, 0.25% of volume fraction of ferrite content was observed. welded and PWHT condition.
In GTA welds having higher FF than A-TIG welds, no ferrite
patches was observed in weld fusion zone and FGHAZ. However,
0.10% of volume fraction of ferrite content was observed near the
fusion boundary of P92 side. In A-TIG welds, higher heat input, hardness was noticed for PWHT of 6 h. The effect of soft ferrite
wide solidification range, and higher weight percentage of ferrite zone formation reduces the overall hardness of the weld fusion
stabilizer (Si: 0.29; Nb: 0.06; V: 0.23) lead to formation of d-ferrite. zone. In A-TIG welds, the peak hardness of weld fusion zone
However, in GTA welds slow cooling rate, narrow solidification was measured to be 490 HV while in GTA welds it was 469 HV.
range and lower weight percentage of ferrite stabilizer (Si: 0.28; The peak hardness in A-TIG weld zone is attributed to higher
Nb: 0.04; V: 0.20) might be the cause of absence of d-ferrite weight percentage of carbon (C) in weld zone due to high heat
patches. The formation of ferrite patches in FGHAZ along the P92 input in single pass. However, the average hardness of weld
side might be the cause of higher weight percentage of ferrite fusion zone was measured to be 436  23 HV, and 442  18 HV
stabilizer like W in P92 steel as compared to P91 steel. for A-TIG and for GTA welds. The reduction in average hardness
Variation in hardness of A-TIG and GTA weld joints for is attributed to formation of soft d-ferrite patches. The variation
different heat treatment condition is shown in Fig. 11. The in hardness of d-ferrite patches is shown in Fig. 12.
heterogeneity in hardness of A-TIG weld zone is occurred After the PWHT of 2 h, average hardness of weld fusion
mainly due to presence of soft-ferrite patches while in GTA zone were measured to be 269  20, and 260  15 HV for A-TIG
welds it occurs mainly due to auto-tempering of martensite. In and GTA welds, respectively. The reduction in hardness is
as-welded condition, higher hardness is attributed to untem- attributed to tempering of martensite and reduction in solid
pered lath martensite. A continuous reduction in heterogeneity solution hardening due to evolution of precipitates. After the
was noticed after the PWHT. Effect of PWHT at 7608C for 2 h is PWHT of 6 h, average hardness of weld fusion zone were
also presented by Pandey et al. [5]. The minimum variation in measured to be 243  15, and 243  10 HV for A-TIG and GTA
archives of civil and mechanical engineering 18 (2018) 713–722 721

Fig. 12 – Variation in hardness of soft d-ferrite zone.

Table 5 – Charpy toughness results.


Weld zone As-welded PWHT-2 h PWHT-6 h

Hardness (HV) Charpy toughness (J) Hardness (HV) Charpy toughness (J) Hardness (HV) Charpy toughness (J)
A-TIG 490 20 269 48 243 58
GTAW 469 42 260 60 243 76

welds, respectively. A considerable lowering in hardness was toughness. In GTA welds, fine grains lead to higher
noticed after the PWHT of 6 h. precipitation of M23C6 type carbide precipitates along the
boundaries that results in higher Charpy toughness of GTA
3.5. Charpy toughness welds fusion zone. Hence, higher content of V and Nb and
poor response to tempering leads to higher tempering
Variation in Charpy toughness value of weld fusion zone along duration of 6 h for A-TIG welds than GTA welds. Schafer [9]
with the hardness are depicted in Table 5. From Table 5, it is had also reported that combined effect of d-ferrite and brittle
clear that, A-TIG welds exhibited the poor Charpy toughness 'dendritic' carbides of M23C6 types (M  65%Cr + 30%Fe + rest)
as compared to GTA welds. In as-welded condition, poor reduced the toughness by encapsulate the d-ferrite.
Charpy toughness is attributed to untempered lath martens-
ite. In GTA welds, higher Charpy toughness is observed mainly 4. Conclusions
due to auto-tempering of brittle martensite. To avoid the
brittle fracture of in P91 and P92 steel weldments, a minimum
average toughness for weld zone was recommended about 47 J  P91 and P92 steel plate of thickness 5 mm were welded
(minimum single value of 38 J) as per EN 1599:1997. In present successfully using autogenous tungsten inert gas (A-TIG)
case, A-TIG and GTA weld exhibit the Charpy toughness value welding process and plates of 8 mm thickness were welded
lower than the minimum required value of 47 J. To meet the using the multi-pass gas tungsten arc welding (GTAW)
minimum required Charpy toughness value PWHT is con- process.
ducted. After the PWHT of 2 h, both A-TIG and GTA welds  A-TIG and GTAW process were found a noticeable effect on
attained the minimum required Charpy toughness value but microstructure evolution and mechanical properties of P91
still A-TIG welds shows the poor Charpy toughness. The PWHT and P92 dissimilar weld joints.
effect on Charpy toughness is also presented by Pandey et al.  In A-TIG weld joint, coarse martensitic microstructure was
[5]. After the PWHT of 6 h, both the weld joints have attained formed due to high peak temperature as compared to GTA
the minimum required Charpy toughness value successfully. welds joint. The coarse grain in fine grained heat affected
Higher V and Nb content in A-TIG weld (Table 1) might be zone (FGHAZ) and coarse martensitic microstructure in weld
the possible cause of poor Charpy toughness value than the zone are attributed to high peak temperature of 1.07 kJ/mm
GTA welds [28]. The packet size of lath martensite also affect in A-TIG process.
the toughness value of weld fusion zone. Fine packet size  In as-welded condition, volume fraction of d-ferrite in weld
results in higher toughness while coarse packet size leads to zone and FGHAZ of P92 side were measured to be 0.60 and
poor toughness value because stress concentration to occur 0.25%, respectively. The d-ferrite patches have observed no
ahead of a crack tip. After the PWHT of 2 h and 6 h, weld any response to PWHT.
fusion zone of A-TIG welds showed the coarse packet size of  The peak hardness measured in A-TIG weld zone was much
tempered martensite laths. That results in poor Charpy higher than GTA weld zone due to higher C content in A-TIG
722 archives of civil and mechanical engineering 18 (2018) 713–722

weld zone. However, average hardness was GTA weld was [11] K.S. Chandravathi, K. Laha, K.B.S. Rao, S.L. Mannan,
higher than A-TIG weld zone due to formation of soft d-ferrite Microstructure and tensile properties of modified 9Cr–1Mo
steel (grade 91), Mater. Sci. Technol. 17 (2001) 559–565.
patches. The hardness was observed to be decreased con-
[12] P.J. Grobenr, W.C. Hagel, The effect of molybdenum on high-
tinuously with increase in PWHT duration due to reduction in
temperature properties of 9 Pct Cr steels, Metall. Trans. A 11A
solid solution hardening and tempering of martensite. (1980) 633–642.
 The uniform grain structure in FGHAZ was observed for both [13] K. Anderko, L. Sch, A.K. Ewaiom, Effect of the d-ferrite phase
P91 and P92 side after 2 h of PWHT in GTA weld and 6 h of on the impact properties chromium steels, J. Nucl. Mater.
PWHT for A-TIG welds. 179–181 (1991) 492–495.
 In as-welded condition, A-TIG weld zone showed the mini- [14] X.Y. Liu, T. Fujita, Chromium chromium, ISIJ Int. 29 (1989)
680–686.
mum Charpy toughness because of higher weight percentage
[15] E.D. Specht, S.M. Allen, Communication Cr Steel Investigated
of V and Nb and presence of d-ferrite patches. However, GTA by In-Situ X-Ray Radiation, 2010, 2–5. , http://dx.doi.org/
weld showed higher Charpy toughness value as compared to 10.1007/s11661-010-0371-7.
A-TIG due to auto-tempering by multi-pass process. [16] X. Tao, J. Gu, L. Han, Characterization of precipitates in
 To meet the minimum specific requirement Charpy tough- X12CrMoWVNbN10-1-1 steel during heat treatment, J. Nucl.
ness value (47 J) as per the EN: 1557:1997 specification, GTA Mater. 452 (2014) 557–564. , http://dx.doi.org/10.1016/
j.jnucmat.2014.06.018.
welds required the PWHT of 2 h at 760 8C while A-TIG
[17] C. Pandey, A. Giri, M.M. Mahapatra, Evolution of phases in P91
required 6 h, at same temperature.
steel in various heat treatment conditions and their effect on
microstructure stability and mechanical properties, Mater.
Sci. Eng. A 664 (2016) 58–74. , http://dx.doi.org/10.1016/
j.msea.2016.03.132.
references
[18] C. Pandey, M. Mahapatra, Evolution of phases during
tempering of P91 steel at 760 for varying tempering time and
their effect on microstructure and mechanical properties, Proc.
[1] C. Pandey, M.M. Mahapatra, P. Kumar, N. Saini, Effect of Inst. Mech. Eng. Part E J. Process Mech. Eng. 664 (2016) 58–74. ,
strain rate and notch geometry on tensile properties and http://dx.doi.org/10.1177/0954408916656678.
fracture mechanism of creep strength enhanced ferritic P91 [19] N. Saini, C. Pandey, M.M. Mahapatra, Characterization and
steel, J. Nucl. Mater. 498C (2017) 176–186. , http://dx.doi.org/ evaluation of mechanical properties of CSEF P92 steel for
10.1016/j.jnucmat.2017.10.037. varying normalizing temperature, Mater. Sci. Eng. A (2017),
[2] B. Silwal, L. Li, A. Deceuster, B. Griffiths, Effect of postweld http://dx.doi.org/10.1016/j.msea.2017.02.022.
heat treatment on the toughness of heat-affected zone for [20] N. Saini, C. Pandey, M. Mohan, Microstructure evolution and
grade 91 steel, Weld. Res. 92 (2013) 80s–87s. mechanical properties of dissimilar welded joint of P911 and P92
[3] C. Pandey, M.M. Mahapatra, P. Kumar, N. Saini, Effect of steel for subsequent PWHT and N&T treatment, Trans. Indian
normalization and tempering on microstructure and Inst. Met. (2017), http://dx.doi.org/10.1007/s12666-017-1145-3.
mechanical properties of V-groove and narrow-groove P91 [21] C.G. Panait, A. Zielinska-Lipiec, T. Koziel, A. Czyrska-
pipe weldments, Mater. Sci. Eng. A 685 (2017) 39–49. , http:// filemonowicz, A.F. Gourgues-Lorenzon, W. Bendick, Evolution
dx.doi.org/10.1016/j.msea.2016.12.079. of dislocation density, size of subgrains and MX-type
[4] Vijaya L. Manugula, Koteswararao V. Rajulapati, G. precipitates in a P91 steel during creep and during thermal
Madhusudhan Reddy, K. Bhanu, Sankara Rao, Role of evolving ageing at 600 8C for more than 100,000 h, Mater. Sci. Eng. A 527
microstructure on the mechanical properties of electron beam (2010) 4062–4069. , http://dx.doi.org/10.1016/j.msea.2010.03.010.
welded ferritic–martensitic steel in the as-welded and post [22] C. Pandey, M.M. Mahapatra, P. Kumar, R.S. Vidyrathy, A.
weld heat-treated states, Mater. Sci. Eng. A 698 (2017) 36–45. , Srivastava, Microstructure-based assessment of creep
http://dx.doi.org/10.1016/j.msea.2017.05.036. rupture behaviour of cast-forged, Mater. Sci. Eng. A 695
[5] C. Pandey, M. Mohan, P. Kumar, N. Saini, Dissimilar joining of (2017) 291–301. , http://dx.doi.org/10.1016/j.msea.2017.04.037.
CSEF steels using autogenous tungsten-inert gas welding and [23] C. Pandey, M.M. Mahapatra, P. Kumar, N. Saini, Characterization
gas tungsten arc welding and their effect on d-ferrite evolution of cast and forged (C&F) Gr. 91 steel in different heat treatment
and mechanical properties, J. Manuf. Process. 31 (2018) 247–259. condition, Trans. Indian Inst. Met. (2017), http://dx.doi.org/
, http://dx.doi.org/10.1016/j.jmapro.2017.11.020. 10.1007/s12666-017-1144-4.
[6] M.L. Santella, R.W. Swindeman, R.W. Reed, J.M. Tanzosh, [24] C. Pandey, M.M. Mahapatra, P. Kumar, A. Giri, Microstructure
Martensite formation in 9Cr–1Mo steel weld metal and its characterization and Charpy toughness of P91 weldment for
effect on creep behavior, n.d. as-welded, PWHT and N&T heat treatment, Met. Mater. Int.
[7] B. Arivazhagan, R. Prabhu, S.K. Albert, M. Kamaraj, S. (2017), http://dx.doi.org/10.1007/s12540-017-6850-2.
Sundaresan, Microstructure and mechanical properties of [25] C. Pandey, M.M. Mahapatra, P. Kumar, N. Saini, Homogenization
9Cr–1Mo steel weld fusion zones as a function of weld metal of P91 weldments using varying normalizing and tempering
composition, J. Mater. Eng. Perform. 18 (2009) 999–1004. , treatment, Mater. Sci. Eng. A 710 (2018) 86–101. , http://dx.doi.org/
http://dx.doi.org/10.1007/s11665-008-9349-7. 10.1016/j.msea.2017.10.086.
[8] C. Pandey, M.M. Mahapatra, Effect of heat treatment on [26] S.W. Shyu, H.Y. Huang, K.H. Tseng, C.P. Chou, Study of the
microstructure and hot impact toughness of various zones of performance of stainless steel A-TIG welds, J. Mater. Eng.
P91 welded pipes, J. Mater. Eng. Perform. 25 (2016) 2195–2210. , Perform. 17 (2008) 193–201. , http://dx.doi.org/10.1007/s11665-
http://dx.doi.org/10.1007/s11665-016-2064-x. 007-9139-7.
[9] L. Sch afer, Influence of delta ferrite and dendritic carbides on [27] J. Onoro, Martensite microstructure of 9–12% Cr steels weld
the impact and tensile properties of a martensitic chromium metals, J. Mater. Process. Technol. 180 (2006) 137–142. , http://
steel, J. Nucl. Mater. 263 (1998) 1336–1339. dx.doi.org/10.1016/j.jmatprotec.2006.05.014.
[10] R.G. Faulkner, J.A. Williams, E.G. Sanchez, A.W. Marshall, [28] T. Onizawa, T. Wakai, M. Ando, K. Aoto, Effect of V and Nb on
Influence of Co, Cu and W on Microstructure of 9% Cr Steel precipitation behavior and mechanical properties of high Cr
Weld Metals, vol. 19, 2003, http://dx.doi.org/10.1179/ steel, Nucl. Eng. Des. 238 (2008) 408–416. , http://dx.doi.org/
026708303225009652. 10.1016/j.nucengdes.2006.09.013.

Вам также может понравиться