Effect of heat treatment on microstructure and property of plasma-sprayed lanthanum hexaaluminate coating

Effect of heat treatment on microstructure and property of plasma-sprayed lanthanum hexaaluminate coating

Accepted Manuscript Effect of heat treatment on microstructure and property of plasma-sprayed lanthanum hexaaluminate coating Junbin Sun, Jinshuang Wa...

10MB Sizes 0 Downloads 40 Views

Accepted Manuscript Effect of heat treatment on microstructure and property of plasma-sprayed lanthanum hexaaluminate coating Junbin Sun, Jinshuang Wang, Shujuan Dong, Yu Hui, Lifen Li, Longhui Deng, Jianing Jiang, Xin Zhou, Xueqiang Cao PII:

S0925-8388(17)34562-0

DOI:

10.1016/j.jallcom.2017.12.347

Reference:

JALCOM 44433

To appear in:

Journal of Alloys and Compounds

Received Date: 12 October 2017 Revised Date:

27 December 2017

Accepted Date: 29 December 2017

Please cite this article as: J. Sun, J. Wang, S. Dong, Y. Hui, L. Li, L. Deng, J. Jiang, X. Zhou, X. Cao, Effect of heat treatment on microstructure and property of plasma-sprayed lanthanum hexaaluminate coating, Journal of Alloys and Compounds (2018), doi: 10.1016/j.jallcom.2017.12.347. 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.

ACCEPTED MANUSCRIPT Effect of heat treatment on microstructure and property of plasma-sprayed

2

lanthanum hexaaluminate coating

3

Junbin Suna, Jinshuang Wanga, Shujuan Donga, Yu Huib, Lifen Lia, Longhui Denga,

4

Jianing Jianga, Xin Zhoua, ∗, Xueqiang Caoa,* a

5

RI PT

1

State Key Laboratory of Silicate Materials for Architectures (Wuhan University of

6

Technology), China.

7

b

8

Sciences, China.

9

Abstract: Amorphous phase is commonly found in LaMgAl11O19 (LMA) coating

10

prepared by atmospheric plasma spraying. In order to improve their performance, the

11

as-sprayed LMA thermal barrier coatings on a nickel-based DZ125 substrate were

12

heat-treated at 900, 1000 and 1100°C for 5~20 h, respectively. The effect of heat

13

treatment on the microstructure, porosity, phase composition, and crystallization

14

behavior of the LMA coatings was investigated using scanning electron microscopy,

15

X-ray diffraction, differential scanning calorimetry analysis. Samples before and after

16

heat treatment for 20 h were thermally cycled at 1127°C. Results indicated that the

17

annealing of the as-sprayed LMA coating prior to thermal cycling is preferable to

18

promote the crystallization of amorphous phase. Variations in microstructure, phase

19

composition and thermal cycling lifetime of LMA coatings have been correlated to

20

the heat treatment. Heat treatment at 900°C improved the lifetime of LMA coatings,

21

whereas their thermal cycling performance did not significantly improved when the

AC C

EP

TE D

M AN U

SC

Atmospheric Environment Research Center, Shenyang Academy of Enviromental

Corresponding author. Tel/Fax: +86-27-87651856. E-mail: [email protected]. (X. Zhou), [email protected].(X. Cao) *

1

ACCEPTED MANUSCRIPT aging temperature increased to 1000 or 1100°C. The corresponding mechanism had

2

been explored by tracing the microstructure and thermo-physical properties evolutions

3

during thermal annealing process.

4

Keywords: Heat treatment; Lanthanum hexaaluminate; Amorphous phase; Thermal

5

cycling lifetime; Plasma spraying; Thermal barrier coatings

6

1 Introduction

RI PT

1

Thermal barrier coatings (TBCs) have exhibited an increasing potential for

8

improving the durability and efficiency of gas turbine engines by allowing an increase

9

in turbine inlet temperature and reducing the amount of cooling air required by the

10

hot-section components [1]. TBCs are usually made up of a two-layered structure: a

11

metallic bond coat for oxidation/corrosion resistance and a ceramic topcoat for

12

thermal protection [2, 3]. MCrAlY (M=Ni, and/or Co) is used as a bond coat to

13

provide a good thermal expansion match between the topcoat and substrate and to

14

inhibit oxidation of the substrate [4]. The current state-of-the-art of topcoat is zirconia

15

partially stabilized with 7-8 wt.% yttria (7-8 YSZ) with high coefficient of thermal

16

expansion (CTE) (10.5~11.5×10-6 K-1, 20–1200°C), fracture toughness (~2 MPa • m1/2)

17

as well as low thermal conductivity (2.1~2.2 W • m-1 • K-1), which are always

18

deposited onto the components either by electron beam-physical vapor deposition

19

(EB-PVD) or by atmospheric plasma spraying (APS) [5-7]. YSZ as the commonly

20

used TBCs material performs quite well up to 1200°C. When the material is used at

21

higher application temperature over 1200°C for a long time, it exhibits a poor service

22

lifetime due to the phase transformation [8, 9], and distinct sintering in combination

AC C

EP

TE D

M AN U

SC

7

2

ACCEPTED MANUSCRIPT with the increase of Youngꞌs modulus [10-13]. As a result, developing new TBCs

2

materials has been a hot spot of present research to improve the temperature capability

3

and durability of the deposited coatings. New TBCs candidate materials such as

4

oxides with pyrochlore (La2Zr2O7, LZ) [14], fluorite (La2Ce2O7) [15], perovskite

5

(SrZO3) [16], magnetoplumbite (LaMgAl11O19, LMA) [10, 13] structures have been

6

researched and developed.

RI PT

1

Among these new TBCs materials, LMA has essential characteristics of low

8

thermal conductivity (0.8~2.6 W • m-1 • K-1), high fracture toughness (~3.59 MPa •

9

m1/2), high CTE (9~11×10-6 K-1, 20–1200°C) and outstanding thermal stability up to

10

1600°C, owing to the magnetoplumbite structure in which spinel blocks are separated

11

by mirror planes containing the large La3+ cation [17]. With the superior structural and

12

thermochemical stability as well as at least equivalent thermophysical properties with

13

YSZ, LMA has been proposed as one of promising candidates for the next generation

14

TBCs to overcome above problems associated with the use of YSZ at higher

15

temperature [18, 19]. However, as a new TBCs candidate material, some problems are

16

urgent to be solved before it can be widely used. For instance, a large amount of

17

amorphous phase often presents in the plasma-sprayed LMA coating due to the rapid

18

quenching from the molten state [20-22]. After the LMA coating is heated to a high

19

temperature for the first time, it suffers large volume shrinkage, and plane tensile

20

stress will be developed in LMA coating due to the restriction of the superalloy

21

substrate. Besides, the recrystallization can give rise to a sudden reduction in CTE

22

value of the LMA coating [17]. Then, thermal expansion mismatches between

AC C

EP

TE D

M AN U

SC

7

3

ACCEPTED MANUSCRIPT substrate and coating materials result in the development of residual stresses during

2

thermomechanical loading, which will lead to the spallation and may put questions on

3

the reliability of the plasma-sprayed LMA coating [17, 23]. It is obvious that the

4

negative effect of amorphous phase crystallization must be removed to improve the

5

reliability and service lifetime of LMA coatings. As we all known, transition from

6

amorphous phase to crystalline phase occurs as the temperature higher or close to its

7

crystallization temperature [24]. Therefore, the negative effect of amorphous phase

8

crystallization on the LMA coating can be removed by heat treatment prior to its using.

9

On the one hand, heat treatment will induce the crystallization of the amorphous

10

phase and avoid the strong thermal stresses caused by fierce heating and cooling cycle

11

alternately. One the other hand, heat treatment also leads to the relaxation of stress

12

formed during the plasma spraying process. These are very helpful to the

13

improvement of the thermal cycling lifetime of the LMA coating in theory. Actually,

14

to the best of our knowledge, no works have been carried out on the thermal cycling

15

behavior of the LMA coatings after heat treatment.

EP

TE D

M AN U

SC

RI PT

1

Since the crystallization temperature of amorphous LMA is between 800 and

17

1200°C [17] and the upper limit temperature of superalloy DZ125 is about 1200°C,

18

therefore, the as-sprayed LMA coatings were heat-treated at 900, 1000 and 1100°C for

19

5~20 h, respectively. The effects of heat treatment on the microstructure and thermal

20

cycling behavior of the coatings were studied, and failure mechanisms of the coatings

21

were also analyzed.

22

2 Experimental

AC C

16

4

ACCEPTED MANUSCRIPT 1

2.1 Preparation of powder and coating As the chemicals in this work, the powders of La2O3, MgO and Al2O3 in proper

3

molar ratio of 0.5:1:5.5 were mixed with deionized water for 24 h by ball-milling

4

and then dried later at 100°C for 48 h. In order to synthesize the LMA powder, the

5

obtained powder mixture of La2O3, MgO and Al2O3 was heated at 1450°C for 12 h

6

and this process was replicated for two times. The as-synthesized LMA powder was

7

mixed with deionized water, Gum Arabic (adhesion agent), and ammonium citrate

8

(dispersing agent), followed by ball-milling (72 rpm) using zirconia balls for 72 h

9

and then the slurry was spray-dried (SFOC-16, Shanghai-Ohkawara Dryers Co.,

10

Ltd.). The free-flowing powders (see Fig. 1) with particle size between 20 and 125

11

µm were collected and used directly for plasma spraying without other treatments.

M AN U

SC

RI PT

2

LMA coating were produced by APS with a Multicoat Plasma Spray Unit

13

(Oerlikon Metco, Switzerland). Argon was used as primary plasma gas and powder

14

carrier gas and hydrogen was used as a secondary gas to adjust the arc voltage [25,

15

26]. The substrate was heated to 120°C using plasma flame prior to the coating

16

deposition. LMA coatings with thickness ~200 µm for thermal cycling tests were

17

prepared by APS on the DZ125 superalloy substrate (30×10×2 mm3) with NiCrAlY

18

bond coat. In addition, LMA coatings with thickness 1000~1200 µm for thermal

19

expansion test and thermal analysis were made by APS on graphite substrate without

20

bond coat. Plasma spraying parameters were listed in Table 1.

21

2.2 Heat treatment and thermal cycling test

22

AC C

EP

TE D

12

The as-sprayed LMA coatings were heat-treated in a tube furnace. The pressure 5

ACCEPTED MANUSCRIPT in the furnace was about 2~8 Pa and no oxygen was introduced to the furnace. The

2

heat treatment process was composed of heating of the coatings from room

3

temperature to 900, 1000 and 1100°C with a heating rate of 5°C/min, respectively,

4

and holding at the temperature for 5, 10, 15 and 20 h, respectively. The as-sprayed

5

LMA coating and coatings after heat treatment at 900, 1000 and 1100°C for 20 h

6

were named as C0, C900, C1000 and C1100, respectively. Thermal cycling

7

experiments, to determine TBCs cyclic lifetime, were performed in a

8

specially-designed, automated furnace. A full 35-min thermal cycle consisted of

9

rapid heating of the samples to 1127°C and holding at temperature (30 min),

10

followed by conventional cooling to near-room temperature (5 min). Each specimen

11

was examined periodically. A specimen was considered to have failed when the

12

topcoat failure area (spallation plus delamination) reached ~10% of the total topcoat

13

area. The specimen was then removed from the furnace and the number of thermal

14

cycles at which it failed was recorded. Three specimens of every kind of coating

15

were tested simultaneously and the thermal cycling lifetime was the average value of

16

the three specimens. Coatings of C0, C900, C1000 and C1100 after thermal cycling

17

failures were named as FC0, FC900, FC1000 and FC1100, respectively.

18

2.3 Characterization

SC

M AN U

TE D

EP

AC C

19

RI PT

1

Phase composition of the samples was identified by X-ray diffraction (XRD,

20

D/MAX-RB RU-200B, Cu-Kα radiation, λ=0.15406 nm) over 2θ values of 15–65°

21

with a scanning rate of 4°·min−1 by a step width of 0.02°. In addition,

22

microstructural studies in this work were performed using a field emission-scanning 6

ACCEPTED MANUSCRIPT electron microscope (SEM, QUANTA FEG 450) equipped with an energy dispersive

2

X-ray spectroscopy (EDS). All the coatings for SEM cross-section analysis were

3

embedded in a transparent epoxy resin and polished with diamond pastes, while the

4

fractured cross-sections were directly analyzed without other treatments. Coating

5

samples for thermal analyses and thermal expansion tests were prepared by abrading

6

the graphite substrate. Differential scanning calorimetry (DSC) of the LMA coatings

7

was performed on a thermoanalyzer (Schleibinger CDF Test) in air atmosphere with

8

a heating rate of 20°C·min−1. Coating samples with dimensions of (23~25)

9

mm×(5~8) mm×(1~1.2) mm were selected for thermal expansion measurements.

10

Thermal expansion of the coatings was determined between room temperature (RT)

11

and 1300°C using the Netzsch 402C high-temperature dilatometer. Amorphous

12

phase contents of the coatings were based on their amorphous indexes, which were

13

the ratios of the amorphous hump areas to the total area of the crystalline peaks and

14

humps in XRD patterns in the 2θ range between 15 and 65° [27, 28]. Porosity of the

15

coatings was evaluated by image analyses. Average thickness of coatings before and

16

after heat treatment was analyzed using SEM micrographs. About 50 measurements

17

were taken for each specimen in the thickness analysis.

18

3 Results and discussion

19

3.1 Crystallization of amorphous phase

AC C

EP

TE D

M AN U

SC

RI PT

1

20

Fig. 2 exhibits XRD patterns of LMA coatings heated at 900, 1000 and 1100°C

21

for 0~20 h, respectively. Characteristic peaks of LMA are observed obviously in these

22

four coatings, and a broad hump corresponding to amorphous phase [23, 25] is also 7

ACCEPTED MANUSCRIPT noticed (2θ=20-40°) in the XRD patterns of the as-sprayed LMA coating. With the

2

heat treatment time increasing from 5 to 20 h, the broad hump gradually flattens. As

3

shown in Fig. 2c, the broad hump seems to disappear after heat treatment at 1100°C

4

for 5 h. However, for the coating heated at 900°C, the broad hump still can be

5

observed obviously after heat treatment for 20 h (see Fig. 2a). In addition, some weak

6

peaks ascribed to LaAlO3, which cannot be observed in the as-sprayed coating, also

7

presents in the patterns of annealed LMA coatings. In fact, the LaAlO3 may be

8

covered by the high amorphous phase content in the as-sprayed coating. After heat

9

treatment, amorphous phase decreases by crystallization and LaAlO3 is detected in the

10

patterns. The amorphous phase contents of LMA coatings heated 900, 1000 and

11

1100°C for 0~20 h are presented in Fig. 3. As shown in Fig. 3, the amorphous phase

12

content decreases from 59 to 33% after heat treatment at 900°C for 20 h. However,

13

for the coating heated at 1100°C, the amorphous phase content suddenly decreases

14

from 59 to 8% after heat treatment for 5 h and then almost has no obvious changes

15

with the increase of time. It is obvious that the crystallization rate of amorphous phase

16

increases with the increase of temperature from 900 to 1100°C. In order to better

17

describe the crystallization rate, Avrami equation was adopted in this work as follows

18

[29]:

19

φ = 1 − ݁ ି௞௧ (1)

20

where k is the crystallization rate factor, n is the Avrami exponent, t is heat treatment

21

time and ϕ is the crystallization fraction. The relationship between the crystallization

22

fraction ϕ and the amorphous phase content can be described as follows:

AC C

EP

TE D

M AN U

SC

RI PT

1



8

ACCEPTED MANUSCRIPT 1

φ=

ூబ ିூ೟ ூబ

× 100%(2)

where I0 is the amorphous phase content of the as-sprayed LMA coating and It is the

3

amorphous phase content of the coating heated in different time. Fig. 4 shows a plot

4

of ϕ as the function of heat treatment time at different temperatures, which is

5

simulated according to Avrani equation. It is obvious that crystallization rate factor k

6

of the LMA coating heated at 1100°C (1.779) is far larger than that of the coatings

7

heated at 900°C (0.017) and 1000°C (0.012). In fact, crystallization rate factor k has

8

dependent on the temperature following the Arrhenius law,

9

kሺTሻ = A݁ ି౎౐ (3)

M AN U

SC

RI PT

2

ಶೌ

where A and activation energy, Ea, are kinetic parameters which are not dependent on

11

temperature. R is the ideal gas constant and T is the temperature. Therefore, k

12

increases with the increase of heat treatment temperature from 900 to 1100°C. The

13

larger crystallization rate factor k at 1100°C also can be used to explain the suddenly

14

decrease of amorphous phase after heat treatment at 1100°C for 5 h.

TE D

10

In order to further investigate the crystallization process of the amorphous phase

16

in the LMA coatings, DSC analyses of LMA coatings heated at different temperatures

17

for 20 h were carried out. As shown in Fig. 5a, there are two sharp exothermal peaks

18

at about 900 and 1170°C in the curve of C0. After heat treatment at 900°C for 20 h,

19

the first sharp peak disappears from the DSC curve of C0. As the heat treatment

20

temperature increases successively, the exothermal peak at about 1170°C gradually

21

gets flat in the DSC curves of the LMA coating, especially in the curve of C1100,

22

only a small exothermal peak is observed (see the inset of Fig. 5a). As discussed

AC C

EP

15

9

ACCEPTED MANUSCRIPT above, the crystallization level of the amorphous phase increased by increasing the

2

heat treatment temperature between 900 and 1100°C. For these reasons, two sharp

3

exothermal peaks can be due to the crystallization of the amorphous phase. In fact, the

4

same viewpoint was also proposed by Friedrich [17]. Although the second exothermal

5

peak is at about 1170°C, when the as-sprayed LMA coating is heat-treated at 1100°C

6

for 20 h, it also can accelerate the transformation to the stable phase transition [24].

7

Besides, compared with the areas of the two peaks, it can be found that the area of

8

peak at 900°C is almost 1.5 times the peak at 1170°C, which shows that the

9

crystallization of amorphous phase mostly occurs at about 900°C [30]. Although

10

about 28% amorphous phase is remained in the coating after heat treatment at 1000°C

11

for 20 h, it is hard to decrease the amorphous phase content by prolong the heat

12

treatment time owing to the kinetic limitations of crystallization [31]. Therefore, in

13

this work, LMA coatings were heated at different temperatures for 20 h before their

14

thermal cycling tests.

15

3.2 Heat capacity and thermal expansion

EP

TE D

M AN U

SC

RI PT

1

The heat capacity (Cp) curves of these coatings are illustrated in Fig. 5b. Molar

17

specific heat capacity of the as-sprayed LMA coating suffers two sharp decreases

18

around 900 and 1170°C, respectively. The sudden changes in Cp values are believed

19

to be caused by the discrete change of the heat resulted from the phase transitions of

20

the as-sprayed LMA coating during heating [23]. Besides, the crystallization of the

21

amorphous phase will induce the volume shrinkage of the coating, leading to the

22

mismatch of CTE between the ceramic coat and the bond coat. As shown in Fig. 6, C0

AC C

16

10

ACCEPTED MANUSCRIPT shows approximately linear expansion firstly, and then it suffers two significant

2

shrinkages ascribed to the crystallization of amorphous LMA phase [17]. The

3

temperature ranges where shrinkages occur are consistent with the results of thermal

4

analyses as discussed above. However, the first shrinkage disappears from the thermal

5

expansion curve of the coating C900. In addition, LMA coating almost show a linear

6

expansion from RT to 1300°C after aging at 1100°C for 20 h. In comparison with the

7

low CTE value (4.68×10-6 K-1) of C0, there is a dramatically growth in CTE value of

8

the coating after heat treatment (see Table 2). For instance, CTE value of C1100

9

reaches to 7.64×10-6 K-1. Furthermore, the first shrinkage of 2.3% is higher than the

10

second shrinkage of 0.99%, because the crystallization of the amorphous phase in the

11

LMA coating mainly occurs at about 900°C. Results show that the adverse effect of

12

the sudden change of Cp and the volume shrinkage on the as-sprayed LMA coating

13

can be removed by heat treatment prior to its using.

14

3.3 Microstructure

TE D

M AN U

SC

RI PT

1

Fig. 7 shows the fractured cross-sections SEM images of LMA coatings before

16

and after heat treatment at 900, 1000 and 1100°C for 20 h, respectively. As shown in

17

Fig. 7a, there is a large amount of amorphous phase in the as-sprayed LMA coating,

18

which is consistent with the previous report [23]. After heat treatment at 900°C for 20

19

h, some plate-like crystals (see Fig. 7b), with the different micromorphology in

20

comparison with the amorphous phase, are embedded in the coating matrix. As the

21

fractured cross-section SEM micrographs shown in Fig. 7c and Fig. 7d, it can be

22

observed that the grain growth process has been significantly accelerated and many

AC C

EP

15

11

ACCEPTED MANUSCRIPT plate-like grains present consequently in C1000 and C1100. Fig. 8 presents EDS

2

spectra and chemical compositions acquired from plate-like crystals (see Fig. 7(c, d))

3

of the LMA coatings C1000 and C1100. It is obvious that the plate-like crystals are

4

composed of La2O3-MgO-Al2O3. Besides, in our previous work [23], the

5

crystallization process was confirmed by high-resolution transmission electron

6

microscopy examination (HR-TEM). The amorphous phase as well as the crystalline

7

phase was present in the as-sprayed LMA coating. After heat treatment, the HR-TEM

8

image exhibited a typical lattice configuration of the hexaaluminate with the

9

magnetoplumbite-type structure. It indicated that a transformation of disorder to order

10

occurred to the as-sprayed LMA coating during the heat treatment process. Therefore,

11

it can be concluded that the platelet crystals are formed by the crystallization of

12

amorphous phase and the crystallization level shows a growing tendency with the

13

temperature increasing, which agrees well with the XRD analysis (see Fig. 2).

TE D

M AN U

SC

RI PT

1

Cross-sectional SEM micrographs of the LMA coatings before and after heat

15

treatment for 20 h are shown in Fig. 9. In the four TBCs systems, the LMA topcoat

16

bonds well with the bond coat. However, in comparison with the microstructure of the

17

as-sprayed coating (see Fig. 9a), coatings after heat treatment show a porous

18

microstructure with more cracks. As shown in Fig. 9b, some fine vertical cracks start

19

to be observed in C900. After increasing the heat treatment temperature further, the

20

vertical-crack tendency is more obvious. Besides, horizontal cracks also appear in

21

C1000 and C1100 (see Fig. 9(c, d)). As shown in Fig. 10, the porosity of the coating

22

increases with the heat treatment temperature increasing. On the one hand, large

AC C

EP

14

12

ACCEPTED MANUSCRIPT numbers of plate-like crystalline grains (see Fig. 7) exist in the coatings after heat

2

treatment, which is a contribution to the improvement of the porosity. On the other

3

hand, the in-plane tensile stress, generated because of thermal expansion mismatch

4

between the ceramic topcoat and the bond coat or metallic substrate, can lead to the

5

formation of vertical cracks [32]. As discussed above, the as-sprayed LMA coating

6

suffers two sudden volume shrinkages by crystallization of the amorphous phase and

7

the average CTE value is about 4.68×10-6 K-1. However, the average CTE value of the

8

nickel-base superalloy substrate is about 14×10-6~16×10-6 K-1 (20~1000°C) [33]. The

9

higher CTE of the substrate ensures that the topcoat is under significant tension at

10

high temperature [34]. For the reason, the number and size of the crack all increase

11

with the heat treatment temperature increasing. Besides, there is an obvious reduction

12

in thickness of the coating after heat treatment compared with that of the as-sprayed

13

coating. As shown in Fig. 10, coatings heated at different temperatures show different

14

thickness such that C900>C1000>C1100. The decrease in thickness is in good

15

agreement with the thermal expansion analysis of the LMA coatings.

16

3.4 Thermal cycling behavior

SC

M AN U

TE D

EP

AC C

17

RI PT

1

Thermal cycling lifetimes of these four coatings are compared in Fig. 11. C900

18

has the longest thermal cycling lifetime (126 cycles), which has been improved by

19

25% compared with C0 (100 cycles). C1000 and C1100 (102 cycles) have similar

20

lifetime with C0. Thermal cycling results indicate that the heat treatment temperature

21

plays an important role in the thermal cycling lifetime of the LMA coatings.

22

Fig. 12 shows the surface morphology of the coatings after thermal cycling 13

ACCEPTED MANUSCRIPT failure. As seen from FC0, visible cracks are found evidently (noted by the black

2

frame). In addition, due to the extreme heating and cooling conditions encountered at

3

the edges, obvious surface spallation is observed on the edge of the sample and more

4

than 10% of surface area spalled after failure. The surface photo indicates the failure

5

of the LMA coatings starts at the edge, and the spallation seems to occur instantly

6

during cycling.

RI PT

1

Fig. 13 exhibits XRD patterns of the coatings after thermal failures. Compared

8

with coatings before thermal cycling tests (see Fig. 2), the broad hump

9

corresponding to amorphous phase has disappeared in the XRD patterns of all these

10

coatings, especially in the pattern of FC0. Although the LMA coatings were

11

heat-treated at different temperatures for 20 h, the crystallization of the amorphous

12

phase still occurs during thermal cycling. Besides, LaAlO3 phase disappears from

13

the XRD patterns of coatings after thermal cycling tests, indicating that LaAlO3

14

phase may re-form to LMA accompanying the crystallization of the amorphous

15

phase during the thermal cycling.

EP

TE D

M AN U

SC

7

Fig. 14 shows cross-sectional SEM images of coatings after thermal cycling

17

failures. In comparison with the microstructures of coatings before thermal cycling

18

tests, large vertical and horizontal cracks are observed in the failed coatings. From

19

Fig.14, the failure of these four LMA coatings occurred by the formation of large

20

horizontal crack in the ceramic top coat which was close to the interface of ceramic

21

top coat and bond coat. In addition, some dark substances with the thickness of 3~5

22

µm are found at the interface between the ceramic coat and the bond coat of FC900

AC C

16

14

ACCEPTED MANUSCRIPT (see the A zone in Fig. 14b) and FC1100 (see the B zone in Fig. 14d), which may be

2

TGO. It is well known that TGO is the result of selective oxidation of elements

3

within the bond coat. The oxidation of the bond coat is conclusively a result of

4

oxygen transport through the interconnected pores within the porous plasma-sprayed

5

ceramic topcoat during thermal cycling tests. In order to demonstrate the existence of

6

TGO, elemental distributions by map scanning at the topcoat-bond coat interface of

7

FC1100 is determined and their results are shown in Fig. 15(b-g). Cr and Al should

8

be uniformly distributed throughout the bond coat. However, after thermal cycling

9

failure there are high levels of Cr (see Fig. 15f) and Al (see Fig. 15c) at the interface

10

between the bond coat and the ceramic topcoat. From the EDS results, it is obvious

11

that TGO has formed in C900 and C1100 during thermal cycling, and its

12

composition may be Cr2O3 and Al2O3.

TE D

M AN U

SC

RI PT

1

Generally, stresses caused by the thermal expansion mismatch and the growth of

14

thermally grown oxides (TGO) lead to the failure of TBCs during thermal cycling. As

15

discussed above, the as-sprayed LMA coating with large amount of amorphous phase

16

has the lowest CTE (4.68×10-6 K-1). Therefore, it suffers the largest thermal stress

17

caused by mismatch of CTEs between the ceramic topcoat and the substrate, which is

18

enough to lead the failure of C0. For the heat-treated LMA coatings, the amorphous

19

phase content decreases dramatically during heat treatment. LMA coatings of C900,

20

C1000 and C1100 have the larger CTEs (between 7.02 and 7.64×10-6 K-1) than that of

21

C0. This effectively mitigates the thermal expansion mismatch stress level. Besides,

22

since the annealing process is preferable to promote the crystallization of amorphous

AC C

EP

13

15

ACCEPTED MANUSCRIPT phase in a mild condition, the higher stress level which will be caused by the rapid

2

recrystallization and the platelet-like LMA grain growth can be avoided. Those can

3

well explain the improvement of thermal durability for C0. However, the emergency

4

of the large cracks in C1000 and C1100 decreases the bond strength of the coating and

5

promote the growth of TGO, which is adverse to the improvement of thermal shock

6

resistance of the coating. Stresses caused by TGO growth also contribute to the failure

7

of these coatings. For these reasons, only the thermal cycling lifetime of C900 has

8

been improved dramatically, and C1000 and C1100 have no obvious improvement in

9

thermal cycling lifetime compared with C0.

SC

M AN U

10

RI PT

1

4. Conclusions

LaMgAl11O19 (LMA) coatings were prepared by plasma spraying and then were

12

heat-treated at 900, 1000 and 1100°C for 0~20 h, respectively. With the heat

13

treatment increasing from 900 to 1100°C, the crystallization rate of the amorphous

14

phase increases. During the heat treatment process, the crystallization of the

15

amorphous phase induces the volume shrinkage, the increase of porosity, and the

16

change in heat capacity (Cp) value of the LMA coating. The changes are more obvious

17

by increasing the heat treatment temperature. Heat treatment prior to use of the

18

plasma-sprayed LMA coating at 900°C weakens the negative effect of the amorphous

19

phase and improves the thermal shock resistance of the coating. In comparison with

20

thermal cycling lifetime (100 cycles) of C0, C900 has an obvious improvement of

21

thermal cycling lifetime (126 cycles). However, C1000 and C1100 with the same

22

thermal cycling lifetime (102 cycles) are similar to that of C0.

AC C

EP

TE D

11

16

ACCEPTED MANUSCRIPT The mismatch in coefficient of thermal expansion (CTE) between the bond coat

2

and the LMA topcoat leads to the failure of the as-sprayed LMA coating (C0). For

3

coatings heated at 900 (C900), 1000 (C1000) and 1100 (C1100), the mismatch of

4

CTE is decreased effectively by recrystallization. The growth of TGO and the weak

5

bond strength lead to the failure of these coatings.

6

Acknowledgement

RI PT

1

The authors thank Dr. Zhitao Shan for the DSC measurements. This work was

8

supported by the National Natural Science Foundation of China (No. 51501137,

9

51702244) and the Natural Science Foundation of Hubei Province (No.

M AN U

EP

TE D

2017CFB285).

AC C

10

SC

7

17

ACCEPTED MANUSCRIPT References

AC C

EP

TE D

M AN U

SC

RI PT

[1] C.G. Zhou, N. Wang, H.B. Xu, Comparison of thermal cycling behavior of plasma-sprayed nanostructured and traditional thermal barrier coatings, Mater. Sci. Eng. A, 452 (2007) 569-574. [2] N.P. Padture, M. Gell, E.H. Jordan, Thermal barrier coatings for gas-turbine engine applications, Science, 296 (2002) 280-284. [3] L.H. Gao, H.B. Guo, L.L. Wei, C.Y. Li, S.K. Gong, H.B. Xu, Microstructure and mechanical properties of yttria stabilized zirconia coatings prepared by plasma spray physical vapor deposition, Ceram. Int., 41 (2015) 8305-8311. [4] C.G. Zhou, N. Wang, Z.B. Wang, S.K. Gong, H.B. Xu, Thermal cycling life and thermal diffusivity of a plasma-sprayed nanostructured thermal barrier coating, Scripta Mater., 51 (2004) 945-948. [5] L. Wang, Y. Wang, X.G. Sun, J.Q. He, Z.Y. Pan, C.H. Wang, Microstructure and indentation mechanical properties of plasma sprayed nano-bimodal and conventional ZrO2–8wt%Y2O3 thermal barrier coatings, Vacuum, 86 (2012) 1174-1185. [6] W.G. Mao, J.M. Luo, C.Y. Dai, Y.G. Shen, Effect of heat treatment on deformation and mechanical properties of 8 mol% yttria-stabilized zirconia by Berkovich nanoindentation, Appl. Surf. Sci., 338 (2015) 92-98. [7] K.D. Bouzakis, A. Lontos, N. Michailidis, Determination of mechanical properties of electron beam-physical vapor deposition-thermal barrier coatings (EB-PVD-TBCs) by means of nanoindention and impact testing, Surf. Coat. Technol., 163 (2003) 75-80. [8] R. Vaßen, F. Cernuschi, G. Rizzi, A. Scrivani, N. Markocsan, L. Östergren, A. Kloosterman, R. Mevrel, J. Feist, J. Nicholls, Overview in the Field of Thermal Barrier Coatings Including Burner Rig Testing in the European Union, Adv. Eng. Mater., 43 (2008) 371-382. [9] E. Withey, C. Petorak, R. Trice, G. Dickinson, T. Taylor, Design of 7wt.% Y2O3–ZrO2/mullite plasma-sprayed composite coatings for increased creep resistance, J. Eur. Ceram. Soc., 27 (2007) 4675–4683. [10] H. Lu, C.A. Wang, C. Zhang, S. Tong, Thermo-physical properties of rare-earth hexaaluminates LnMgAl11O19 (Ln: La, Pr, Nd, Sm, Eu and Gd) magnetoplumbite for advanced thermal barrier coatings, J. Eur. Ceram. Soc., 35 (2015) 1297-1306. [11] Z.G. Liu, J.H. Ouyang, Y. Zhou, J. Li, X.L. Xia, Influence of ytterbium- and samarium-oxides codoping on structure and thermal conductivity of zirconate ceramics, J. Eur. Ceram. Soc., 29 (2009) 647-652. [12] Z.G. Liu, J.H. Ouyang, Y. Zhou, X.L. Xia, Hot corrosion behavior of V2O5-coated Gd2Zr2O7 ceramic in air at 700–850°C, J. Eur. Ceram. Soc., 66 (2009) CD008046-CD008046. [13] X. Chen, Y. Zhao, X. Fan, Y. Liu, B. Zou, Y. Wang, H. Ma, X. Cao, Thermal cycling failure of new LaMgAl11O19/YSZ double ceramic top coat thermal barrier coating systems, Surf. Coat. Technol., 205 (2011) 3293-3300. [14] R. Vaßen, F. Traeger, D. Stöver, New Thermal Barrier Coatings Based on Pyrochlore/YSZ Double-Layer Systems, Int. J. Appl. Ceram. Technol., 1 (2004) 18

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

351–361. [15] X.Q. Cao, R. Vassen, W. Fischer, F. Tietz, W. Jungen, D. Stoever, Lanthanum–Cerium Oxide as a Thermal Barrier‐Coating Material for High‐ Temperature Applications, Adv. Mater., 15 (2003) 1438-1442. [16] W. Ma, D.E. Mack, R. Vassen, D. Stöver, Perovskite-Type Strontium Zirconate as a New Material for Thermal Barrier Coatings, J. Am. Ceram. Soc., 91 (2008) 2630–2635. [17] C. Friedrich, R. Gadow, T. Schirmer, Lanthanum hexaaluminate—a new material for atmospheric plasma spraying of advanced thermal barrier coatings, J. Therm. Spray Technol., 10 (2001) 592-598. [18] R. Gadow, M. Lischka, Lanthanum hexaaluminate — novel thermal barrier coatings for gas turbine applications — materials and process development, Surf. Coat. Technol., s 151–152 (2002) 392-399. [19] P. Jana, P.S. Jayan, S. Mandal, K. Biswas, Microstructural design of neodymium-doped lanthanum–magnesium hexaaluminate synthesized by aqueous sol–gel process, J. Mater. Sci., 50 (2015) 344-353. [20] X.Q. Cao, Y.F. Zhang, J.F. Zhang, X.H. Zhong, Y. Wang, H.M. Ma, Z.H. Xu, L.M. He, F. Lu, Failure of the plasma-sprayed coating of lanthanum hexaluminate, J. Eur. Ceram. Soc., 28 (2008) 1979-1986. [21] X.L. Chen, Y.F. Zhang, X.H. Zhong, Z.H. Xu, J.F. Zhang, Y.L. Cheng, Y. Zhao, Y.J. Liu, X.Z. Fan, Y. Wang, H.M. Ma, X.Q. Cao, Thermal cycling behaviors of the plasma sprayed thermal barrier coatings of hexaluminates with magnetoplumbite structure, J. Eur. Ceram. Soc., 30 (2010) 1649-1657. [22] R. Vaßen, M.O. Jarligo, T. Steinke, D.E. Mack, D. Stöver, Overview on advanced thermal barrier coatings, Surf. Coat. Technol., 205 (2010) 938-942. [23] X.L. Chen, Y. Zhao, W.Z. Huang, H.M. Ma, B.L. Zou, Y. Wang, X.Q. Cao, Thermal aging behavior of plasma sprayed LaMgAl11O19 thermal barrier coating, J. Eur. Ceram. Soc., 31 (2011) 2285-2294. [24] V.M. Fokin, E.D. Zanotto, N.S. Yuritsyn, J.W.P. Schmelzer, Homogeneous crystal nucleation in silicate glasses: A 40 years perspective, J. Non-Cryst. Solids, 352 (2006) 2681-2714. [25] H.Z. Liu, J.H. Ouyang, Z.G. Liu, Y.M. Wang, Microstructure, thermal shock resistance and thermal emissivity of plasma sprayed LaMAl11O19 (M = Mg, Fe) coatings for metallic thermal protection systems, Appl. Surf. Sci., 271 (2013) 52-59. [26] J.S. Wang, J.B. Sun, H. Zhang, S.J. Dong, J.N. Jiang, L.H. Deng, X. Zhou, X.Q. Cao, Effect of spraying power on microstructure and property of nanostructured YSZ thermal barrier coatings, J. Alloy Compd., 730 (2018) 471-482. [27] F. Tarasi, Suspension plasma sprayed alumina-yttria stabilized zirconia nano-composite thermal barrier coatings: formation and roles of the amorphous phase, in, Concordia University, 2010. [28] F. Tarasi, M. Medraj, A. Dolatabadi, R.S. Lima, C. Moreau, Thermal Cycling of Suspension Plasma Sprayed Alumina- YSZ Coatings Containing Amorphous Phases, J. Am. Ceram. Soc., 95 (2012) 2614-2621. [29] E.B. Peixoto, E.C. Mendonça, S.G. Mercena, A.C.B. Jesus, C.C.S. Barbosa, C.T. 19

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

Meneses, J.G.S. Duque, R.A.G. Silva, Study of the dynamic of crystallization of an amorphous Fe40Ni40P14B6 ribbon through Johnson-Mehl-Avrami model, J. Alloy Compd., 731 (2018) 1275-1279. [30] L.L. Huang, H.M. Meng, J. Tang, Crystallization behavior of plasma-sprayed lanthanide magnesium hexaaluminate coatings, Int. J. Miner. Metall. Mater., 21 (2014) 1247-1253. [31] A. Rosenflanz, M. Frey, B. Endres, T. Anderson, E. Richards, C. Schardt, Bulk glasses and ultrahard nanoceramics based on alumina and rare-earth oxides, Nature, 430 (2004) 761. [32] A.C. Fox, T.W. Clyne, Oxygen transport by gas permeation through the zirconia layer in plasma sprayed thermal barrier coatings, Surf. Coat. Technol., 184 (2004) 311-321. [33] X.Q. Cao, R. Vassen, D. Stoever, Ceramic materials for thermal barrier coatings, J. Eur. Ceram. Soc., 24 (2004) 1-10. [34] Y. Bai, L. Zhao, J.J. Tang, S.Q. Ma, C.H. Ding, J.F. Yang, L. Yu, Z.H. Han, Influence of original powders on the microstructure and properties of thermal barrier coatings deposited by supersonic atmospheric plasma spraying, part II: Properties, Ceram. Int., 39 (2013) 5113–5124.

20

ACCEPTED MANUSCRIPT

M AN U

SC

RI PT

Figures

AC C

EP

TE D

Fig. 1. XRD pattern and SEM image of the starting powder.

21

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

Fig. 2. XRD patterns of LMA coatings before and after heat treatment at different temperatures: (a) 900°C; (b) 1000°C; (c) 1100°C. A, B, C, D, E represent the heat treatment time of 0, 5, 10, 15, 20 h, respectively.

22

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

Fig. 3. Amorphous phase content of LMA coatings heated at 900, 1000 and 1100°C in different time.

23

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

Fig. 4. Crystallization fraction of LMA coatings heated at 900, 1000 and 1100°C in different time.

24

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

Fig. 5. DSC curves of coatings (a) and heat capacities (Cp) curves (b) of the LMA coating before and after heat treatment at different temperatures. The insets in (a) and (b) are the DSC and Cp curves of C1100 with magnification, respectively.

25

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

Fig. 6. Thermal expansion curves of the LMA coatings before and after heat treatment at different temperatures.

26

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

Fig. 7. SEM micrographs of the fractured cross-sections of coatings: (a) C0; (b)-(bꞌ) C900 with the different magnifications of 5000× and 20000×; (c) C1000; (d)-(dꞌ) C1100 with the different magnifications of 5000× and 20000×.

27

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

Fig. 8. EDS spectra acquired from plate-like crystals of the LMA coating after heat treatment at different temperatures for 20 h: (a) 1000°C; (b) 1100°C.

28

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

Fig. 9. Cross-sectional SEM micrographs of the LMA coating: (a) C0; (b) C900; (c) C1000 and (d) C1100.

29

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

Fig. 10. Porosity and thickness of the LMA coatings before and after heat treatment at different temperatures.

30

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

Fig. 11. Thermal cycling lifetime of LMA coatings before and after heat treatment at different temperatures.

31

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

Fig. 12. Surface photos of LMA coatings before and after thermal cycling tests at 1127°C.

32

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

Fig. 13. XRD patterns of LMA coatings after thermal cycling failure.

33

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

Fig. 14. Cross-sectional SEM micrographs of LMA coatings after thermal cycling failure: (a) FC0; (b) FC900; (c) FC1000; (d) FC1100.

34

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

Fig. 15. SEM images of the elemental maps for FC1100.

35

ACCEPTED MANUSCRIPT List of Table Table 1 Plasma spraying parameters Current

Power

Plasma gas,

Carrier gas,

Powder feeding rate

(mm)

(A)

(kW)

(SLPM*)

Ar (SLPM)

(%)

100

620

42

35/12

3.2

AC C

EP

TE D

M AN U

SC

*SLPM: standard liter per minute.

RI PT

Spray distance

36

15

ACCEPTED MANUSCRIPT Table 2 Measured coefficients of thermal expansion for the LMA coatings before and after heat treatment at different temperatures Average CTE (×10-6 K-1)

C0

4.68 (RT to 736°C)

C900

7.02 (RT to 1014°C)

C1000

7.08 (RT to 1096°C)

C1100

7.64 (RT to 1224°C)

AC C

EP

TE D

M AN U

SC

RI PT

Coating

37

ACCEPTED MANUSCRIPT Highlights • LaMgAl11O19 (LMA) coatings were pretreated at different temperatures. • Microstructures, porosities, and thermal properties of the coatings were studied. • Large cracks were observed in coatings after heat treatment at 1000 and 1100°C.

AC C

EP

TE D

M AN U

SC

RI PT

• Thermal cycling lifetime of the coating heated at 900°C was increased by 25%.