You are on page 1of 13

Materials and Design 196 (2020) 109138

Contents lists available at ScienceDirect

Materials and Design

journal homepage: www.elsevier.com/locate/matdes

Sliver defect formation in single crystal Ni-based superalloy castings


Wenliang Xu a,1, Fu Wang a,b,⁎,1, Dexin Ma b, Xintao Zhu b, Dichen Li a, Andreas Bührig-Polaczek b
a
State Key Laboratory for Manufacturing System Engineering, School of Mechanical Engineering, Xi'an Jiaotong University, Xi'an 710049, China
b
Foundry Institute, RWTH Aachen University, Aachen 52072, Germany

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• During directional solidification of SX


castings, dendrite fragmentation arose
the formation of sliver defect.
• The macroscopic and microscopic non-
uniform stress played significant roles
in the dendrite fragmentation.
• The possible solid fraction range for
fragmentation was suggested to be
0.6–0.8.
• The sliver occurrence may be correlated
to the formation of converging interface
between dendrite tips and mold wall.

a r t i c l e i n f o a b s t r a c t

Article history: Sliver is a typical grain defect in single crystal Ni-based superalloy castings produced by Bridgman directional so-
Received 27 July 2020 lidification. However, its formation mechanism has yet to be fully understood. In this study, the formation of
Received in revised form 2 September 2020 sliver defects was investigated with respect to the defect morphological features, the microstructure of defect ini-
Accepted 8 September 2020
tiation points, and the solidification conditions. Experiment results showed that sliver defects were originated
Available online 9 September 2020
from fragmentation of dendrite trunks and could developed to various morphologies. The misorientation range
Keywords:
of sliver grains was between 3.2° to 12°. Microstructural analysis results indicated that the dendrites fragmented
Single crystal Ni-based superalloys abruptly without significant plastic deformation. And the dendrite fragmentation could be facilitated by oxide in-
Directional solidification clusions and solidification pores. It was also found that the sliver defects were prone to occur at the interface
Dendritic solidification where primary dendrite converge to the mold inner wall. Based on the low misorientation level of sliver grains
Sliver defect in the experimental observation, the critical solid volume fraction range for fragmentation was suggested to be
0.6–0.8. The dendrite fragmentation was attributed to the non-uniform stress distribution in dendrite during so-
lidification. In addition, the effect of airfoil geometry of a typical blade casting and defect prevention methodology
were discussed.
© 2020 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://
creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction

Single crystal (SX) Ni-based superalloys have been widely applied to


produce turbine blades for advanced gas turbine engines [1,2]. At pres-
ent, the Bridgman directional solidification (DS) is the major production
⁎ Corresponding author at: State Key Laboratory for Manufacturing System technique of SX blades. With the demands on the blade durability and
Engineering, School of Mechanical Engineering, Xi'an Jiaotong University, Xi'an 710049,
China.
reliability increasing, high crystal perfection of blade casting is being
E-mail address: fuwang@xjtu.edu.cn (F. Wang). pursued. During the DS process, many factors including alloy composi-
1
The two authors contributed equally to this work. tion, process parameters and blade geometries have impacts on the

https://doi.org/10.1016/j.matdes.2020.109138
0264-1275/© 2020 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
W. Xu, F. Wang, D. Ma et al. Materials and Design 196 (2020) 109138

grain growth. Hence the solidification process should be well controlled geometric sensitive. However, its formation mechanism remains less
to avoid grain defects which destroy the microstructural integrity by in- well understood owing to the complexity of dendrite growth behavior
troducing grain boundaries (GBs). Owing to the SX Ni-based superalloys during solidification. It was observed that slivers would occur not only
are lack of GB strengthening elements, the GBs usually become the in specific regions on castings, but also in the airfoil region without
weakest positions under high temperature operating condition, structural discontinuity. Therefore, more efforts should be made to pro-
resulting in degradation of blade mechanical properties [3,4]. vide new insights into the defect formation mechanism. In this work,
Over the past few decades, great efforts have been made to provide sliver morphological features, misorientation range, and microstructure
insight to the grain defect formation [5–12], including stray grain (SG), of defect initiation points were examined. The effect of solidification
freckle, and low angle grain boundary (LAGB). However, limited study condition and airfoil geometries on the dendrite misorientation were
on sliver can be found in literature. Sliver is generally characterized as also analyzed. This investigation provided a direct experiment evidence
the misoriented grain extending along the solidification direction. for dendrite deformation during directional solidification and a better
Their formation and evolution are not fully understood until now. In understanding of the sliver defect formation mechanism.
the early study on the process window of single crystal castings [12],
slivers were considered as stray grains arise from the undesired nucle- 2. Experimental procedures
ation in the undercooling melt at the front of the solid/liquid interface,
due to poor process condition. It has been also reported that slivers usu- 2.1. Experiment and characterization
ally originated at the inner surface defects of ceramic mold. The
misoriented grain formation could be attributed to the heterogeneous Various SX castings were prepared by Bridgman directional solidifi-
nucleation facilitated by mold surface defects [13]. However, the nucle- cation (DS) combined with grain selection technique. Commercial
ated grains are normally of random orientation [14], whereas the major- nickel-based superalloy CMSX-4 was used in the experiments. The DS
ity of sliver grains exhibit low deviation from the matrix. Thus, the experiments were conducted with fixed process parameters in an in-
nucleation mechanism may be difficult to support the observation. dustrial Bridgman furnace (ALD Vacuum Technologies, Inc.). The mold
It has been suggested that dendrite deformation in the mushy zone preheating temperature and the alloy melt pouring temperature were
during the solidification process could cause sliver defects. Various fac- 1500 °C. A withdrawal rate of 3 mm min−1 was used. After the DS pro-
tors including the instability of temperature field, the complex shape of cess, the castings were cleaned and macro-etched by an etchant (a
castings, the fluid flow during solidification, oxide inclusions, and ther- mixed solution of 50 vol% H2O2 and 50 vol% HCl) to reveal the grain
mal stress have been correlated to the dendrite misorientation. Newell morphology. Then the castings with sliver defect were picked out for in-
et al. [7] found that dendrites exhibited random orientation deviation vestigation. For microstructural observation, the regions containing
under steady growth conditions and the maximum deviation was ap- sliver defect were extracted by electro discharged machining. The sam-
proximately 2.3°, while those under non-steady-state growth condi- ple surfaces with sliver defect were carefully polished and etched using
tions had the accumulated misorientation up to 6°. The increased a solution containing 20 mL C2H5OH, 40 mL HCl, and 2 g CuSO4‧5H2O.
misorientation was attributed to the plastic deformation of dendrites The dendritic structure was characterized by light microscopy (LM,
within the mushy zone. In addition, Siredey et al. [15] suggested that Zeiss Axio Vert.A 1 and Stemi 305). The microstructures of grain bound-
the thermal stress caused by precipitation process of the γ' phase gave aries were observed by a scanning electron microscope (SEM, Zeiss
rise to the dendrite deformation. The study by Xie et al. [16] on solidifi- GeminiSEM 500). Energy Dispersive Spectroscopy (EDS) analysis was
cation process of large-scale SX blades confirmed that the dendrite array performed to investigate the element distribution within the casting
in the longitudinal direction gradually became disordered with the samples. After the LM and SEM observation, the samples were polished
height of casting increase during the DS process. And the occurrence using an iron beam milling system (Leica EM RES102) For electron back-
of dendrite misorientation tended to increase with the height. scatter diffraction (EBSD) analysis. The crystal orientation was mea-
Hallensleben et al. [17] provided an experimental evidence for dendrite sured by using the HKL channel 5 package according to the EBSD data
bending by measuring the dendrite growth deviation angles. Their re- collected by an EBSD detector (Oxford instruments). Furthermore, a
sults showed that the dendrite continuous bending angle could reach thin specimen across the GB between misoriented dendrite and matrix
1.7° over a distance of 25 mm. From above mentioned studies, it can was prepared by focus ion beam technique (FIB, Tescan Lyra3) and then
be summarized that the dendrite orientation would gradually vary dur- observed by a transmission electron microscope (TEM, FEI F200X Talos).
ing successive growing. The FIB specimen was also examined by a EBSD system in TEM in order
Furthermore, Carney et al. [18] reported that slivers were frequently to reveal the GB features.
observed where cross section of castings reduced. And the formation of
defect was attributed to oxide inclusions and the enhanced shrinkage 2.2. Numerical simulation
driven fluid flow. However, recent studies [19–21] showed that the
growing dendrites would be deformed at the enlarged cross section po- The directional solidification process of blade casting was simulated
sition due to local thermal stress concentration, and the deformed den- using a commercial finite element software (ProCAST, ESI Group), in
drites further developed to misorientation grains. An in-depth study by order to investigate the solidification condition for defect formation. Fi-
Aveson et al. [22] suggested that slivers arose as a result of the local nite element model was built according to the actual dimensions of fur-
thermal stress at a sensitive position on the casting (the thin channel nace, ceramic shell, and blade casting. The mold temperature, alloy
upon seed) during seeded Bridgman DS process. They quantified the melting temperature, and withdrawal rate used in simulation were con-
dendrite misorientation using a tailored analysis method and evaluated sistent with those applied in the experiment. A 1/6 model of the six-
the stress field by numerical simulation. The experiment results indi- blade cluster mold was used in consideration of the calculation effi-
cated that the differential thermal contraction between the mold and ciency. The thermophysical parameters of CMSX-4 superalloy was ob-
the semi-solid alloy produced bending moments and torques on the tained from ProCAST calculation and Ref. [23], as shown in Table 1.
growing dendrites. The dendrites at the mold wall deflected and finally Some interpolation and extrapolation have been used to estimate the
developed to sliver grains. The defect formation was also correlated to data up to 1500 °C. Other simulation parameters were properly defined
the geometric relationship between the dendrite growth direction and based on literatures [24–26] and temperature measurement in our pre-
the inner surface of mold wall. vious study [27]. In the stress simulation for alloy, isotropic elasto-
According to the previous studies, sliver is characterized as a type of plastic model was employed. The mechanical properties of superalloy
grain defect formed by misoriented dendrite on the casting surface. The in [001] orientation were used in the simulation, thus the anisotropy
defects observed at specific positions indicated its formation is [28,29] was ignored. The availability of this model to semi-

2
W. Xu, F. Wang, D. Ma et al. Materials and Design 196 (2020) 109138

Table 1
Material data of superalloy from ProCAST calculation and Ref. [23] (20−1500 °C).

Density (g/cm3) Thermal expansion (×10−6 K−1) Poison's ratio Young's modulus (GPa) Yield stress (MPa) Ultimate stress (MPa)

7.5~8.5 6~19 0.36~0.5 0.5~127 0~960 0~1195

quantitatively study the stress and strain characteristics of casting dur- shown in Fig. 2(b–d). The inversed pole figures show that its orientation
ing directional solidification has been checked by previous studies in the longitudinal growth direction was almost the same as that of the
[20,23,24]. In addition, the ceramic shell was set as a rigid body in matrix, whereas the orientation varied in the other two directions, indi-
order to eliminate shell deformation, thus the stress in simulation cating that this misoriented dendrite trunk rotated around the crystal
mainly reflects the thermal stress caused by shrinkage. For evaluation axis during solidification. The point to origin misorientation profile
of the stress in mushy zone, the stress calculation initiated at solid frac- along the dendrite growth direction shows that the sliver grain deviated
tion (fs) of 0.5. from the matrix at about 3.2° and kept a constant misorientation (Fig. 2
(d)).
It is interesting to note that the misorientation of dendrite trunks
3. Results and discussion
within matrix was identified by EBSD measurement, as shown in
Fig. 2(b), and the typical low angle grain boundaries (LAGBs) were indi-
3.1. Defect morphology
cated by white dash lines. A slightly misoriented dendrite adjacent to
the sliver was adopted for comparison of misorientation level. By
3.1.1. Defect classification
EBSD line scanning across the LAGBs, the misorientation values of the
Based on the grain defect inspection of many SX castings, it was
two dendrite trunks were measured to be around 1°. These dendrites
found that the slivers had various morphologies. They were linear sur-
may deflect during solidification and then fracture after the deformation
face defects extending along the solidification direction. The width
reaches a critical value. This result indicates that the sliver defect origi-
ranged from a single dendrite to several dendrites, and the length
nated from a segment of misoriented dendrite trunk and the dendrite
ranged from a few millimeters to tens of millimeters. For most of the de-
segment detached from its parent arm owing to deformation.
fects, the initial point can be clearly identified. The sliver defects can be
classified into three types according to the morphological features. The
first type was a subgrain formed by a segment of dendrite arm, the sec- 3.1.3. Defect evolution
ond type was that containing an elongated misoriented dendrite, and An elongated sliver (close to the leading edge of a dummy blade)
the third was a developed grain containing a cluster of misaligned den- was adopted to explain the defect evolution, as shown in Fig. 3(a).
drites which derived from a misoriented dendrite, as shown in Fig. 1(a– This sliver was observed at the side where dendrites converged to the
c). Generally, the development of sliver grains causes more serious mold wall. It was found that the sliver originated from a misoriented
damage to the crystal imperfection of SX castings. dendrite and developed to a narrow misoriented grain. Competitive
growth between the sliver and matrix dendrite trunks was observed.
3.1.2. Defect initial point The sliver dendrite was eliminated after it developed to a height of ap-
An undeveloped sliver and its EBSD results are shown in Fig. 2. This proximately 40 mm. The defect initial site can be readily observed, indi-
defect was observed near the trailing edge on a blade convex surface. cating that the grain has an abrupt change in orientation (Fig. 3(b)).
The dendrite growth direction converged to the mold wall and the den- Compared with the growing direction of matrix dendrites, this
drite trunks showed a continuous growth morphology. The dendrite misoriented dendrite deviated to the right side at approximately 8°.
array grew unidirectionally such that the misorientation defect was Therefore, a diverging GB is formed at the left side, while a converging
hard to be found, as shown in Fig. 2(a). It was observed that a part of pri- GB formed at the right side, as shown in Fig. 3(c). During directional
mary dendrite differed from the matrix in optical contrast. The EBSD growth, the secondary arms (S) of the misoriented primary dendrite
measurement confirmed this dendrite segment were about 1.5 mm in trunk (P) acquired sufficient space to branch ternary arms (T) at the di-
length and had an abrupt change in crystallographic orientation, as verging GB. The sliver grain grew upwards by deriving higher order

Fig. 1. (a) A sliver defect formed by a segment of dendrite (type 1) on blade convex surface. (b) Longitudinally developed sliver defects (type 2) near blade leading edge. (c) A
longitudinally and horizontally developed sliver defect (type 3) on blade concave surface.

3
W. Xu, F. Wang, D. Ma et al. Materials and Design 196 (2020) 109138

Fig. 2. (a) Macrostructures of a sliver grain and a deflected dendrite. (b) EBSD orientation image map. Typical LAGBs along dendrite trunks are highlighted by the white dash lines.
(c) Corresponding inverse pole figure. (d) Point to origin misorientation line scanning along the white arrow in (b).

arms to form new longitudinal dendrite trunks (Fig. 3(c) and (d)) at this terminated owing to branching was restricted and the dendrite tip
side. However, at the converging GB, the secondary arms were blocked was geometrically constrained by the matrix dendrites and mold wall
by matrix dendrites. This observation was consistent with the widely (Fig. 3(e)).
accepted model of competitive grain growth [30,31]. In fact, the den- The development of sliver defects can be correlated to the local
drites competitively grew in three-dimensions. In the observation growing condition, including orientation [30], thermal field [31], and
plane of this misoriented dendrite, the longitudinal growth was growing space [20]. Thus, if the misoriented dendrite was under a

Fig. 3. (a) Overview of a typical longitudinally developed silver defect. (b) Defect initiation point. (c) and (d) Defect developing morphologies (in sequence). (e) Defect terminal.

4
W. Xu, F. Wang, D. Ma et al. Materials and Design 196 (2020) 109138

preferred condition for transverse and longitudinal propagation, it could angle grain boundaries and show less morphological correlation to the
develop to a columnar grain with large size in both of cross section and ordered dendrites.
length. For instance, the developed sliver (type 3) shown in Fig. 1(c). It
is interesting to note that this sliver even grew through the platform and 3.2. Microstructure analysis
into the blade root region.
3.2.1. Defect origin
The morphology and orientation measurements of a typical sliver
3.1.4. Misorientation features defect are shown in Fig. 5. This sliver was observed near the trailing
Morphology analysis and crystallographic orientation measure- edge on a blade convex surface. It can be observed that the dendrite
ments were carried out for a variety of sliver defects, in order to inves- growth direction converged to the mold wall and a segment of primary
tigate the misorientation features. Misorientation of typical slivers is dendrite trunk with a length of about 4 mm detached from the original
shown in Fig. 4(a), the measurements were conducted within the defect dendrite trunk. The location of the fracture can be clearly observed and
origin region. The deflected dendrite and the sliver defect shown in the dendrite array showed successive growth morphology, indicating
Fig. 2 were also adopted for comparison (Fig. 4(b) and (c)). Two typical the dendrite detached without significant movement, as shown in
slivers and corresponding inverse pole figures are shown in Fig. 4(d)– Fig. 5(a) and (b). The EBSD analysis results illustrated this segment of
(g), respectively. It was found that most of the slivers formed at the dendrite maintained the original orientation in longitudinal direction,
side where dendrites converged to the mold wall. These slivers derived but slightly deviated in the secondary orientation, in comparison to
from misoriented dendrite trunks and showed relatively small devia- the matrix (Fig. 5(c)).
tion with respect to matrix. And the sliver orientations changed The local magnified SEM image of sliver origin shows a layer of
abruptly rather than deviated with an accumulative misorientation coarse γ' precipitated along the GB and the coarse γ' were approxi-
from the matrix. The abrupt-changed misorientation angles of sliver mately 2 mm in length and 1 mm in width (Fig. 5(d)). The morphology
grains ranged from 3.2° to 11.4°, whereas the typical accumulative mis- illustrates that the sliver originated from a fragmented dendrite and the
orientation of dendrite bending was reported to be 0.68°/cm [17]. Addi- detached dendrite rejoined with the matrix during solidification. A TEM
tionally, the misorientation could be observed in all the three sample was cut off at the fragmentation interface by FIB for microstruc-
crystallographic orientation ([001], [010], and [100]), indicating that ture analysis (Fig. 5(e)). It was observed that the γ' precipitates near GB
the sliver dendrites deviated from the matrix three-dimensionally within the TEM sample showed coarse and irregular morphology,
(Fig. 4(f) and (g)). which were different to those previous observed in Fig. 5(d), as
The experiment results support the defect formation mechanism shown in Fig. 5(f). This could be attributed to the variation of GB mor-
that a sliver originates from a segment of dendrite detached from the phology within different section.
matrix and indicate the angle of 3.2° may be the critical value for den-
drite fracture. Moreover, the small deviation partly explained why 3.2.2. Validation of dendrite deformation
slivers could co-grow with the matrix dendrites for a certain distance. The TEM sample was analyzed by EBSD technique in order to deter-
In comparison to freckle, which is also a type of surface defect arising mine the local misorientation near the GB at a microscopic scale. As
from dendrite fragmentation during directional solidification, slivers shown in Fig. 6 (a) and (b), the GB was not straight but exhibited a zig-
generally occur individually and show successive growing morphol- zag morphology. In the light of γ' precipitates morphology shown in
ogies similar to normal dendrites of matrix. Whereas freckles are char- Fig. 5(f), it was found that the irregular γ' precipitates distributed at
acterized as discontinuous and long chains composed by fine equiaxed each side of the GB and their edges were constrained by the GB. This re-
grains. Owing to the strong thermosolutal convection, the orientations sult indicates that the precipitates occurred after the GB formed.
of freckle grains usually deviate from the matrix significantly and ran- The local misorientation map and corresponding orientation distri-
domly. Some freckle grains could develop to elongated grains similar bution range were illustrated in Fig. 6(c) and (d), respectively. It was ob-
to slivers in morphology [10,32]. However, they usually form high served that each side of GB has no significant accumulative orientation,

Fig. 4. (a) Misorientation distribution of typical slivers. (b)–(e) Defect morphologies of sample 1, 2, 4, 8, respectively. (f) and (g) Inverse pole figures for sample 4 and 8, respectively.

5
W. Xu, F. Wang, D. Ma et al. Materials and Design 196 (2020) 109138

Fig. 5. (a) Macrostructure of a sliver defect. (b) Magnified image of defect origin. (c) EBSD orientation image map (inverse pole fig. X and Y). (d) SEM image of GB. (e) FIB extracting position
at the GB. (f) FIB prepared sample showing the microstructure of GB region.

but maintained the inherent orientation, indicating that the dendrite measured to be about 5.7°. The pole figure of the investigated region
misorientation can be characterized by brittlely fragment of dendrite shows that the orientation deviation mainly occurred in the secondary
rather than dendrite deform with large plastic strain. The point to origin orientation of dendrite, as shown in Fig. 6(f), which was consistent
misorientation line scanning result along the white arrow in Fig. 6 with the EBSD measurement results in dendrite scale (Fig. 5(c)).
(b) reveals that the misorientation value jumped at the GB (Fig. 6(e)). The selected area electron diffraction (SAED) result of the GB region
The orientation difference between the original grain and sliver was in TEM sample suggests that the GB angle was about 5.7° and formed by

Fig. 6. (a) Band contrast image of the GB region within TEM sample. (b) Corresponding EBSD orientation image map. (c) Local misorientation map. (d) Local misorientation distribution.
(e) EBSD line scanning result along the white arrow showing in (b). (f) Corresponding pole figure.

6
W. Xu, F. Wang, D. Ma et al. Materials and Design 196 (2020) 109138

Fig. 7. (a) TEM image exhibiting the GB and (b) selected area electron diffraction (SAED) pattern of the GB region (white circle in (a)).

a grain rotating around the [001] zone axis (Fig. 7). This result confirms were discovered after mechanical polishing (pointed out by yellow ar-
that the segment of dendrite trunk rotated during solidification and rows). Such pores were formed by insufficient liquid feeding into the
finally fragmented and developed to a sliver defect in turn. interdendritic region during the last stage of solidification and pro-
A high-resolution TEM (HRTEM) image of the GB region and magni- moted by solidification shrinkage [36]. The SEM morphology of a typical
fied images of selected regions are shown in Fig. 8(a–f). Fast Fourier pore is shown in Fig. 10(b). It was observed that the microstructure was
Transform (FFT) was applied to the high-resolution image and then In- separated and partially rejoined. In the rejoined region, complex micro-
verse Fast Fourier Transform (IFFT) was applied to the paired (200) structure including coarse γ', oxide inclusions, and cracks was observed,
spots in the FFT spectrum in order to obtain the corresponding lattice as shown in Fig. 10(c). EDS map shows the oxides contained alumina,
fringe image of the crystallography plane, as shown in Fig. 8(g–k). The which may come from mold surface or oxide film during melting
GB shows tilt characteristics, with periodic dislocation can be observed (Fig. 10(d)). In the separation region near the pore, successive micro-
at the GB, as shown in Fig. 8(b) and (g). These dislocations were the ge- structural tearing was observed. The morphology of a crack with
ometry necessary dislocations (GNDs) to form the GB. Furthermore, it rough interface shown in Fig. 10(c) indicates that the crack occurred
should be noted that various dislocations and lattice distortions can be after solidification and was caused by thermal stress during cooling. Fur-
observed near the GB. These crystallographic defects were caused by a thermore, the surface within the pore shows the presence of many
certain amount of plastic deformation, which would be correlated to spherical particles, which were considered to be the low melting point
solidification stress. The white dotted circle in the IFFT figure shows a phases formed at the last stage of solidification [37], suggesting that
dislocation dipole (Fig. 8(i)). And dislocation loops were also observed, the pore formed during solidification (Fig. 10(f)). This observation re-
as shown in white circle in Fig. 8(j). It is interesting to note that the ob- sult demonstrates that dendrites would be subjected to complex stress
served dislocation density was lower than that within the as-cast micro- during solidification due to the presence of inclusions and pores, and the
structure which would cause recrystallization [24]. The critical plastic dendrite fragmentation could be induced by the stress. Additionally,
strain introduced in solidification process (above 1000 °C) for recrystal- whether the pore formation before or after dendrite deformation is
lization of CMSX-4 superalloy was estimated to be 1.5–2.0%, which is at not clear according to the present experimental results. Both of the
a relatively low level of deformation. Therefore, it suggests that the den- two situation would cause higher susceptibility of dendrite fragmenta-
drites were vulnerable during solidification and the required plastic tion. Further investigation is required to improve the understanding of
deformation would be quite small for the fragmentation. the interaction between the pores and slivers.
Besides, a tertiary dendrite arm detached from its parent trunk was So far, only the deviation of crystallographic orientation can be
observed, as shown in Fig. 9, further confirming that dendrites could found in previous studies on dendrite deformation, while the typical mi-
fragment mechanically and give rise to sliver defects. The dendrite frag- crostructure to reveal the mechanical deformation mechanism has been
mentation occurred at the junction between the primary and secondary rarely reported. The direct observation of dendrite deformation by X-
arm. And the tertiary arm exhibited a few misorientation level, indicat- ray radiation imaging during the growing process [38] provided an ex-
ing that it detached after branching and was mechanically constrained perimental evidence for the dendrite deformation, however it did not
by neighboring dendrites. This observation is consistent with the analy- identify the inducing factors of defect formation and examine the mi-
sis of deformation susceptibility of dendrite arms by Aveson [22] and crostructure features of the defect initiation point. The experimental re-
Takaki [33]. Niederberger [34] indicated that misorientation could be sults in this section support the sliver formation mechanism which
more concentrated in dendrite branching necks due to the increased involves the dendrite mechanical deformation and fragmentation.
local stress in such thinnest sections.
According to the above analysis, dendrites would mechanically 3.3. Critical solid fraction range
deform and fragment if the stress exceeds the mechanical strength
limit [35], resulting in the formation of sliver grains. In this study, it The small deviation range of slivers suggests that the dendrites
was also found that oxide inclusions and pores could facilitate the sliver should be constrained by neighbor dendrites and have limited deflec-
formation. For instance, a longitudinal developed sliver defect accompa- tion space if they detached from the parent dendrites during solidifica-
nied with oxide inclusions and pores was observed on the casting sur- tion. It would be reasonable to assume that the dendrite deflection
face (pointed out by a black arrow in Fig. 10(a)). The sliver grain had a occurred at the last stage of solidification, at which the solid phase in
misorientation angle of 11.1° (Sample 7 in Fig. 4). The irregular pores the mushy zone reached a relatively high content, and the dendrite

7
W. Xu, F. Wang, D. Ma et al. Materials and Design 196 (2020) 109138

Fig. 8. (a) High resolution TEM image of the GB region. (b–f) Magnified image of selected region in a. (g–k) Corresponding IFFT images of (b-f) constructed by using the (200) frequencies
in the FFT image.

network formation was almost complete. At this stage, the inter- reached 0.9~0.94, the dendrites were not bridged but surrounded by a
dendrite permeability sharply decreased [36], therefore the dendrite very thin liquid film. Hence the dendrites were vulnerable to the me-
deformation may not be caused by the liquid flow. However, the shrink- chanical stress at this stage. At the bottom of mushy zone, the dendrites
age and deformation of the fully solidified part of the casting would be coalesced with each other and acquired sufficient strength to withstand
transferred to the solid-liquid phase in the mushy zone [39]. Wang the solidification stress. Whereas the dendrites at the top of mushy zone
et al. [40] indicated that when the solid fraction (fs) in the mushy zone were surrounded by liquid and had few mechanical constrains, such

8
W. Xu, F. Wang, D. Ma et al. Materials and Design 196 (2020) 109138

occurred when the solid fraction reached 0.6 to 0.8 during directional
solidification.

3.4. Solidification deformation analysis

According to the above analysis, the dendrites in the mushy zone


would be affected by solidification deformation. Moreover, the observed
dendrite misorientation showed a strong tendency at the converging
side where the dendrites grew to the mold inner wall. This result is con-
sistent with the observations of Aveson et al. [22]. It may be attributed to
an extra load or constraint on dendrite arm due to the interaction with
the mold wall. Bogdanowicz et al. [42] reported that the mechanical in-
Fig. 9. The sliver defect formed by a misoriented tertiary dendrite arm (pointed out by a
red arrow). The defect originated from the thin neck of a secondary dendrite arm
teraction between growing dendrites and mold wall could induce a
indicating the dendrite deflection was associated with stress concentration. forced orientation effect on dendrites at the converging side of thin-
walled blade, resulting in the misorientation defect.
In order to investigate the influence of solidification deformation, the
solidification process was analyzed by ProCAST simulation. The simulated
thermal profile of airfoil region during solidification process is shown in
that they were less affected by solidification deformation. Note that the Fig. 12(a). It was found that a concave solidification front formed. And
freckles are prone to form at the top of mushy zone due to high perme- the solidification front inclined to the heater side due to the convex surface
ability for liquid flow [32]. of airfoil faced to the heater. The curved solidification front could influence
Our previous work [41] on the directional solidification characteris- the dendrite growth in two aspects [17,43]. First, primary dendrites tend to
tics of CMSX-4 superalloy shows that the secondary dendrite arms par- grow perpendicular to the liquid isotherm. Hence dendrites within high
tially contacted when the solid fraction reached above 0.6. At a solid curvature region may be prone to deform. Second, a concave solidification
fraction of 0.8, the secondary dendrite arms coarsened and overlapped front would assist secondary dendrite arms to preferentially branch to-
between the primary dendrite arms, resulting in a morphology similar ward the casting center. This may result in overgrowth of sliver grains
to that of the solidification completion. Fig. 11(a) is a schematic diagram from outer region to the casting central region. In the present study, no suf-
of dendrites in the mushy zone showing solid volume fraction change ficient evidence was found to indicate that the high curvature inevitably
during directional solidification. The transverse sections of mushy leads to sliver formation or offers slivers more favorable growth conditions.
zone at solid fraction of 0.4, 0.6, and 0.8 are shown in Fig. 11(b), (c), In consequence, there is a need to further identify the influence of thermal
and (d), respectively. The dendrite morphologies of different solid frac- field on sliver formation and evolution.
tions combined with the low deviation level of sliver grains (mentioned The stress contours within an airfoil cross section at a solid fraction
in 3.1.4 section) suggest that dendrite fragmentation causing slivers of around 0.7 is shown in Fig. 12(b). Owing to the unstable thermal

Fig. 10. (a) Macrostructure of the sliver defect. (b) SEM image of a typical pore. (c) Magnified image of the rejoined region in (b). (d) EDS mapping analysis result on O and Al distribution
for the region shown in (c). (e) Crack interface showing microstructural tearing. (f) The inner surface of a pore.

9
W. Xu, F. Wang, D. Ma et al. Materials and Design 196 (2020) 109138

Fig. 11. (a) Schematic of dendrites in the mushy zone showing solid volume fraction (fs) change during directional solidification. (b–d) Transverse sections of mushy zone at solid fraction
of 0.4, 0.6, and 0.8, respectively [41].

Fig. 12. (a) thermal profile of airfoil during solidification process. (b) Effective stress contour in mushy zone at solid fraction of ~0.7. (c–e) Corresponding displacement of X, Y, and Z,
respectively. (f) Schematic illustration of the component forces applied on the dendrite near casting surface.

10
W. Xu, F. Wang, D. Ma et al. Materials and Design 196 (2020) 109138

Fig. 13. (a) Schematic of typical blade cross-sections, (b) schematic of blade cross sections (S1 and S2), and (c) longitudinal section (S3).

field, the effective stress within mushy zone distributed non-uniformly. the growing dendrites contact with the mold wall due to the airfoil ge-
The corresponding displacement in X, Y, and Z axis are shown in Fig. 12 ometry. As shown in a typical longitudinal cross-section (S3) parallel to
(c), (d), and (e), respectively. It can be observed that the casting had the vertical axis, the airfoil surface had an inclination (Fig. 13(c)). If the
non-uniform deformation in both transverse and longitudinal direc- primary dendrites grow parallel to the vertical axis, a converging inter-
tions during solidification. The airfoil tended to shrink inward and face will form on the convex surface, while a diverging interface will
downward. Thus, the complex shrinkage should have impact on the form on the concave surface. The converging and diverging angle
dendrites in mushy zone. Therefore, the vulnerable dendrites are were denoted by α and β (Fig. 13(c)), respectively. This model predicts
prone to deflect if suffer from an unbalanced stress field. For an individ- that the sliver has a higher probability to occur on the convex surface,
ual dendrite on the casting surface, the complex interaction with sur- owing to the dendrite deformation susceptibility at converging inter-
rounding dendrites and mold wall involves in the deformation face. The occurrence of sliver defects near the leading edge on the con-
process. The stress acted on the dendrite could be considered as the re- vex surface, as shown in Fig. 1(a), was consistent with the prediction. In
sultant of multiple stresses in different directions and equivalents to a this region, the local dendrite growth direction slightly converged to the
stress parallel to Z axis (σZ, tensile or compression) and a stress in X-Y mold wall (almost parallel to the surface), according to the observation
plane (σXY, shear), as shown in Fig. 12(f). The inward and downward so- of dendrite growth orientation within casting. In addition, other factors
lidification shrinkage contributes to the resultant stress. For instance, including the grain orientation, solid-liquid interface morphology, and
downward shrinkage leads to the tensile stress on dendrite arm, while the size of dendrite arms [20,47,48] would also affect the dendrite de-
the torque in X-Y plane which could arise from inward shrinkage as flection as well as sliver formation. Thus, the sliver defect could occur
well as the interaction between dendrite and mold wall results in the at the concave surface, as shown in Fig. 1(c), when the selected grain
rotation of dendrite trunk. orientation largely deviates to this surface (the interface becomes
Another factors to take into consideration for the dendrite deflection converging).
is the width of mushy zone. The width of mushy zone is inversely pro- Statistics on sliver defects in the SX blade casting production showed
portional to the longitudinal thermal gradient within mushy zone. Dur- that various slivers occurred more frequently on the convex surface,
ing the solidification process of SX blade by conventional DS method, whereas a few defects were detected on the concave surface [49]. To
the thermal gradient would gradually decrease with the increase of some extent, the influence of airfoil structure discussed in this section
casting height, hence the mushy zone will become wider [26,44]. More- can be used to explain the defect formation susceptibility in the indus-
over, the solidification front on shadow side usually possesses a wider trial production. Besides, several studies [14,50,51] reported that slivers
mushy zone [45,46]. With lower longitudinal thermal gradient within could occur at the diverging interface. Their formation mechanisms may
mushy zone, the stay period of dendrites in the vulnerable zone be slight different, but they are still associated with dendrite deforma-
would be prolonged. Hence the risk of dendrite deflection further tion. For fully understand the dendrite deformation mechanism, further
increases. investigation is required.

3.5. Effect of airfoil geometry 3.6. Defect prevention

The airfoil of turbine blade possesses complex three-dimensional Previous studies have reported that low angle grain boundaries de-
curved surfaces, i.e. convex surface and concave surface. In the direction teriorate mechanical properties, particularly the GB angle exceeds 9°
along the blade vertical axis, the cross section from the airfoil tip to the [52]. The GB misorientation tolerance of several alloys is even lower
bottom is characterized by a continuous twist. Within the airfoil region than 6° [53]. According to the experimental results, the misorientation
of the blade casting shown in Fig. 1, the typical cross-sections (S1 and of sliver defects could exceed 9°. Some fragmented dendrites would de-
S2) stacked along the blade vertical axis (Z) from the tip to bottom velop to narrow misoriented grains or large three-dimensional colum-
(Fig. 13(a)). It can be seen that the twist made the leading edge contour nar grains to damage the SX structure. Generally, misorientation
moved to the side of convex surface, while the position of trailing edge tolerance and defect occurrence position are established to allow the
changed insignificantly (Fig. 13(b)). presence of slivers in order to reduce the production cost of SX blades.
In the directional solidification process of SX blade, the selected However, for new generation engines demanding higher durability, it
grain is usually introduced from the blade tip and the withdrawal direc- is necessary to minimize the occurrence of the sliver defects within SX
tion is parallel to the blade vertical axis. Therefore, it is inevitable that turbine blades. The formation and evolution mechanism of such defect

11
W. Xu, F. Wang, D. Ma et al. Materials and Design 196 (2020) 109138

is complex, which involves solidification stress, temperature field, ge- acknowledge the support of the Youth Innovation Team of Shaanxi
ometry, crystal orientation, etc. Eliminating the defects is still challeng- Universities.
ing. Several process optimization techniques can be taken into account
to reduce the rejection of castings. Except for grain orientation control References
technique [22], optimization of process parameters [26] as well as the
[1] R.C. Reed, T. Tao, N. Warnken, Alloys-by-design: application to nickel-based single
casting arrangement in cluster mold [54] are needed in order to reduce crystal superalloys, Acta Mater. 57 (2009) 5898–5913.
the width of mushy zone and the inclination of solid–liquid interface. [2] T.M. Pollock, S. Tin, Nickel-based superalloys for advanced turbine engines: chemis-
Several modification of DS equipment, such as square heater [3] and ra- try, microstructure, and properties, J. Propuls. Power 22 (2006) 361–374.
[3] Innovations in casting techniques for single crystal turbine blades of superalloys, in:
diation baffle [46] can be considered to reduce the curvature of solidifi- D. Ma, F. Wang, Q. Wu, S. Bogner, A. Bührig-Polaczek, M. Hardy, et al., (Eds.), Super-
cation front. Moreover, the mold quality must be strictly controlled, alloys 2016, TMS, Warrende 2016, pp. 237–246.
especially the inner surface of the mold should be smooth and clean, [4] J. Zhang, L. Lou, Basic research in development and application of cast superalloy,
Acta Metall. Sin. 54 (2018) 1637–1652.
in order to prevent the localized stress during solidification. Also, reduc-
[5] R.E. Napolitano, R.J. Schaefer, Convergence-fault mechanism for low-angle boundary
tion of oxide inclusions in melting and pouring process is necessary. formation in single-crystal castings, J. Mater. Sci. 35 (2000) 1641–1659.
[6] M. Newell, N. D’Souza, N.R. Green, Formation of low angle boundaries in Ni-based
4. Conclusions superalloys, Int. J. Cast Met. Res. 22 (2009) 66–69.
[7] M. Newell, K. Devendra, P.A. Jennings, N. D’Souza, Role of dendrite branching and
growth kinetics in the formation of low angle boundaries in Ni-base superalloys,
In the present work, we investigated the formation of sliver defects Mater. Sci. Eng. A 412 (2005) 307–315.
in single crystal Ni-based superalloy castings by analyzing the macro/ [8] P. Hallensleben, H. Schaar, P. Thome, N. Jöns, A. Jafarizadeh, I. Steinbach, G. Eggeler, J.
Frenzel, On the evolution of cast microstructures during processing of single crystal
micro morphologies of defects and the solidification conditions. The Ni-base superalloys using a Bridgman seed technique, Mater. Des. 128 (2017)
main conclusions can be summarized as follows: 98–111.
[9] W. Bogdanowicz, J. Krawczyk, A. Tondos, J. Sieniawski, Subgrain boundaries in single
(1) During directional solidification of SX castings, dendrite frag- crystal blade airfoil of aircraft engine, Cryst. Res. Technol. 52 (2017) article no.:
mentation arose the formation of sliver defect. The misorienta- 1600372.
[10] L. Qin, J. Shen, G.X. Yang, Q.D. Li, Z. Shang, A design of non-uniform thickness mould
tion of sliver grain ranged from 3.2° to 12°. The minimum
for controlling temperature gradient and S/L interface shape in directionally solidi-
misorientation of 3.2° may be the critical value for dendrite frag- fied superalloy blade, Mater. Des. 116 (2017) 565–576.
mentation. The EBSD orientation measurement and TEM obser- [11] D.X. Ma, A. Bührig-Polaczek, Application of a heat conductor technique in the pro-
vation of dendrite with slight deviation indicated that the duction of single-crystal turbine blades, Metall. Mater. Trans. B Process Metall.
Mater. Process. Sci. 40 (2009) 738–748.
dendrites fragmented brittlely rather than with large plastic de- [12] A. De Bussac, C.A. Gandin, Prediction of a process window for the investment casting
formation. of dendritic single crystals, Mater. Sci. Eng. A 237 (1997) 35–42.
(2) The dendrite fragmentation can be attributed to macroscopic and [13] Z.X. Shi, S.Z. Liu, J.R. Li, Sliver fomation mechanism of single crystal superalloy during
directional solidification process, Hot Work. Technol. 42 (2013) 31–33.
microscopic non-uniform stress. And the dendrite deformation [14] Y.Z. Zhou, Formation of stray grains during directional solidification of a nickel-
included torsion and bending. Moreover, the oxide inclusions based superalloy, Scr. Mater. 65 (2011) 281–284.
and pores could induce dendrite fragmentation. [15] N. Siredey, M. Boufoussi, S. Denis, J. Lacaze, Dendritic growth and crystalline quality
of nickel-base single grains, J. Cryst. Growth 130 (1993) 132–146.
(3) The low misorientation level of sliver grain indicated that the [16] G. Xie, S.H. Zhang, W. Zheng, G. Zhang, J. Shen, Y.Z. Lu, H.Q. Hao, L. Wang, L.H. Lou, J.
dendrite fragmentation occurred when the solid fraction of Zhang, Formation and evolution of low angle grain boundary in large-scale single
mushy zone reached a high level. According to the dendrite mor- crystal superalloy blade, Acta Metall. Sin. 55 (2019) 1527–1536.
[17] P. Hallensleben, F. Scholz, P. Thome, H. Schaar, I. Steinbach, G. Eggeler, J. Frenzel, On
phologies of different solid fractions in mushy zone, the possible
crystal mosaicity in single crystal ni-based superalloys, Crystals 9 (2019) article no.:
solid fraction range for fragmentation was suggested to be 149.
0.6–0.8. [18] The origin of sliver defects in single crystal turbine blades, in: C.A. Carney, J. Beech, J.
(4) The observation in this experiment showed slivers frequently oc- Beech (Eds.), Solidification Processing 1997, Depatment of Engineering Materials,
University of Sheffield, Sheffield 1997, pp. 33–36.
curred at the side where the primary dendrites converged to the [19] W. Bogdanowicz, R. Albrecht, J. Sieniawski, K. Kubiak, The subgrain structure in tur-
mold inner wall. The sliver defects on the blade convex surface bine blade roots of CMSX-4 superalloy, J. Cryst. Growth 401 (2014) 418–422.
may be correlated to the formation of converging interface. [20] M. Huo, L. Liu, W.C. Yang, D.J. Sun, S.S. Hu, J. Zhang, H.Z. Fu, Formation of slivers in
the extended cross-section platforms of Ni-based single crystal superalloy, Adv.
Eng. Mater. 20 (2018) 1701189, https://doi.org/10.1002/adem.201701189.
[21] D.J. Sun, L. Liu, T.W. Huang, W.C. Yang, C. He, Z.R. Li, J. Zhang, H.Z. Fu, Formation of
Credit author statement lateral sliver defects in the platform region of single-crystal superalloy turbine
blades, Metall. Mater. Trans. A Phys. Metall. Mater. Sci. 50 (2019) 1119–1124.
[22] J.W. Aveson, P.A. Tennant, B.J. Foss, B.A. Shollock, H.J. Stone, N. D’Souza, On the origin
Wenliang Xu: Data curation, Formal analysis, Investigation, Writing - of sliver defects in single crystal investment castings, Acta Mater. 61 (2013)
original draft. Fu Wang: Methodology, Funding acquisition, Writing - 5162–5171.
[23] C. Panwisawas, J.C. Gebelin, R.C. Reed, Analysis of the mechanical deformation aris-
original draft. Dexin Ma: Conceptualization, Validation, Writing - review ing from investment casting of directionally solidified nickel-based superalloys,
& editing. Xintao Zhu: Data curation, Investigation. Dichen Li: Supervi- Mater. Sci. Technol. 29 (2013) 843–853.
sion, Project administration. Andreas Bührig–Polaczek: Project [24] C. Panwisawas, H. Mathur, J.C. Gebelin, D. Putman, C.M.F. Rae, R.C. Reed, Prediction
of recrystallization in investment cast single-crystal superalloys, Acta Mater. 61
administration. (2013) 51–66.
[25] Y.F. Li, L. Liu, D.J. Sun, Q.C. Yue, T.W. Huang, B. Gan, J. Zhang, H.Z. Fu, Quantitative
analysis of withdrawal rate on stray grain formation in the platforms of a Ni-
based single crystal dummy blade, J. Alloys Compd. 773 (2019) 432–442.
Declaration of Competing Interest [26] D. Szeliga, Effect of processing parameters and shape of blade on the solidification of
single-crystal CMSX-4 Ni-based superalloy, Metall. Mater. Trans. B Process Metall.
The authors declare that they have no known competing financial Mater. Process. Sci. 49 (2018) 2550–2570.
[27] F. Wang, Z.N. Wu, C. Huang, D.X. Ma, J. Jakumeit, A. Bührig-Polaczek, Three-
interests or personal relationships that could have appeared to influ- dimensional dendrite growth within the shrouds of single crystal blades of a
ence the work reported in this paper. nickel-based superalloy, Metall. Mater. Trans. A 48 (2017) 5924–5939.
[28] L. Zhang, L. Zhao, A. Roy, V.V. Silberschmidt, G. McColvin, Low-cycle fatigue of single
crystal nickel-based superalloy – mechanical testing and TEM characterisation,
Acknowledgements
Mater. Sci. Eng. A 744 (2019) 538–547.
[29] L. Zhang, L. Zhao, R. Jiang, C. Bullough, Crystal plasticity finite-element modelling of
This work was supported by the National Natural Science Founda- cyclic deformation and crack initiation in a nickel-based single-crystal superalloy
tion of China (No. 91860103) and the Key Program of Equipment Pre- under low-cycle fatigue, Fatigue Fract. Eng. Mater. Struct. 43 (2020) 1769–1783.
[30] X.B. Meng, Q. Lu, X.L. Zhang, J.G. Li, Z.Q. Chen, Y.H. Wang, Y.Z. Zhou, T. Jin, X.F. Sun,
Research Field Foundation of China (No. 61409230405). The authors Z.Q. Hu, Mechanism of competitive growth during directional solidification of a

12
W. Xu, F. Wang, D. Ma et al. Materials and Design 196 (2020) 109138

nickel-base superalloy in a three-dimensional reference frame, Acta Mater. 60 [43] A.J. Elliott, S. Tin, W.T. King, S.C. Huang, M.F.X. Gigliotti, T.M. Pollock, Directional so-
(2012) 3965–3975. lidification of large superalloy castings with radiation and liquid-metal cooling: a
[31] C.B. Yang, L. Liu, X.B. Zhao, N. Wang, J. Zhang, H.Z. Fu, Competitive grain growth comparative assessment, Metall. Mater. Trans. A 35 (2004) 3221–3231.
mechanism in three dimensions during directional solidification of a nickel-based [44] G. Matache, D.M. Stefanescu, C. Puscasu, E. Alexandrescu, A. Bührig-Polaczek, Inves-
superalloy, J. Alloys Compd. 578 (2013) 577–584. tigation of solidification microstructure of single crystal CMSX-4 superalloy – exper-
[32] D.Y. Han, W.G. Jiang, J.H. Xiao, K.W. Li, Y.Z. Lu, W. Zheng, Investigation on freckle for- imental measurements and modelling predictions, Int. J. Cast Met. Res. 28 (2015)
mation and evolution of single-crystal nickel-based superalloy specimens with dif- 323–336.
ferent thicknesses and abrupt cross-section changes, J. Alloys Compd. 805 (2019) [45] W.L. Xu, F. Wang, D.X. Ma, A. Bührig-Polaczek, Effect of Ru on macro-/micro-struc-
218–228. ture evolution within platform of Ni-based superalloy single crystal blades, J. Alloys
[33] T. Takaki, M. Ohno, T. Shimokawabe, T. Aoki, Two-dimensional phase-field simula- Compd. 817 (2020) article no.: 153337.
tions of dendrite competitive growth during the directional solidification of a binary [46] D. Szeliga, K. Gancarczyk, W. Ziaja, The control of solidification of Ni-based superal-
alloy bicrystal, Acta Mater. 81 (2014) 272–283. loy single-crystal blade by mold design modification using inner radiation baffle,
Adv. Eng. Mater. 20 (2018) article no.: 1700973.
[34] C. Niederberger, J. Michler, A. Jacot, Origin of intragranular crystallographic misori-
[47] D.J. Sun, L. Liu, W.C. Yang, T.W. Huang, M. Huo, S.S. Hu, J. Zhang, H.Z. Fu, Influence of
entations in hot-dip Al-Zn-Si coatings, Acta Mater. 56 (2008) 4002–4011.
secondary dendrite orientation on the evolution of misorientation in the platform
[35] D.G. Eskin, Suyitno, L. Katgerman, Mechanical properties in the semi-solid state and region of single crystal superalloy turbine blades, Adv. Eng. Mater. 21 (2019) article
hot tearing of aluminium alloys, Prog. Mater. Sci. 49 (2004) 629–711. no.: 1800933.
[36] E. Plancher, P. Gravier, E. Chauvet, J. Blandin, E. Boller, G. Martin, L. Salvo, P. Lhuissier, [48] S.S. Hu, L. Liu, W.C. Yang, D.J. Sun, M. Huo, T.W. Huang, J. Zhang, H.J. Su, H.Z. Fu, For-
Tracking pores during solidification of a Ni-based superalloy using 4D synchrotron mation of accumulated misorientation during directional solidification of Ni-based
microtomography, Acta Mater. 181 (2019) 1–9. single-crystal superalloys, Metall. Mater. Trans. A 50 (2019) 1607–1610.
[37] Z.P. Zhou, L. Huang, Y.J. Shang, Y.P. Li, L. Jiang, Q. Lei, Causes analysis on cracks in [49] D.X. Ma, F. Wang, H.Y. Sun, W.T. Xu, Study on sliver defects in single crystal castings
nickel-based single crystal superalloy fabricated by laser powder deposition addi- of superalloys, Foundry 68 (2019) 567–573.
tive manufacturing, Mater. Des. 160 (2018) 1238–1249. [50] Z.X. Shi, J.R. Li, S.Z. Liu, J.Q. Zhao, Formation mechanism of low angle boundary of
[38] J.W. Aveson, G. Reinhart, H. Nguyen-Thi, N. Mangelinck-Noël, A. Tandjaoui, B. Billia, dd6 single crystal superalloy blades, Adv. Mater. Res. 535–537 (2012) 1019–1022.
K. Goodwin, T.A. Lafford, J. Baruchel, H.J. Stone, N. D’Souza, Dendrite bending during [51] J. Zhang, L. Wang, D. Wang, G. Xie, Y.Z. Lu, J. Shen, L.H. Lou, Recent progress in re-
directional solidification, Superalloys 2012 (2012) 615–624. search and development of nickel-based single crystal superalloys, Acta Metall.
[39] M. Rappaz, J. Drezet, M. Gremaud, A new hot-tearing criterion, Metall. Mater. Trans. Sin. 55 (2019) 1077–1094.
A 30 (1999) 449–455. [52] M. Huo, L. Liu, W.C. Yang, Y.F. Li, S.S. Hu, H.J. Su, J. Zhang, H.Z. Fu, Formation of low-
[40] N. Wang, S. Mokadem, M. Rappaz, W. Kurz, Solidification cracking of superalloy angle grain boundaries under different solidification conditions in the rejoined plat-
single- and bi-crystals, Acta Mater. 52 (2004) 3173–3182. forms of Ni-based single crystal superalloys, J. Mater. Res. 34 (2019) 251–260.
[41] F. Wang, D.X. Ma, S. Bogner, A. Bührig-Polaczek, Influence of processing parameters [53] Y.S. Zhao, J. Zhang, Y.S. Luo, B. Zhang, G. Sha, L.F. Li, D.Z. Tang, Q. Feng, Improvement
on the solidification behavior of single-crystal CMSX-4 superalloy, Metall. Mater. of grain boundary tolerance by minor additions of Hf and B in a second generation
Trans. A 47 (2016) 3703–3712. single crystal superalloy, Acta Mater. 176 (2019) 109–122.
[54] Y.Y. Lian, D.C. Li, K. Zhang, Effects of the location of a cast in the furnace on flatness of
[42] W. Bogdanowicz, J. Krawczyk, R. Paszkowski, J. Sieniawski, Primary crystal orienta-
the solidification front in directional solidification, J. Cryst. Growth 451 (2016)
tion of the thin-walled area of single-crystalline turbine blade airfoils, Materials 12
33–41.
(2019) article no.: 2699.

13

You might also like