You are on page 1of 14

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/337408870

Microstructure-based modelling of rubbing in polycrystalline honeycomb


structures

Article  in  Continuum Mechanics and Thermodynamics · November 2019


DOI: 10.1007/s00161-019-00852-5

CITATIONS READS

0 58

4 authors, including:

Tim Fischer Oliver Munz


Technische Universität München Karlsruhe Institute of Technology
6 PUBLICATIONS   7 CITATIONS    8 PUBLICATIONS   15 CITATIONS   

SEE PROFILE SEE PROFILE

E. Werner
Technische Universität München
441 PUBLICATIONS   2,783 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Advanced micromechanical modelling of titanium alloys View project

All content following this page was uploaded by Tim Fischer on 25 November 2019.

The user has requested enhancement of the downloaded file.


Continuum Mech. Thermodyn.
https://doi.org/10.1007/s00161-019-00852-5

O R I G I NA L A RT I C L E

Tim Fischer · Sonun Ulan kyzy · Oliver Munz ·


Ewald Werner

Microstructure-based modelling of rubbing in polycrystalline


honeycomb structures

Received: 12 November 2019 / Accepted: 13 November 2019


© Springer-Verlag GmbH Germany, part of Springer Nature 2019

Abstract A microstructure-based modelling approach was used to study the deformation behaviour of poly-
crystalline honeycomb structures under rubbing loading. Rubbing originates from the sliding contact between
sealing surfaces in a gas turbine engine. As a stationary component of the sealing system, the honeycomb struc-
ture’s role is to prevent catastrophic failure of the rotating component. Therefore, normal forces and surface
deformations of the honeycomb structure need to be minimised to limit the heat input into the rotating com-
ponent. To achieve a detailed representation of the honeycomb material response, the constitutive behaviour
of the employed nickel-based superalloys Hastelloy X and Haynes 214 was modelled with a crystal plasticity
approach, utilising a finite element framework. Uniaxial tensile tests at relevant temperatures and strain rates
resembling the rubbing allowed the identification of crucial model parameters. The simulative studies based
on the unit cell of the honeycomb structure revealed that the normal forces and the surface deformations are
strongly affected by the microstructural features (size and orientation of the grains) and the applied deformation
rate. In addition, a significant amount of residual stresses could be found for the macroscopically unstressed
state after cooling and unloading.

Keywords Micromechanical modelling · Crystal plasticity · Rubbing contact · Honeycombs · Nickel-based


superalloys

1 Introduction

Honeycomb structures play a major role in the efficiency of gas turbines. They are frequently used as stationary
components in labyrinth seal systems (Fig. 1, left). In order to minimise leakage during engine operation, the
gap between the honeycomb structure and the rotating component (labyrinth seal fin) needs to be minimised.
Under extreme operating conditions, this can lead to a sliding contact between the components which is called
rubbing.

Communicated by Andreas Öchsner.


T. Fischer (B) · E. Werner
Institute of Material Science and Mechanics of Materials, Technische Universität München, Boltzmannstraße 15, 85748 Garching,
Germany
E-mail: fischer@wkm.mw.tum.de

S. Ulan kyzy
Metals and Alloys, Universität Bayreuth, Ludwig-Thoma-Straße 36b, 95447 Bayreuth, Germany

O. Munz
Institute of Thermal Turbomachinery, Karlsruher Institut für Technologie, Kaiserstraße 12, 76131 Karlsruhe, Germany
T. Fischer et al.

Fig. 1 Labyrinth seal system in a gas turbine from the macroscopic component scale to the microscopic scale of the polycrystalline
nickel-based superalloy honeycomb structure. During rubbing, the honeycombs are highly deformed and fractured. Adapted from
[27,35]

As a consequence of rubbing, both components may experience damage caused by the emerging thermo-
mechanical loads. While any type of damage on the rotor is unacceptable for safety reasons, local damage on the
honeycomb material is permitted to a certain degree. Unacceptable damage mechanisms on the rotor material
comprise, for example, excessive heat development, wear or hot cracking [28]. For the honeycomb material,
large breakouts are undesirable, since this increases the gap and, thus, the leakage between the components.
Moreover, this can affect the global structural stability. Severe localised deformations and fracture in the
presence of temperatures near the melting point are commonly found for the honeycomb material (Fig. 1,
bottom right). If severe surface deformations on the honeycomb structure occur, the heat input into the rotor
may be increased and hence threatens the integrity of the rotor material. By keeping these deformations at
a low level, the heat input is limited and the risk of failure of the rotor material is reduced. The extent of
surface deformation during rubbing defines, together with the material properties and the overall geometry,
the magnitude of the normal force acting on the honeycomb structure. Investigations conducted by [10,37]
demonstrated the great importance of the rubbing surface.
The polycrystalline nickel-based superalloys Hastelloy X and Haynes 214 are the most frequently used
honeycomb alloys [27,34,35]. While Hastelloy X offers a good high-temperature strength, Haynes 214 features
excellent oxidation resistance, particularly at temperatures above 955 ◦ C [12,13]. Both alloys differ in their
microstructure. While Hastelloy X exhibits a single-phase γ -microstructure with a face-centred cubic (fcc)
lattice structure, Haynes 214 shows a two-phase γ /γ  -microstructure with γ  precipitates [12,13,26].
As a result of the small dimensions of the honeycomb structure, which are in the order of the characteristic
length of the microstructure (Fig. 1, top right), a microstructure-based modelling approach for the rubbing
simulation is set up in the present work. The primary objective is to gain an understanding of the effect
of microstructural properties, such as size and orientation of grains, on the deformation behaviour of the
honeycomb structure. The current work is an extension of the studies of Fischer et al. [9,10]. Initially, the
studies were inspired by the work from Zhang et al. [42,43], who developed a microstructure sensitivity model
to analyse fretting contact.
Within this contribution, we present a rubbing simulation model for polycrystalline honeycomb structures.
The constitutive behaviour of the used materials is based on a crystal plasticity framework which is described
in detail in Sect. 2.1. Material model parameters were derived from macroscopic tensile tests conducted at
various temperatures and strain rates. The parameter identification was accompanied by numerical simula-
tion on a synthetically generated microstructure (Sect. 2.2). In Sect. 3, the microstructure-based modelling
Microstructure-based modelling of rubbing

approach for the rubbing simulation is presented. The simulation results are discussed in Sect. 4 along with
considerations of different effects on the rubbing behaviour. Finally, a conclusion of the work is given in
Sect. 5.

2 Constitutive model of the polycrystalline materials

With the overall goal to provide a microstructure-based model for the rubbing process, we use a crystal
plasticity approach to describe the temperature-dependent visco-plastic behaviour of the polycrystalline nickel-
based superalloys. The model was implemented into the finite element code Abaqus/Standard [5] with the
multi-physics software tool DAMASK [29–31], using the user–material interface (UMAT). Isotropic thermal
expansion of the material was considered in the model as presented by [19,20,29,41]. To calibrate the material
parameters, experimental data of high-temperature tensile tests were utilised.

2.1 Phenomenological crystal plasticity model

According to the classical concept [6], a multiplicative decomposition of the deformation gradient F = Fe Fp
has been used, where Fe is the elastic and Fp the plastic part. The rate Lp of the plastic velocity gradient is
introduced as the sum of the shear rates γ̇ i over all twelve slip systems of the underlying fcc lattice:


12
Lp = Ḟp Fp−1 = γ̇ i mi ⊗ ni , (1)
i=1

where mi and ni denote the slip direction and the slip plane normal of slip system i, respectively. The
temperature-dependent shear rate γ̇ i (T ) on a given slip system evolves as
 i   
 τ − χ i n(T )
i ∗ 
γ̇ (T ) = γ̇0 (T )  i  sgn τ i
− χ i
, (2)
τc (T ) 
using the common phenomenological power-law description [15,16]. The temperature-dependent reference
shear rate γ̇0∗ (T ) is an adjustable parameter comprising the temperature dependence as an Arrhenius term
[33]:
⎧  
⎨γ̇0 · exp − Q 0 T /Tm ≥ 0.8
γ̇0∗ (T ) =  R ·T  (3)
⎩γ̇0 · exp − Q 0
T /Tm < 0.8,
R ·0.8T m

where γ̇0 is the pre-exponential factor, Q 0 the apparent activation energy and
R the universal gas constant.
Moreover, Eq. (2) incorporates the resolved shear stress τ i = S : mi ⊗ ni with S being the second Piola–
Kirchhoff stress, the temperature-dependent critical resolved shear stress τci (T ), a back stress χ i and the
temperature-dependent stress exponent n (T ) = n 0 (T /Tm ) p . The term T /Tm denotes the homologous tem-
perature, where Tm stands for the melting temperature of the material. The parameters n 0 and p serve fitting
purposes. The evolution of the temperature-dependent critical resolved shear stress is calculated from:
a 

12
τ
j
(T )  
c  
τ̇ci (T ) = qi j h 0 1 − · γ̇ j (T ) , (4)
τs Π (T )
j=1

with the matrix qi j providing for the latent hardening behaviour between the individual slip systems i and
j, the slip hardening parameters h 0 and a, and τs which limits the shear stress to a saturation value. Grain
size dependence is captured in the model by means of the Hall–Petch relationship [23,40]. Therefore, static
changes in the grain size are considered by adapting the critical resolved shear stress and the saturation shear
stress following:
K c/s
τc/s = τc/s,0 + √ , (5)
d
T. Fischer et al.

Fig. 2 Calculated stress–strain curves at 1100 ◦ C compared to experimental results for different strain rates and mean grain
sizes (left). Representative volume element (RVE) constructed by employing periodic Voronoï tessellation. The crystallographic
orientation for each of the 150 grains was randomly assigned (right)

where τc/s,0 and K c/s are material parameters and d is the equivalent grain size diameter [24]. It should
be mentioned that Eq. (5) usually applies to many polycrystals at temperatures below the recrystallisation
temperature (T /Tm < 0.3−0.5) [17]. This should be borne in mind when applying it to elevated temperatures.
The temperature-dependent softening parameter Π (T ) with 0 ≤ Π ≤ 1 in Eq. (4) was introduced into the
model to adjust the critical shear stress and, thus, the shear rate to the prevailing temperature level.
To capture the deformation behaviour under reverse loading conditions, the model is enhanced by a kinematic
hardening model. According to the Armstrong–Frederick formulation [11,14], consisting of a hardening and
dynamic recovery term, the back stress χ̇ i on each slip system evolves as:
 
 
χ̇ i = h γ̇ i − r χ i γ̇ i  , (6)

where h and r are coefficients for hardening and dynamic recovery, respectively. As the present work predom-
inantly addressed monotonic loading conditions, kinematic hardening was not considered (χ i = 0). This also
reduced the computational effort significantly.

2.2 Identification of model parameters

This section presents the values for the previously introduced model parameters. These were primarily derived
from experimental data of macroscopic high-temperature tensile tests. In addition to test results at different
temperatures (20 ◦ C ≤ T ≤ 1250 ◦ C), data for different strain rates (1.25, 12.5 and 50 1/s) were available. An
excerpt of the results at 1100 ◦ C for the nickel-based superalloys is shown in Fig. 2, left. The identified model
parameters are found in Tables 1, 2 and 3.
By plotting the measured ultimate tensile strength as a function of the applied strain rate at different tem-
peratures, the temperature-dependent stress exponent n (T ) is obtained via estimation of the individual slopes.
As often observed in crystalline matter, the stress exponent decreases with the increasing temperature [17].
A small value of the stress exponent indicates an increasing time dependence of the deformation mechanism
and a more ductile material behaviour [2]. At temperatures around 1100 ◦ C, the values of our alloys are in
the range of 3 ≤ n ≤ 5 (cf. Table 3). The apparent activation energy Q is obtained by plotting the measured
ultimate tensile strength as a function of the imposed temperature at different strain rates and then estimating
the individual slopes [2]. The determined value of the activation energy is about 250 kJ/mol for both alloys
Microstructure-based modelling of rubbing

Table 1 Elastic material parameters for Hastelloy X and Haynes 214 at room temperature (elastic constants C11 , C12 and C44 ,
Young’s modulus E and Poisson’s ratio ν)

Alloy C11 , GPa C12 , GPa C44 , GPa E, GPa ν, –


Hastelloy X 302 134 91 200 0.31
Haynes 214 324 137 84 217 0.31

Table 2 Softening parameters Π (T ) with 0 ≤ Π ≤ 1 determined from macroscopic tensile tests for Hastelloy X and Haynes 214

Alloy 20 ◦ C 600 ◦ C 1100 ◦ C 1200 ◦ C 1230 ◦ C 1250 ◦ C


Hastelloy X 1.000 0.610 0.498 0.324 0.243 –
Haynes 214 1.000 0.816 0.253 0.132 – 0.130

Table 3 Crystal plasticity model parameters for the Hastelloy X and Haynes 214 alloy at room temperature

Alloy γ̇0 , s−1 Q 0 , kJ mol−1 n0 , – p, – a, – h 0 , MPa τc/s,0 , MPa K c/s , MPa mm

Hastelloy X 5 × 109 250 4.7 − 1.5 1.75 1000 121 / 221 7.27 / 13.26 [4]
Haynes 214 5 × 109 250 2.8 − 1.9 1.75 1000 220 / 416 7.27 / 13.75 [4]

(cf. Table 3). Besides providing further information about the prevailing deformation mechanism of the mate-
rial, the activation energy enables the introduction of an additional temperature dependency into the material
model [2]. The temperature-dependent softening Π (T ) parameter was obtained by normalising the flow stress
at elevated temperatures to the corresponding value at room temperature σy (T ) /σy (cf. Table 2). The material
constants required to describe the elastic material behaviour are listed in Table 1 (at room temperature). To
introduce the temperature dependence of the individual elastic constants (C11 , C12 and C44 ), the equation
proposed in [38] was employed.
To identify the remaining model parameters and for the material model validation, a series of numerical
simulations on a representative volume element (RVE) [3,7,36,39,41] were conducted. The RVE, illustrated
in Fig. 2, right, is fully periodic, formed of 150 grains and is generated by standard Voronoï tessellation [32].
The generation procedure of the RVE and the associated convergence study are described in detail by [10]. On
average, each grain was meshed with 867 tetrahedron quadratic elements (C3D10). The equivalent grain size
diameter di for each grain i in the RVE was assigned as:
 
di,0
di = d̄ · i = 1, 2, . . . , 150, (7)
d̄0
where d̄ is the mean grain size diameter taken from experimental data. The equivalent grain size diameter di,0
and the mean grain size diameter d̄0 comprise the corresponding initial dimensionless values. The measured
average grain size diameters for Hastelloy X were 29 µm and 59 µm. Grain size diameters of 28 µm and 71 µm
were determined for Haynes 214. Two different grain sizes for each alloy result from the fact that two different
sample geometries were used. It should be mentioned that the crystallographic orientation of each grain was
randomly assigned, according to experimental results. In Fig. 2, left, the calculated representative stress–strain
curves for different grain sizes and strain rates are added to the experimental results at 1100 ◦ C. This was done
after fitting of the remaining crystal plasticity model parameters (γ̇0 , a, h 0 and τc/s,0 ). The results show that
the model is quite capable of reproducing the experimental data. Due to the static nature of the implemented
grain size dependence, dynamic microstructural phenomena, e.g. dynamic recrystallisation and grain growth,
cannot be captured with the present model. This explains the differences between experiments and simulations
observed especially at high strain rates.

3 Simulation set-up for the rubbing model

The previously introduced modelling framework has been developed to accurately predict the constitutive
response of polycrystalline materials under thermo-mechanical loading. The capability of the framework was
used to simulate the rubbing process.
T. Fischer et al.

Fig. 3 Geometry and boundary conditions (BCs) of the hexagonal honeycomb structure for the rubbing simulation. The rubbing
surface is located in the centre of the periodic unit cell on the top face (z = h). Paths A and B are used for the evaluation of the
numerical results and also located on the rubbing surface

Table 4 Geometric dimensions of the investigated honeycomb structure

l, mm t, µm x, mm y, mm θ, ◦ h, mm
0.92 75.0 1.60 2.772 60 2.0
The height of the core part is given by h c = h/4

3.1 Geometry and boundary conditions

Figure 3 shows the geometry of the hexagonal honeycomb structure used for the simulation. The geometric
dimensions are given in Table 4. These dimensions are typical for honeycomb seal components and result in
a relative density of about 12.5 %. It is worth mentioning that the wall thickness is only 75 µm. Hence, the
macroscopic scale of the honeycomb structure is of the same magnitude as the characteristic length of the
microstructure of the material. In addition, the macroscopic or structural domain was restricted to the unit cell
of the honeycomb structure (highlighted by the dashed box) and periodic displacement boundary conditions
(BCs) to the outer faces of the cell [7,36] were applied. A similar procedure was presented in [21]. Thereby, an
infinite extension in the x- and y-direction was mimicked, reducing the computational effort [18]. Displacement
on the lower surface (z = 0) was fixed in the z-direction. The rubbing surface was assumed in the centre of
the honeycomb structure (indicated by the dark grey area) on the top face (z = h). Preceding investigations
showed that the loads at this site of the honeycomb structure were the highest due to the pronounced rubbing
surface area resulting from the double wall thickness (2 × t).
The rubbing loading parameters under typical service conditions are illustrated in Fig. 4, left, and given in
Table 5. Here, a differentiation between the heating/loading cycle (0 ≤ t/t0 ≤ 0.5) and the cooling/unloading
cycle (0.5 ≤ t/t0 ≤ 1) is required. The total rubbing time t0 is 8 s. During unloading, all loading parameters
were inactive, except for the initial temperature T0 acting on the bottom of the honeycomb structure. Moreover,
a thermo-mechanically coupled simulation strategy was employed. Within the preceding thermal analysis, the
initial temperature and a uniform heat flow rate q̇ on the rubbing surface were imposed. Thermo-physical
properties of the materials were taken from [12,13]. Subsequently, the resulting temperature distribution was
transferred to the mechanical analysis. Here, an absolute total displacement |u z | in the z-direction and a shear
stress τ0 , both resembling the movement of the rotating component, were prescribed on the rubbing surface.
Note that the total displacement was specified as a linear function of time during the loading cycle. This resulted
in an absolute constant displacement rate |u̇ z |. In contrast, the heat flow rate and the shear stress were applied
almost instantaneously. To impose the displacement and to evaluate the reaction forces, a reference point was
used. As indicated in Table 5, proportional relations were assumed following: τ0 = μ · p0 and q̇ = α̃ · τ0 · vs
[8], with the contact pressure p0 and the coefficient of friction μ determined from experiments on a test rig
[22]. Furthermore, α̃ is the heat partitioning factor (equal partitioning assumed) and vs is the sliding velocity.
Microstructure-based modelling of rubbing

Table 5 Rubbing parameters during the loading cycle (0 ≤ t/t0 ≤ 0.5)

p0 , MPa μ, − τ0 , MPa α̃, − vs , m/s q̇, W/mm2 T0 , ◦ C |u z |, µm


2.5 0.1 = μ · p0 0.5 110 = α̃ · τ0 · vs 350 40
During the unloading cycle (0.5 ≤ t/t0 ≤ 1), only T0 on the bottom of the investigated honeycomb structure remains active

Fig. 4 Thermal and mechanical loading imposed on the rubbing surface (left). Embedded cell approach consisting of a core part
with a fully resolved microstructure embedded in a macroscopic outer part (right)

3.2 Microstructure-based modelling approach

Since rubbing is a phenomenon spatially limited near the surface, the embedded cell approach can be used
to study the influence of microstructural properties on the rubbing behaviour of the nickel-based superalloys
[3,41]. The geometrical model consists of a core part with a discrete microstructure embedded in a non-discrete
or continuous outer part (Fig. 4, right). By separating it into a core and an outer part, an efficient calculation of
the micromechanical fields of inhomogeneous materials was possible. The core part was designed so that the
virtually average grain size diameter matches the experimental one of the real honeycomb structure [24,25].
The measured average grain size diameter for the Hastelloy X honeycomb structure was 71 µm (≈ 900 grains
with 134 elements per grain), and that for the Haynes 214 honeycomb structure was 101 µm (≈ 330 grains with
180 elements per grain). This means that only one to two grains make up the wall thickness of the honeycomb
structure [37]. To ensure that the transition of the local stress and strain fields between the core part and the outer
part was as smooth as possible, the material behaviour of the outer part (isotropic J2 plasticity) was chosen
to match the homogenised material behaviour of the core part. Besides, a congruent mesh at the interface of
the individual parts was ensured. The phenomenological crystal plasticity model described in Sect. 2.1 was
employed for the core part.

4 Results and discussion

In this section, the constitutive model was employed to simulate the rubbing process. The simulation results
of different effects on the rubbing behaviour of the nickel-based superalloys (Hastelloy X and Haynes 214)
are discussed. The considered effects were the size and orientation of the grains, the applied deformation rate
and the unloading cycle.
First, we compare the thermo-mechanical response of the alloys for the previously mentioned initial
configuration. In Fig. 5, left, a typical temperature history consisting of a heating and a cooling part is displayed.
The depicted temperature history corresponded to a point in the centre of the rubbing surface, at which location
the highest temperatures in the honeycomb structure were found. For a better comparison, the temperature
of the materials is expressed in terms of the homologous temperature. Melting of the alloys starts at around
1260 ◦ C for Hastelloy X and 1355 ◦ C for Haynes 214 [12,13]. As a result of the higher thermal conductivity
T. Fischer et al.

Fig. 5 Temperature history within the rubbing surface (left) and macroscopic load–displacement curve (right) of the nickel-based
superalloys Hastelloy X and Haynes 214

Fig. 6 Distribution of the temperature (left) and the von Mises equivalent stress (right) in the half-section of the core part for
Hastelloy X after heating/loading (t/t0 = 0.5)

and heat capacity, the alloy Haynes 214 revealed lower temperatures. After heating (t/t0 = 0.5), the maximum
homologous temperatures reached 0.86 (≈ 1050 ◦ C) for the alloy Hastelloy X and 0.78 (≈ 1000 ◦ C) for the
alloy Haynes 214. Figure 6, left, depicts the predicted inhomogeneous temperature distribution in the half-
section of the core part for Hastelloy X after heating. Within the rubbing surface, the temperatures decreased
from the centre to the outer sides in the x-direction (T ≈ 140 K). The macroscopic load–displacement curves
of the alloys are presented in Fig. 5, right. Three regions characterise the shape of the curves: (i) an initial
linear elastic progression of the load, (ii) a subsequent softening region after the onset of plastic deformation
in the honeycomb cell walls and (iii) a region with a slight increase in loads caused by further compressing
the honeycomb structure. Haynes 214 exhibited a highly pronounced softening region arising from the rapid
drop of its strength in the range of 600 ◦ C ≤ T ≤ 1000 ◦ C (cf. Table 2). Due to its high strength at low
temperatures (< 600 ◦ C), Haynes 214 was also exposed to the highest loads (≈ 98 N). Figure 6, right, presents
the distribution of the equivalent von Mises stress in the half-section of the core part for the Hastelloy X
alloy after loading. The highest stresses (≈ 927 MPa) were found at the side edges of the rubbing surface.
Here, the wall thickness of the honeycomb structure was the thinnest and the temperatures were below the
maximum, occurring in the centre of the rubbing surface. A strong localisation of the deformation prevailed.
Hence, at these sites the materials’ ability to plastically deform was restricted, which led to microcracks
that can propagate from the rubbing surface into the material. Due to the high temperatures and strain rates,
intergranular crack growth is expected [1,27]. This damage mechanism can additionally be promoted by the
accompanying oxidation at high temperatures.
In the following, the local deformation behaviour of the honeycomb structure on the rubbing surface is
taken into account. Figure 7 depicts the displacement in the y-direction after loading for the paths A and B
(as indicated in Fig. 3). The mean value and the standard deviation of five rubbing simulations with different
random assignments of the crystallographic orientation of the grains were evaluated. Both the alloys exhibited
maximum displacements near the middle of the path. This resulted mainly from the high temperatures located
at the centre of the surface and the associated lowered strength of the material. A difference in the standard
deviation between the two materials was not obvious. On the other hand, the mean value of the displacement
indicated a slight difference in the results. The maximum mean displacement of Haynes 214 (≈ 22 µm) was
Microstructure-based modelling of rubbing

Fig. 7 Local displacement in the y-direction after loading on the rubbing surface along the paths A and B for Hastelloy X (left)
and Haynes 214 (right)

Fig. 8 Macroscopic load–displacement curve after loading for Hastelloy X (left) and Haynes 214 (right) for two mean grain sizes

more pronounced compared to Hastelloy X (≈ 16 µm). This can be explained by the fact that Haynes 214
suffers from the pronounced loss of strength due to elevated temperatures.

4.1 Effect of grain size and grain orientation

Thermal effects during the manufacturing process of the honeycomb structure cause substantial grain growth
in the material [37]. This effect was assessed by investigating various types of microstructural realisations.
To study the effect of the grain size on the rubbing behaviour, we restrict the mean grain size to a minimum
of 50 µm and to a maximum of 174 µm. As in the previous section, for each microstructural realisation, five
simulations with different random orientation of the grains were performed. A notable impact of the grain size
on the loads is observed in Fig. 8. The regions with the highest forces are shown enlarged. As the grain size
increases, the forces of the two alloys decrease because of the grain size effect and the associated lower flow
stress of the material. For Hastelloy X, the forces drop by approximately 7 N (≈ 11.3 %) and for Haynes 214
by 10 N (≈ 9.3 %). Moreover, a larger grain size leads to greater deviation of the loads. This deviation was even
more apparent when considering the local displacement (Fig. 9). The increasing deviation can be attributed to
a weaker mutual interaction between the grains and their loss of strength. In conclusion, in case of an adverse
orientation of the grains, the largest rubbing surface was achieved by a coarse grain structure of Haynes 214
(≈ 30 µm). However, a significant influence of the grain size on the mean value of the local displacement was
not evident. This was valid for both alloys. The absolute mean displacement for Hastelloy X was about 12 µm
and for Haynes 214 close to 19 µm, neglecting the initial and final regions of the paths.

4.2 Effect of grain orientation

In order to study the grain orientation effect, the grain size was set to the initial configuration, i.e. 71 µm
(Hastelloy X) and 101 µm (Haynes 214). The experimentally observed predominant grain orientation in the
T. Fischer et al.

Fig. 9 Local displacement in the y-direction after loading on the rubbing surface along the paths A and B for Hastelloy X (left)
and Haynes 214 (right) for two mean grain sizes

Fig. 10 Macroscopic load–displacement curve after loading for Hastelloy X (left) and Haynes 214 (right) for two grain orientations

Fig. 11 Local displacement in the y-direction after loading on the rubbing surface along the paths A and B for Hastelloy X (left)
and Haynes 214 (right) for two grain orientations

real honeycomb structure was random (rnd). The non-random orientations were aligned parallel to the y-axis
(designated as
y). Similar to the grain size study presented in the previous section, an effect of the grain
orientation was clearly observable (Figs. 10, 11). For the [001]-direction, the macroscopic loads of both alloys
were least pronounced. The maximum drop in force for Hastelloy X was about 11 N (≈ 16.9 %) and for
Haynes 214 almost twice as much, with approximately 21 N (≈ 18.9 %). Moreover, only slight fluctuations in
the local displacement were noted for the [001]-direction. Over a large fraction of the path, the displacement
was nearly uniform. Again, the largest local deformations were found for the Haynes 214 alloy (≈ 20 µm).
Microstructure-based modelling of rubbing

Fig. 12 Macroscopic load–displacement curve after loading for Hastelloy X (left) and Haynes 214 (right) with variation of the
displacement rate

Fig. 13 Local displacement in the y-direction after loading on the rubbing surface along the paths A and B for Hastelloy X (left)
and Haynes 214 (right) with variation of the displacement rate

4.3 Effect of deformation rate

In this section, we point out that the rubbing behaviour not only depends on the microstructural properties,
but also on the external displacement rate. For a realistic study of the deformation mechanisms occurring
during rubbing, different deformation rates, similar to those found during the operation of the gas turbine,
were imposed. The results are shown in Figs. 12 and 13. As the deformation rate increased, the loads increased
significantly, whereby the softening region (region (ii) in Fig. 5, right) partially vanished (|u̇ z | > 0.01 mm/s)
and the local displacement on the rubbing surface decreased. The maximum forces (≈ 180 N) and the maximum
local displacements (≈ 30 µm) occurred for Haynes 214. To avoid large deformations during rubbing and
therefore prevent overheating of the rotor material, high deformation rates appeared to be more preferable. It
should be mentioned that additional heating due to the rise in contact pressure was not captured in the model.
However, higher deformation rates caused lower temperatures in the material due to the reduced exposure
time. This, in return, was taken into account by the model.

4.4 Effect of cooling and unloading

As a final result, the effect of cooling and unloading on the microstress state of the honeycomb structure is
discussed. The accurate prediction of residual stresses at the micro-scale is of interest to evaluate the tendency
for (hot-)crack formation, therefore a possible loss in the efficiency of the turbine due to leakage losses. These
cracks primarily occur during cooling at weakened grain boundaries in the presence of high tensile stresses
[28]. In Fig. 14, the maximum principal stresses from the macroscopically stress-free state on the rubbing
surface are shown along the paths A and B for both alloys. Both alloys revealed a notable increase in the local
stress level after the cooling and unloading cycle. This is indicated by the second-order polynomial fit, which
was introduced for enhanced visibility. (The illustrated dots refer to the calculated stresses at each integration
point.) An increase in the maximum principal stress is particularly pronounced at the initial and final regions
T. Fischer et al.

Fig. 14 Maximum principal stress on the rubbing surface along the paths A and B for Hastelloy X (left) and Haynes 214
(right) after heating (loading) and cooling (unloading). For enhanced visibility, the numerical results are fitted with second-order
polynomials

of the paths. Haynes 214 reaches the highest microscopic residual stresses (> 1000 MPa), slightly exceeding
the tensile strength of the material at the corresponding temperature (= T0 ). Therefore, cracks are likely to
form during the unloading cycle.

5 Conclusions

The rubbing behaviour of two polycrystalline nickel-based superalloy honeycomb structures was characterised
by employing a microstructure-based modelling approach. To adequately describe the material response of
the superalloys (Hastelloy X and Haynes 214), a crystal plasticity description was used. Model parameters
were identified from high-temperature macroscopic tensile tests. The material model was designed for a wide
range of temperatures and strain rates. It also takes into account the microstructural features of the material
(size and orientation of the grains). Subsequently, a series of numerical rubbing simulations were performed.
The studies showed that the normal forces and the surface deformation occurring during rubbing were much
greater for Haynes 214 than for Hastelloy X. However, the forces and deformations proved dependent on the
grain size, the orientation of the grains and the applied deformation rate. In addition, it could be demonstrated
that in the macroscopic unloaded state considerable residual stresses prevailed in the material. These stresses
can have a negative effect on the performance of the honeycomb structure. Future investigations aim to extend
the constitutive model to describe the localised damage process at the grain boundaries.

Acknowledgements This work is part of the research project WE 2351/14-1 funded by the Deutsche Forschungsgemeinschaft
(DFG). The authors are grateful to the Max Planck Institut für Eisenforschung (MPIE) in Düsseldorf for providing the flexible
and easy-to-use multi-physics simulation kit DAMASK.

References

1. Alabort, E., Barba, D., Sulzer, S., Lißner, M., Petrinic, N., Reed, R.C.: Grain boundary properties of a nickel-based superalloy:
characterisation and modelling. Acta Mater. 151, 377–394 (2018)
2. Alabort, E., Reed, R.C., Barba, D.: Combined modelling and miniaturised characterisation of high-temperature forging in a
nickel-based superalloy. Mater. Des. 160, 683–697 (2018)
3. Böhm, H.: A short introduction to basic aspects of continuum micromechanics. ILSB report, Vienna University of Technology,
vol. 206 (1998)
4. Cordero, Z.C., Knight, B.E., Schuh, C.A.: Six decades of the Hall–Petch effect—a survey of grain-size strengthening studies
on pure metals. Int. Mater. Rev. 62(8), 495–512 (2016)
5. Dassault Systèmes: Abaqus 6.13 analysis user’s guide (2013)
6. Dunne, F., Petrinic, N.: Introduction to Computational Plasticity. Oxford University Press, New York (2005)
7. Fillafer, A., Krempaszky, C., Werner, E.: On strain partitioning and micro-damage behavior of dual-phase steels. Mater. Sci.
Eng. A 614, 180–192 (2014)
8. Fischer, F., Werner, E., Knothe, K.: The surface temperature of a half-plane subjected to rolling/sliding contact with convec-
tion. J. Tribol. 122, 864–866 (2000)
9. Fischer, T., Werner, E., Ulan kyzy, S., Munz, O.: Modeling the rubbing contact in honeycomb seals. Contin. Mech. Thermodyn.
30(2), 381–395 (2018)
Microstructure-based modelling of rubbing

10. Fischer, T., Werner, E., Ulan kyzy, S., Munz, O.: Crystal plasticity modeling of polycrystalline Ni-base superalloy honeycombs
under combined thermo-mechanical loading. Contin. Mech. Thermodyn. 31(3), 703–713 (2019)
11. Frederick, C., Armstrong, P.: A mathematical representation of the multiaxial Bauschinger effect. G.E.G.B. report RD/B/N,
vol. 731 (1966)
12. Haynes International, I.: Hastelloy X alloy (UNS N06002). High-temperature alloys (1997)
13. Haynes International, I.: Haynes 214 alloy (UNS N07214). High-temperature alloys (2008)
14. Hennessey, C., Castelluccio, G.M., McDowell, D.L.: Sensitivity of polycrystal plasticity to slip system kinematic hardening
laws for Al 7075-T6. Mater. Sci. Eng. A 687, 241–248 (2017)
15. Hutchinson, J.W.: Bounds and self-consistent estimates for creep of polycrystalline materials. Proc. R. Soc. A Math. Phys.
Eng. Sci. 348(1652), 101–127 (1976)
16. Kalidindi, S.R.: Incorporation of deformation twinning in crystal plasticity models. J. Mech. Phys. Solids 46(2), 267–271,
273–290 (1998)
17. Kalpakjian, S., Schmid, S.R., Werner, E.: Werkstofftechnik. Pearson Studium, München (2011)
18. Kouznetsova, V.: Computational homogenization for the multi-scale analysis of multi-phase materials. Ph.D. thesis, Technical
University of Eindhoven (2002)
19. Meier, F.: Influence of the aluminum-microstructure on the damage behavior of integrated circuits. Ph.D. thesis, Technical
University of Munich (2017)
20. Meier, F., Schwarz, C., Werner, E.: Crystal-plasticity based thermo-mechanical modeling of Al-components in integrated
circuits. Comput. Mater. Sci. 94, 122–131 (2014)
21. Mohr, D.: Multi-scale finite-strain plasticity model for stable metallic honeycombs incorporating microstructural evolution.
Int. J. Plast. 22, 1899–1923 (2006)
22. Munz, O., Pychynski, T., Schwitzke, C., Bauer, H.J.: Continued experimental study on the friction contact between a labyrinth
seal fin and a honeycomb stator: slanted position. Aerospace 5(3), 82 (2018)
23. Petch, N.J.: The cleavage strength of polycrystals. J. Iron Steel Res. Int. 174, 25–28 (1953)
24. Quey, R., Dawson, P., Barbe, F.: Large-scale 3D random polycrystals for the finite element method: generation, meshing and
remeshing. Comput. Methods Appl. Mech. Eng. 200, 1729–1745 (2011)
25. Quey, R., Renversade, L.: Optimal polyhedral description of 3D polycrystals: method and application to statistical and
synchrotron X-ray diffraction data. Comput. Methods Appl. Mech. Eng. 330, 308–333 (2018)
26. Reed, R.C.: The Superalloys: Fundamentals and Applications. Cambridge University Press, Cambridge (2006)
27. Rossmann, A.: Die Sicherheit von Turbo-Flugtriebwerken: Reibverschleiß, Anstreifen und Spalthaltung, Labyrinthdichtun-
gen, Bürstendichtungen, Containment, Feuer und Explosionen. Turbo Consult, Karlsfeld (2000)
28. Rossmann, A.: Probleme der Maschinenelemente erkennen, verhüten und lösen—Band 2B: Versagensformen, typische
Schadensbilder, Mechanismen, Ursachen von Elementen durchströmter Systeme: Berührende und berührungsfreie Dichtun-
gen, Rohrleitungen, Schlauchleitungen. Turbo Consult, Karlsfeld (2010)
29. Roters, F., Diehl, M., Shanthraj, P., Eisenlohr, P., Reuber, C., Wong, S., Ma, D., Jia, N., Kok, P., Fujita, N., Ebrahimi, A.,
Hochrainer, T., Grilli, N., Janssens, K., Stricker, M., Weygand, D., Meier, F., Werner, E., Fabritius, H.O., Nikolov, S., Friak,
M., Raabe, D.: Damask—the Düsseldorf advanced material simulation kit for modelling multi-physics crystal plasticity,
damage and thermal phenomena from the single crystal up to the component scale. Comput. Mater. Sci. 158, 420–478 (2019)
30. Roters, F., Eisenlohr, P., Hantcherli, L., Tjahjanto, D., Bieler, T., Raabe, D.: Overview of constitutive laws, kinematics,
homogenization and multiscale methods in crystal plasticity finite-element modeling: therory, experiments, applications.
Acta Mater. 58(4), 1152–1211 (2010)
31. Roters, F., Eisenlohr, P., Kords, C., Tjahjanto, D., Diehl, M., Raabe, D.: Damask: the Düsseldorf advanced material simulation
kit for studying crystal plasticity using an FE based or a spectral numerical solver. Procedia IUTAM 3, 3–10 (2012)
32. Rycroft, C.H.: Voro++: a three-dimensional Voronoi cell library in C++. Chaos 19, 041111 (2009)
33. Sedláček, R.: Finite Elemente in der Werkstoffmechanik. Verlag Dr. Hut, München (2009)
34. Smarsly, W., Zheng, N., Buchheim, C., Nindel, C., Silvestro, C., Sporer, D., Tuffs, M., Schreiber, K., Bomba, C.L., Anderson,
O., Goehler, H., Simms, N., McColvin, G.: Advanced high temperature turbine seals materials and designs. Mater. Sci. Forum
492–493, 21–26 (2005)
35. Sporer, D., Shiembob, L.: Alloy selection for honeycomb gas path seal systems. Proc. AMSE Turbo Expos. 4, 763–774
(2004)
36. Taxer, T., Schwarz, C., Smarsly, W., Werner, E.: A finite element approach to study the influence of cast pores on the
mechanical properties of the Ni-base alloy MAR-M247. Mater. Sci. Eng. A 575, 144–151 (2013)
37. Ulan kyzy, S., Völkl, R., Munz, O., Fischer, T., Glatzel, U.: The effect of brazing on microstructure of honeycomb liner
material Hastelloy X. J. Mater. Eng. Perform. 28(4), 1909–1913 (2019)
38. Varshni, Y.P.: Temperature dependence of the elastic constants. Phys. Rev. B 2(10), 3952–3958 (1970)
39. von Kobylinski, J., Lawitzki, R., Hofmann, M., Krempaszky, C., Werner, E.: Micromechanical behaviour of Ni-based super-
alloys close to the yield point: a comparative study between neutron diffraction on different polycrystalline microstructures
and crystal plasticity finite element modelling. Contin. Mech. Thermodyn. 31(3), 691–702 (2019)
40. Werner, E., Siegmund, T., Weinhandl, H., Fischer, F.D.: Properties of random polycrystalline two-phase materials. Appl.
Mech. Rev. 47(1S), 231–240 (1994)
41. Werner, E., Wesenjak, R., Fillafer, A., Meier, F., Krempaszky, C.: Microstructure-based modelling of multiphase materials
and complex structures. Contin. Mech. Thermodyn. 28(5), 1325–1346 (2015)
42. Zhang, M., McDowell, D.L., Neu, R.W.: Microstructure sensitivity of fretting fatigue based on computational crystal plasticity.
Tribol. Int. 42, 1286–1296 (2009)
43. Zhang, M., Neu, R.W., McDowell, D.L.: Microstructure-sensitive modeling: application to fretting contacts. Int. J. Fatigue
31, 1397–1406 (2009)

Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional
affiliations.

View publication stats

You might also like