Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

Accepted Manuscript

Title: Effect of long-term aging on the microstructure and


mechanical properties of T23 steel weld metal without
post-weld heat treatment

Authors: Xue Wang, Yong Li, Huijun Li, Sanbao Lin, Yaoyao
Ren

PII: S0924-0136(17)30414-4
DOI: http://dx.doi.org/10.1016/j.jmatprotec.2017.09.015
Reference: PROTEC 15392

To appear in: Journal of Materials Processing Technology

Received date: 11-3-2017


Revised date: 29-8-2017
Accepted date: 9-9-2017

Please cite this article as: Wang, Xue, Li, Yong, Li, Huijun, Lin, Sanbao, Ren, Yaoyao,
Effect of long-term aging on the microstructure and mechanical properties of T23
steel weld metal without post-weld heat treatment.Journal of Materials Processing
Technology http://dx.doi.org/10.1016/j.jmatprotec.2017.09.015

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Effect of long-term aging on the microstructure and mechanical

properties of T23 steel weld metal without post-weld heat treatment

Xue WANG a, b*,Yong LI a,Huijun LI c, Sanbao LIN b,Yaoyao REN a

a) School of power and mechanics, Wuhan University, Wuhan 430072, China


b) State Key Laboratory of Advanced Welding and Joining, Harbin Institute of Technology, Harbin 150001, China
c) Faculty of Engineering and Information Sciences, University of Wollongong, NSW 2522, Australia;

*Corresponding author. Tel.: +86 13554693820; fax: +86 02768772253.


E-mail address: wangxue2011@whu.edu.cn (Xue Wang)

Graphical abstract

ABSTRACT: T23 steel welds without post-weld heat treatment (PWHT) were aged for 3,000 h at temperatures in the range

773–923 K. Microstructural evolution in the aged weld metal (WM) was observed and its influence on the hardness and Charpy

impact toughness were investigated. The as-welded WM exhibits a martensitic/bainitic microstructure with moderate hardness

less than 350 HB and impact toughness value above 110 J/cm2. Aged WM exhibits deterioration of mechanical properties

dependent on the temperature of aging. When the aging temperature is below 823 K, the WM shows little change in hardness but

its impact toughness drops significantly. When the aging temperature is above 823 K, the hardness of the WM starts to decrease

quickly while its impact toughness increases gradually. As the aging temperature increases to 923 K, its impact toughness further

increases and recovers to a value comparable to that of the as-welded condition. Embrittlement occurring during aging at 773–823

K results from tempered martensite embrittlement (TME), which is caused by the decomposition of retained austenite. When the

1
aging temperature exceeds 823 K, M23C6 carbides precipitate rapidly inside grains and at grain boundaries and MX carbides begin

to precipitate inside grains. These cause the release of carbon from the matrix, reduction of lattice distortion, and accelerated

recovery of laths causing the hardness to decrease rapidly and the impact toughness to increase. W-rich carbides are observed in a

minority of grain boundaries in the WM aged at temperatures higher than 823 K, which may cause intergranular fracture and

retard the improvement of impact toughness.

KEYWORDS: T23 steel; weld metal; microstructure; mechanical properties

1. Introduction

Masuyama et al. (1994) developed the steel grade T23 (low C-2.25Cr-1.6W-V-Nb), which has been widely

used in boiler tubes for advanced fossil-fuel-fired power plants. According to Vaillant et al. (2008), its main

fields of application are water wall panels for ultra-supercritical boilers (USCB), and superheaters and reheaters

for conventional boilers and heat recovery steam generators (HRSG). Steel T23 was developed on the basis of

steel T/P22 (10CrMo9-10) to enhance creep rupture strength by substituting W for a portion of Mo and adding

the elements V, Nb, and N. After proper normalizing and tempering (NT) or quenching and tempering (QT)

treatment, T23 exhibits a stable fully bainitic microstructure with creep rupture strength about 1.8 times higher

than that of T22 steel. Masuyama et al. (1994) reported that T23 welds produced by gas tungsten arc welding

(GTAW) and shielded metal arc welding (SMAW) had hardness lower than 350 HV 10 in both the weld metal

(WM) and heat-affected zones (HAZs) without post-weld heat treatment (PWHT). Bendick et al. (2007) also

confirmed hardness in the WM lower than 350 HV 10, and proposed that T23 steel allows the welding of thin

wall tubes (δ < 10 mm) without PWHT. The welding procedure is simplified and the cost of boiler manufacture

is reduced if the PWHT can be avoided.

However, it is not sufficient for the as-welded condition to just fulfill the requirements of welds in terms of

hardness and toughness. Recently, Ji et al. (2009) reported that many water wall panels manufactured using T23

2
tubes without PWHT burst during service in a USCB operating in China and cracks were found in the WM of

the welds. These potential premature failures of welds threaten the operational safety of power plants. Wang et

al. (2015) investigated the failures of T23 WM in water walls and found that creep embrittlement occurred in

the WM during service, causing cracking. Mohyla and Foldyna (2009) found that un-tempered weld joints of

both T23 and T24 steels undergo a process of secondary hardening during long-term exposure at 823 K and

concluded that secondary hardening causes embrittlement. Secondary hardening cannot be used to explain the

failure that took place beyond 1,000 h of service because it occurred within 1,000 h of aging. Therefore, the

underlying mechanism for embrittlement in T23 welds has not been revealed entirely.

Zieliński et al. (2016) reported the microstructure evolution and mechanical properties of T23 steel after

long-term aging up to 70,000 h at temperatures of 823 K and 873 K; the results indicated that long-term aging of T23

steel with metastable microstructure resulted in deterioration of mechanical properties and the magnitude of these

changes depended on aging temperature. Sawada et al. (2009) investigated the creep behavior of T/P23 steel and

found that the high initial hardness accelerates microstructural degradation, which causes accelerated drop of

creep strength. This may be consistent with the results provided by Okada et al. (2010) that the creep strength of

samples with lower original hardness provide creep properties similar to those with higher hardness when long-term

creep rupture properties are considered. These results indicate that different initial microstructures have different

performances in terms of microstructure evolution and effect on creep strength. T23 steel was used in an NT

condition. Through tempering at 1,003–1,053 K after normalizing at 1,333 ± 10 K, a relatively stable

microstructure can be obtained with the precipitation of carbides, but degradation will still occur after

long-term service. Considering the WM in the as-welded condition similar to a quenched condition, most of the

carbon is dissolved in the matrix and will precipitate in the form of carbides following aging or service

exposure. As-welded WM has a more unstable microstructure than as-received T23 steel and may cause

3
significant changes in mechanical properties during long-term exposure. Zieliński et al. (2017) reported that

steel grade T24—a new low alloy steel that has alloying elements similar to T23—exhibits very good

performance in the delivery state and after long-term temperature influence. However, it has been reported that

numerous leaks occurred in welds without PWHT after short-term service (Nowack et al., 2011) and

embrittlement also occurred in its WM, as in T23 (Mohyla and Foldyna, 2009). Therefore, it is necessary to

study the evolution of microstructure and corresponding changes in mechanical properties for T23 WM during

long-term aging, to clarify the mechanisms for the failure of T23 welds and offer references for the use of T24

steel welds.

Zieliński et al. (2016) presented a procedure to determine the creep strength based on results obtained from

abridged creep tests, which can result in a significant reduction in the waiting time. Furthermore, the practical

application of abridged creep tests has advantages when estimating the residual life and the useful residual life, but

an impediment to applying the short-term creep tests is the limitation of the range of test temperatures to be used,

which results from the precipitation process in the material (Zieliński et al., 2016). Thus, the precipitating behaviors

of carbides at different aging temperatures should be investigated to clarify the effects of microstructure on the

mechanical properties of the WM and to provide reference for the temperature choice in short-term creep tests.

In this paper, based on the service temperatures of different components fabricated with T23 steel in

USCB, the aging tests of as-welded WM were conducted in the temperature range 773–923 K for up to 3,000 h.

Subsequently, microstructure evolution and the changes of mechanical properties were investigated to analyze

the reasons for the premature failure of T23 welds during service.

2. Experimental

T23 tubes with outer diameter of 38.1 mm and wall thickness of 6.5 mm were used as the base metal (BM)

for the preparation of butt welds. The filler metal used was a Thyssen WZCr2WV wire 2.4 mm in diameter. The

4
chemical composition (in mass percent) of the WM was C0.037, Si0.33, Mn0.68, Cr2.34, Mo0.17, W1.47,

V0.20, B0.0018, Nb0.038, N0.015, and Fe (balance). The composition of the filler metal was close to the

nominal composition of T23 steel, except that the carbon content was lower than the minimum value (0.04%)

of the ASME SA-213 standard. The joints were welded in the fixed vertical position (2G) with a V-groove. The

circumferential weld was performed in three layers using a manual GTAW process with heat input of 12–15

kJ/cm without preheating and with inter-pass temperature ≤ 523 K. The welded joints were cooled slowly in

ceramic wool to avoid hydrogen-induced cold cracking. Magnetic particle inspection and X-ray inspection of

the welds did not reveal welding defects. Subsequently, the aging test was conducted. The aging temperatures

selected were 773 K, 823 K, and 873 K, corresponding to the equivalent service exposure temperatures of tubes

for water walls, reheaters, and superheaters, respectively. Furthermore, the temperature of 923 K was chosen to

study the microstructure change due to overheating. The aging tests were conducted for 3,000 h at each

temperature.

After aging, impact tests at room temperature were carried out on Charpy V specimens of size 5 mm × 10

mm × 55 mm, and notches were machined in the centers of the welds. The hardness of the WM was examined

using a 320HBS-300/0035 Brinell hardness tester with load of 187.5 kg.

Specimens for microstructure examination were cut from the center of the WM. Optical microscopy (OM)

was carried out using an OLYMPUS-PMG3. The observation of fine microstructure and precipitates, and

analysis of morphologies of impact fracture surfaces were conducted using a QUANTA400 scanning electron

microscope (SEM). Thin foil samples and carbon extraction replicas for transmission electron microscopy

(TEM) were prepared to investigate the substructure and fine precipitates in the WM. Thin foils were examined

on a JEOL-2010 operating at 200 kV. The precipitates were identified by TEM in the energy-dispersive X-ray

spectroscopy (EDX) and selected area electron diffraction (SAED) modes. Moreover, changes in the phase

5
components in the T23 steel WM with aging temperature were calculated using Thermo-Calc.

3. Results and discussion

3.1 Microstructures in as-welded condition

Microstructures of BM and as-welded WM in T23 welds are shown in Fig. 1. BM exhibited a fully bainitic

microstructure decorated with carbides precipitated both at the prior austenite grain boundaries and inside

grains (Fig. 1a and b). WM exhibited mixed microstructures of un-tempered martensite/bainite with a coarse

prior austenite grain size (Fig. 1c). Lath martensite was observed in the SEM micrograph (Fig. 1d) and many

short rod-shaped martensite-austenite (M-A) constituents lined up along lath boundaries with bright contrast

(highlighted with arrows). In addition, a few short rod-shaped or granular M-A constituents could also be

observed at the prior austenite grain boundary.

Fig. 1. Microstructural morphologies of (a), (b) BM and (c), (d) as-welded WM.

Figure 2 shows thin foil TEM images of the WM. It shows that the as-welded WM had a high proportion

6
of un-tempered martensite with lath substructure and high density of dislocations. Small amounts of fine M3C

carbides were found in lath interiors (Fig. 2a), probably as a result of self-tempering due to relatively slow

cooling or short-time tempering produced by the following welding layer. Furthermore, precipitates were not

formed at grain boundaries and the lamellar retained austenite was distributed along lath boundaries or prior

austenite grain boundaries, as shown in Fig. 2b, which is consistent with the SEM results (Fig. 1d).

Fig. 2. TEM images from a thin foil of as-welded WM: (a) laths, (b) three-grain boundaries.

The above investigation shows that the microstructure of the as-welded WM is characterized by a large

portion of un-tempered martensite with a small amount of carbides precipitated at lath interiors. This is

different from BM in the as-received state.

3.2 Microstructures after aging

Figure 3 shows OM images of the WM aged at different temperatures for 3,000 h. WM still had visible

martensitic laths when aged at temperatures below 823 K, while the lath microstructure appeared to reduce

markedly at aging temperatures higher than 873 K. When the aging temperature increased to 923 K, the lath

microstructures further reduced and precipitates could be found at the prior austenite grain boundaries and

inside grains.

7
Fig. 3. Optical micrographs of WM aged at different temperatures for 3,000 h:
(a)773 K, (b) 823 K, (c) 873 K, and (d) 923 K.

Figure 4 shows SEM micrographs of the WM aged at different temperatures for 3,000 h. After aging at

773 K for 3,000 h, most of the M-A constituents decomposed but few precipitates could be observed at the

prior grain boundaries or in the grain interiors (Fig. 4a), indicating that most alloy elements still remained in the

solid solution matrix. After aging at 823 K for 3,000 h, the precipitate increased but was still small (Fig. 4b).

When the aging temperature increased to 873 K, both the quantity and size of precipitates increased markedly

at the grain boundaries and within the grains (Fig. 4c). At 923 K, the carbides at the grain boundaries increased

and seemed to grow (Fig. 4d).

8
Fig. 4. SEM images of WM aged at different temperatures for 3,000 h:
(a)773 K, (b) 723 K, (c) 873 K, and (d) 923 K.

Figure 5 shows TEM images of thin foils of WM aged at 773 K and 923 K for 3,000 h. Unlike the

as-welded condition (Fig. 2a), the retained austenite was resolved and a few tiny carbides were found in the

grain interiors in WM aged at 773 K (Fig. 5a and b). However, there was no obvious increase in lath width and

high-density dislocations remained, indicating the lack of recovery. When aged at 923 K, obvious recovery and

polygonization were observed, sub-grains started forming and large carbides of size 200–500 nm precipitated at

grain boundaries (Fig. 5c). Additionally, carbides precipitated within the grains. Larger carbides exhibited a

granular shape while some small carbides exhibited an acicular shape (Fig. 5d). The above results suggest that

WM did not undergo change in microstructure when aged at temperatures below 773 K, which is attributed to

the extremely limited diffusion of carbon at this temperature.

9
Fig. 5. TEM images from a thin foil in WM aged at 773 K and 923 K for 3,000 h:
(a), (b) 773 K and (c), (d) 923 K.

Figure 6 shows typical images of extraction replicas of WM aged at 823 K for 3,000 h, and Table 1 lists

the corresponding EDX results. Blocky carbides of size about 300 nm were seen present at the grain boundary.

These carbides contained mainly Cr and Fe (exceeding 80% in atom percentage), about 15% Mn, V, W, and

very little Nb and Mo. Deng et al. (2007) found that coarse M23C6 carbides precipitated at the grain boundary

following tempering after the normalizing of T23 steel. Based on their results, the intergranular carbides shown

in Fig. 6a may be M23C6 carbides. There were some fine carbides of about 100 nm present inside grains, as

shown in Fig. 6b, and their compositions were similar to those of intergranular carbides, which means that they

were also M23C6. No V- or Nb-rich MX phase was found in the WM aged at this condition.

10
Fig. 6. TEM images of precipitate from a replica of the WM aged at 823 K for 3,000 h:
(a) precipitates on grain boundaries and (b) precipitates inside grain.

Table 1 EDS compositions of precipitates in the WM aged at 823 K for 3,000 h


(atomic fraction /%).

Element V Cr Mn Fe Mo W

Precipitates in Fig. 6a 1.94±1.36 49.92±13.34 6.69±0.89 35.75±10.93 1.01±0.62 4.69±2.33

Precipitates in Fig. 6b 3.77±2.39 56.25±8.75 6.59±1.07 29.56±8.59 1.15±0.45 2.68±1.46

Figure 7 shows typical images of extraction replicas of WM aged at 923 K for 3,000 h and Table 2

provides the corresponding EDX results. As the WM aged at 823 K for 3,000 h, M23C6 carbides rich in Cr and

Fe were also found but were coarser. Furthermore, another precipitate observed was rich in V and Nb and was

tiny and dispersed, which is identified as the MX-phase.

Fig. 7. TEM images of precipitate inside grain from a replica of the WM aged at 923 K for 3,000 h:

11
(a) low magnification and (b) high magnification.

Table 2 EDS compositions of precipitates inside grains in the WM aged at 923 K for 3,000 h
(atomic fraction /%).

Element V Cr Mn Fe Nb Mo W

Large precipitate in
6.74±1.43 60.81±1.34 3.78±0.45 26.03±0.79 - 0.65±0.26 1.61±0.23
Fig. 7a

Fine precipitate in
47.47±10.53 12.52±2.90 - 19.77±9.13 7.55±3.41 2.83±1.65 9.87±2.13
Fig. 7b

Figure 8a shows the morphology of intergranular carbides of WM aged at 923 K. They grew to sizes

exceeding 500 nm and seemed to agglomerate. These carbides were also rich in Cr and Fe, such as those

precipitated at 823 K, therefore they should be M23C6 carbides. The SAED result further confirmed that they

were M23C6 carbides (Fig. 8b).

Fig. 8. TEM images of extraction replica of precipitate at the grain boundaries from the WM aged at 923 K for 3,000 h:
(a) low magnification and (b) high magnification.

3.3 Thermodynamic calculations of phases

Figure 9 shows the phase diagrams of the WM calculated at equilibrium and at metastable equilibrium.

From Fig. 9a it can be seen that the amount of M23C6 carbide is very low (approximately zero) when the

temperature is above 773 K, while M6C carbide has a higher presence in the range 773–823 K even though its

amount decreases with increasing temperature. Figure 9a predicts the constituents of phases of WM as a

function of temperature at equilibrium but cannot predict those of phases before reaching equilibrium. Komai et

12
al. (2005) found that there was only a small quantity of M6C carbides precipitated in T23 superheater tubes

after service at 847 K for 79,102 h. Wang et al. (2006) also reported that only a small portion of M23C6 carbides

evolved to M6C carbides when T23 steel was crept at 923 K for 5,109 h. Miyata and Sawaragi. (2001)

compared the precipitate evolution behavior between 2.25Cr-Mo-V-Nb and 2.25Cr-W-V-Nb steels during creep

at 923 K and found that W retards the formation of M6C carbide at the grain boundary and thus improves the

creep rupture life. From their investigations, it can be concluded that M6C carbide precipitates during the late

stage of aging of T23 steel. This was also true for the WM, as very little M6C carbides were observed even after

aging at 923 K for 3,000 h, which shows that tungsten addition to the WM contributed to stabilizing the M23C6

carbides and delayed the formation of M6C carbides. Figure 9b presents the constituents of phases where the

formation of M6C carbide was not observed; compared to the equilibrium, the amount of M23C6 carbides

increased remarkably while the other carbides remained unchanged at temperatures higher than 773 K,

indicating that M23C6 carbides transform to M6C carbides at equilibrium.

Fig. 9. Calculated phase diagram of the WM of T23 steel:


(a) with the precipitation of M6C carbide and (b) without the precipitation of M6C carbide.

The amount of Mo2C carbide was very low because of the low content of Mo in T23 WM. Moreover,

Mo2C carbide was resolved when the aging temperature exceeded 783 K; hence, Mo2C carbide was not

observed in the T23 WM. Figure 9b shows that MX carbides increased when temperature exceeded 823 K,

which is consistent with the results of the TEM microstructural examination. When the temperature was lower

13
than 823 K, the lack of MX carbides in the WM may be summarized by thermodynamic instability and the low

diffusion rates of V and Nb. A longer aging time is required to precipitate MX carbide at temperatures below

823 K, and thus carbon preferentially combines with Cr to form the M23C6 carbide. From Fig. 9b it is seen that

M23C6 carbide reduced markedly while MX carbide increased as the temperature exceeded 823 K as a result of

the dissolution of M23C6 carbides and the increase in diffusion rates of V and Nb.

3.4 Mechanical properties

Figure 10 shows the effects of aging temperature on the hardness and Charpy impact toughness. From Fig.

10a it can be seen that the as-welded WM exhibited high hardness of 325 HB as a result of its un-tempered

martensite microstructure with supersaturated carbon content and high dislocation density. After aging at 773 K

and 823 K, the hardness decreased slightly owing to the retention of most martensite lath and high-density

dislocations. When the temperature exceeded 873 K, the hardness decreased markedly; this can be attributed to

the recovery of laths and reduction of the degree of supersaturation of carbon due to the precipitation of

carbides. Although the as-welded WM exhibited sufficient impact toughness (exceeding 110 J/cm2), its mean

value dropped rapidly to only 10 J/cm2 and 15 J/cm2 when aged at 773 K and 823 K, respectively, for 3,000 h,

indicating severe embrittlement. When the aging temperature increased to 873 K, the impact toughness

improved relative to the aging conditions of 773 K or 823 K, but was still lower than the value in the as-welded

condition. When the aging temperature increased to 923 K, the impact toughness improved and was close to the

value of the as-welded condition. Dobrzański et al. (2009) studied the effects of tempering time and

temperature on the hardness and impact toughness of T23 steel after NT treatment and showed that the hardness

decreased slightly after aging for 3,000 h both at 823 K and 873 K, whereas the impact toughness almost did

not change. Zieliński et al. (2016) also reported that the hardness and impact toughness reduced slightly during

aging of the as-received T23 steel for 3,000 h at 823 K and 873 K. Therefore, it is clear that there were large

14
differences in the responses of mechanical properties during long-term aging between T23 steel and WM

without PWHT.

Fig. 10. Effect of aging temperature on the hardness and toughness of WM after 3,000 h aging:

(a) hardness and (b) impact toughness.

The fracture impact surfaces of the as-welded WM and after aging at 773 K and 823 K are shown in Fig.

11. Apparent macro-deformation was observed in the as-welded WM (Fig. 11a). The radical zone exhibited

intra-crystalline fracture with quasi-cleavage features. Many fine tearing ridges were observed among the river

features and some secondary cracks were found, similar to the impact fracture features of un-tempered lath

martensite (Fig. 11b). Macro-deformation in the WM aged at 773 K and 823 K reduced significantly (Fig. 11c

and d). The areas of shear lip and fibrous zones dropped rapidly to near zero. The fracture modes of radical

zones displayed intra-crystalline quasi-cleavage features, but the tearing ridges of rival features were flatter

than that of the as-welded condition (Fig. 11d and f), indicating that the WM had a tendency toward

embrittlement during aging at these temperatures.

15
Fig. 11. Impact fractographs of the WM aged at 773 K and 823 K for 3,000 h:
(a)–(b) before aging, (c)–(d) aged at 773 K and (e)–(f) aged at 823 K.

Figure 12 shows the impact fracture surface of WM aged at 923 K. It showed appreciable macro

deformation as in the as-welded specimen and the proportion of ductile fracture zones increased (Fig. 12a).

Dispersion precipitates identified as M23C6 carbides (discussed above) were observed at the bottoms of dimples

(Fig. 12b), indicating that M23C6 precipitated during aging at 923 K. The fracture mode of the brittle fracture

16
zone was intra-crystalline fracture with the quasi-cleavage feature (Fig. 12c), and the tearing ridges increased

slightly when compared to the as-welded condition. A few inter-granular fracture surfaces were found in this

area (highlighted with arrows). Figure 12d and e show BSE images at low magnification and high

magnification, respectively. It can be seen that bright and coarse carbides rich in W (size: 500–600 nm) were

present at the grain boundary (as shown in Fig. 12f). Nawrocki et al. (2001) investigated the evolution of

precipitation in the coarse grain heat-affected zones (CGHAZs) of T23 welds during PWHT and found that fine

and dense W2C carbides precipitated within grains after tempering at 848 K for 5 h, while coarse M23C6

carbides precipitated both at grain boundaries and inside grains after tempering at 998 K for 10 h. Hence, the

W-rich carbides in the WM cannot be W2C carbides because they precipitate at grain boundaries with a relative

large size. Thus, they are probably M6C carbides evolved from preferentially precipitated M23C6 carbides at the

grain boundary, as mentioned above. In addition, some grey Cr-rich carbides were found at grain boundaries,

which are inferred to be M23C6 carbides. Parts of the M23C6 carbides were connected to M6C carbides, proving

the transformation of M23C6 carbide to M6C carbide, consistent with the thermodynamic calculations. Dense

precipitation of M6C carbide at a minority of grain boundaries may be related to the local segregation of W

during quick solidification of the WM. The higher content of W atoms produced by segregation at grain

boundaries could satisfy the requirement of M23C6 evolving to M6C by significantly shortening the diffusion

distance of atoms, and therefore accelerating the precipitation of M6C carbide during the early stage of aging.

The content of M6C carbide was found to be low in the carbon extraction replicas because segregation occurred

at minority grain boundaries beyond the observation scale of TEM.

17
Fig. 12. Impact fractographs of the WM aged at 923 K for 3,000 h: (a) macro-scope, (b) dimple in fibrous zone, (c) cleavage in
radical zone, (d)–(e) grain boundary and (f) EDS spectrum of bright precipitate in (e).

The fracture surface of the WM aged at 873 K was similar to that aged at 923 K. Two distinctions are that

the area of the fibrous zone was smaller and the precipitates decreased in quantity and size at the bottoms of

dimples. There were also inter-granular fracture surfaces in the radical zone and precipitates at the grain

boundaries, but their size was finer.

The as-welded WM had hardness value less than 350HB and sufficient impact toughness, in agreement

18
with the research results of Bendick et al. (2007). In terms of meeting the requirement of mechanical properties

at room temperature, it is feasible to eliminate PWHT when welding T23 steel. The mechanical properties of

WM without PWHT are attributed to its martensitic/bainitic microstructure. Despite the coarse prior austenite

grain size, this mixed microstructure equally refines the initial microstructures of WM, which gives it a smaller

equivalent grain size and can retard crack propagation. Finally, this refined microstructure ensures high strength

and improves toughness.

In general, boiler components of power plants work for a long time at high temperature. The initial

mechanical properties cannot be maintained during long-term service because of the microstructural instability

of WM without PWHT. The evolution of microstructure during service may cause the degradation of

mechanical properties. The working temperatures of different components in a boiler vary on a relatively large

scale. As an example, the temperature of a water wall panel is relatively low with equivalent temperature of

673–773 K, not exceeding 823 K. Serious embrittlement of WM without PWHT takes place during long-term

service at this temperature range. From the microstructure examinations, this is possibly related to tempered

martensite embrittlement (TME). Hu and Xie (1993) summarized the mechanisms of TME: (1) intra-granular

brittleness caused by carbides distributed on the lath boundary due to the decomposition of retained austenite,

and (2) inter-granular brittleness caused by segregation of impurities or precipitation of carbides at prior

austenite grain boundaries. From the fracture morphology shown in Fig. 11, the fracture mode for the aged WM

was intra-granular but not inter-granular indicating that decomposition of retained austenite may be the

dominant mechanism. In addition, the increase in rival features and decrease of tearing ridges in the fracture

also demonstrate this. Mohyla and Foldyna (2009) reported that secondary hardening during aging may lead to

reduction of impact toughness in CrMoV steel. Secondary hardening is characterized by fine and cohesive

carbides such as VC and Mo2C acting as dispersion strengthening agents. However, secondary hardening

19
cannot be one of the main reasons for the embrittlement of T23 WM because its hardness did not increase and

almost no VC or Mo2C precipitated during aging at 773 K and 823 K.

When the aging temperature increased above 823 K the increasing diffusion rate of alloying elements

precipitates increased both at grain boundaries and inside grains in the WM. M23C6 carbides precipitated at the

grain boundaries and inside grains and V- or Nb-rich MX carbides began to precipitate inside grains, causing

the reduction of carbon content in the matrix and the degree of lattice distortion, eventually leading to

significant drop in hardness. Laths gradually disappeared and equiaxed grains formed along with the process of

recovery and recrystallization. This microstructural evolution avoids TME, consequently increasing the

toughness. The toughness of WM aged at 923 K for 3,000 h recovered to a value comparable to that of the

as-welded condition, which may be attributed to incomplete elimination of TME and premature formation of

M6C in the segregation zone of tungsten. The equilibrium calculations for the WM shown in Fig. 9a predicted

that most of the M23C6 carbides at grain boundaries would gradually transform to M6C during aging at 923 K

for times longer than 3000 h and the toughness would reduce markedly.

4. Conclusions

The microstructure and mechanical properties of WM in the as-welded condition and that aged in the

temperature range 773 K–923 K for 3,000 h were investigated to understand the mechanisms of premature

failure in T23 steel welds during service exposure. The results can be summarized as follows:

(1) T23 WM in the as-welded condition exhibits a martensitic/bainitic microstructure, ensuring that it has

hardness less than 350 HB as well as sufficient toughness.

(2) Depending on the aging temperature, long-term aging affects the mechanical properties relative to the

as-welded condition. Aging temperatures lower than 823 K cause little change in hardness but induce a

significant reduction in impact toughness. When the aging temperature is higher than 823 K, hardness begins to

20
reduce but impact toughness gradually increases.

(3) The embrittlement occurring in the WM aged at temperatures below 823 K is related to tempered

martensite embrittlement (TME). M23C6 carbides increase at grain boundaries and inside grains when the aging

temperature is higher than 823 K; V- or Nb-rich MX carbides also precipitate in grain interiors, and thus the

hardness declines and the impact toughness increases.

(4) W segregation in the WM accelerates the precipitation of W-rich carbides at grain boundaries during

aging at temperatures higher than 823 K, which retards improvement of the impact toughness during exposure.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (No. 51574181 and No.

51374153) and State Key Lab of Advanced Welding and Joining, Harbin Institute of Technology (No.

AWJ-Z15-02).

References

Bendick, W., Gabrel, J., Hahn, B., Vandenberghe, B., 2007. New low alloy heat resistant ferritic steels T/P23

and T/P24 for power plant application. International Journal of Pressure Vessels and Piping84 (1–2),

13-20.

Deng, Y. Q., Zhu, L. H., Wang, Q. J., Zhou, F. M., 2007. Effect of Microstructure Evolution During Creep on

Properties of Domestic-Made T23 Steel. Journal of Iron and Steel Research 19 (8), 46-50.

Dobrzański, J., Zieliński, A., Sroka, M., 2009. Microstructure, properties investigations and methodology of the

state evaluation of T23 (2.25 Cr-0.3 Mo-1.6 WV-Nb) steel in boilers application. Journal of Achievements

in Materials and Manufacturing Engineering32 (2), 142-153.

Hu, G. L., Xie, X. W., 1993. Heat treatment of steels, Northwestern Polytechnical University Press, Xi'an, pp.

220-222.

21
Ji, X. W., Duan, P., Li, J., 2009. Application of T23 steel to 1000 MW ultra-supercritical units. Symposium on

Sinicization of New Type steels for 600MW/1000MW Ultra-Supercritical Units, Yangzhou, China, pp.

311-316.

Komai, N., Masuyama, F., Igarashi, M., 2005. 10-Year Experience with T23 (2.25 Cr-1.6 W) and T122

(12Cr-0.4 Mo-2W) in a Power Boiler. Journal of Pressure Vessel Technology 127 (2), 190-196.

Masuyama, F., Yokoyama, T., Sawaragi, Y., Iseda, A., 1994. Development of a Tungsten-Strengthened

Low-Alloy Steel with Improved Weldability. Materials for Advanced Power Engineering (Part 1),

173-181.

Miyata, K., Sawaragi, Y., 2001. Effect of Mo and W on the phase stability of precipitates in low Cr heat

resistant steels. ISIJ International 41 (3), 281-289.

Mohyla, P., Foldyna, V., 2009. Improvement of reliability and creep resistance in advanced low-alloy steels.

Materials Science and Engineering: A 510–511, 234-237.

Nawrocki, J. G., DuPont, J. N., Marder, A. R., Robino, C. V., 2001. The postweld heat-treatment response of

simulated coarse-grained heat-affected zones in a new ferritic steel. Metallurgical and Materials

Transactions A 32 (10), 2585-2594.

Nowack, R., Götte C., Heckmann S., 2011. Quality management at RWE, using T24 boiler material as an

example. VGB Power Tech 91 (11), 1-5.

Okada, H., Igarashi, M., Yoshizawa, M., Matsumoto, S., Nakashima, T., 2011, Long-term creep properties of

2.25 Cr-1.6 W-VNbB steel (T23/P23) for fossil fired and heat recovery boilers. Advances in Materials

Technology for Fossil Power Plants: Proceedings from the Sixth International Conference, August

31-September 3, 2010, Santa Fe, New Mexico, USA. ASM International, 153-163.

Sawada, K., Tabuchi, M., Kimura K., 2009. Creep strength degradation of ASME P23/T23 steels. Materials

22
Science and Engineering: A 513, 128-137.

Vaillant, J. C., Vandenberghe, B., Hahn, B., Heuser, H., Jochum, C., 2008. T/P23, 24, 911 and 92: New grades

for advanced coal-fired power plants—Properties and experience. International Journal of Pressure Vessels

and Piping 85 (1), 38-46.

Wang, Q. J., Zhou, F. M., Deng, Y. Q., Zhu, L. H., 2006. Effect of T23 steel's microstructural Evolution on Its

Properties. Bao Steel Technology (3), 18-22.

Wang, X., Zhu, D., Hu, L., Li, X. Q., Yang, C., Ge, Z. X., Yang, Q. X., 2015. Analysis on Short-term Failure of

T23 Steel Welded Joints in Water Wall of USC Tower Boilers. Proceedings of the CSEE 35 (9), 154-161.

Zieliński, A., Golański, G., Sroka, M., Dobrzański, J., 2016. Estimation of long-term creep strength in austenitic

power plant steels. Materials Science and Technology 32 (8), 780-785.

Zieliński, A., Golański, G.,Sroka, M., Skupień, P., 2016. Microstructure and mechanical properties of the T23

steel after long-term ageing at elevated temperature. Materials at High Temperatures 33 (2), 154-163.

Zieliński, A., Golański, G., Sroka. M., 2017. Influence of long-term ageing on the microstructure and

mechanical properties of T24 steel. Materials Science and Engineering: A 682, 664-672.

Zieliński, A., Golański, G., Sroka, M., 2017. Comparing the methods in determining residual life on the basis of

creep tests of low-alloy Cr-Mo-V cast steels operated beyond the design service life. International Journal

of Pressure Vessels and Piping 152,1-6.

23
Table captions

Table 1 EDS compositions of precipitates in the WM aged at 823 K for 3,000 h (atomic fraction /%).
Table 2 EDS compositions of precipitates inside grains in the WM aged at 923 K for 3,000 h (atomic fraction /%).

24
Figure captions

Fig. 1. Fig.1.Microstructural morphologies of (a), (b) BM and (c), (d) as-welded WM.

Fig. 2. TEM images from a thin foil of as-welded WM: (a) laths, (b) three-grain boundaries.

Fig. 3. Optical micrographs of WM aged at different temperatures for 3,000 h: (a)773 K, (b) 823 K, (c) 873 K and (d) 923 K.

Fig. 4. SEM images of WM aged at different temperatures for 3,000 h: (a)773 K, (b) 723 K, (c) 873 K and (d) 923 K.

Fig. 5. TEM images from a thin foil in WM aged at 773 K and 923 K for 3,000 h: (a), (b) 773 K and (c), (d) 923 K.

Fig. 6. TEM images of precipitate from a replica of the WM aged at 823 K for 3,000 h: (a) precipitates on grain boundaries and (b)

precipitates inside grain.

Fig. 7. TEM images of precipitate inside grain from a replica of the WM aged at 923 K for 3,000 h: (a) low magnification and (b)

high magnification.

Fig. 8. TEM images of extraction replica of precipitate at the grain boundaries from the WM aged at 923 K for 3,000 h: (a) low

magnification and (b) high magnification.

Fig. 9. Calculated phase diagram of the WM of T23 steel: (a) with the precipitation of M6C carbide and (b) without the

precipitation of M6C carbide.

Fig. 10. Effect of aging temperature on the hardness and toughness of WM after 3,000 h aging: (a) hardness and (b) impact

toughness.

Fig. 11. Impact fractographs of the WM aged at 773 K and 823 K for 3,000 h: (a)–(b) before aging, (c)–(d) aged at 773 K and

(e)–(f) aged at 823 K.

Fig. 12. Impact fractographs of the WM aged at 923 K for 3,000 h: (a) macro-scope, (b) dimple in fibrous zone, (c) cleavage in

radical zone, (d)–(e) grain boundary and (f) EDS spectrum of bright precipitate in (e).

25

You might also like