You are on page 1of 15

Engineering Failure Analysis 143 (2023) 106931

Contents lists available at ScienceDirect

Engineering Failure Analysis


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

Interfacial wear damage mechanism between Ti-alloy and Al-alloy


in interference-fit joint and influence of surface coatings:
Experimental and numerical study
Haoyuan suo a, Zhaohui Wei a, Bin Luo a, *, Linxuan Wang b, Kaifu zhang a,
Biao Liang a, Kelin Deng a, Hui Cheng a
a
Northwestern Polytechnical University, Xi’an, China
b
Shanghai Spaceflight Precision Machinery Institute, Shanghai, China

A R T I C L E I N F O A B S T R A C T

Keywords: Interfacial wear in Ti-alloy/Al-alloy interference-fit joint resulting from micro-displacement is


Ti/Al interference-fit joint considered one of the primary causes of structure failure. This paper described an experimental
Interfacial wear and numerical study on the interfacial wear damage mechanism between Ti-alloy and Al-alloy in
Surface coating
interference-fit joints and revealed the influence of surface coatings. The evolution of friction
Wear damage
coefficients (COFs), wear topography and element transfer were adopted as the characterization
Wear prediction model
methods. Ploughing effect (Ti-Al) and adhering effect (Al-Ti) are two main wear mechanisms for
uncoated samples. Hard-anodized film has more stable COFs and the film integrity is well
maintained. Wear layer has lower COFs but is easier to be peeled off due to better lubrication and
poor adhesion characteristics. A wear model considering variable COFs is established by
combining energy consumption and adaptive grid technique to simulate the wear scar and predict
the wear depth and volume, the predication results are in good agreement with experimental
results.

1. Introduction

As two irreplaceable structural materials in modern aviation industry, aluminum alloy (Al-alloy) and titanium alloy (Ti-alloy) are
widely used in various load-bearing structures in aircraft [1–6]. In actual assembly process, these two materials are usually connected
by bolt fastener, which is convenient to assemble and disassemble, inspect and repair thus can well meet the requirement of aircraft
assembly structures [4,6–10]. Usually, in order to effectively improve the fatigue life and sealing performance of the structure,
interference-fit would be introduced into the bolted joints [11,12]. During service, these interference-fit structures are often subjected
to vibration load, which induces the interfacial wear between the connected materials, resulting in serious interfacial damage and even
structural failure [13–15]. In fact, about 60%-80% of structural failures occur at the fastening joints [13,16], and the interfacial wear
damage is one of the main causes for joint failures thus should be paid more attention [17].
This interfacial wear in Al/Ti interference-fit joint is quiet complicated which takes place between two contacting surfaces. Due to
the use of the interference-fit technology, the interfacial wear generally occurs in a limited area of which these two contacting surfaces
subjected to the normal load undergo micro-displacement oscillatory motion (1–100 μm) [18].

* Corresponding author.
E-mail address: luobin@nwpu.edu.cn (B. Luo).

https://doi.org/10.1016/j.engfailanal.2022.106931
Received 2 September 2022; Received in revised form 4 November 2022; Accepted 7 November 2022
Available online 11 November 2022
1350-6307/© 2022 Elsevier Ltd. All rights reserved.
H. suo et al. Engineering Failure Analysis 143 (2023) 106931

As shown in Fig. 1, the interfacial wear areas in Al/Ti interference-fit joint is illustrated, which are mainly around the fastener. The
wear and degradation of these surfaces would lead to the loosening of the bolted structure and then induce the structure’s premature
failure and catastrophic destruction. Noteworthy, in recently years relevant scholars have conducted several meaningful studies on the
interfacial wear behaviors in assembly structures. On the one hand, these studies revealed the influence of the parameters choice and
surface coatings on the wear behaviors, and on the other hand the material removal and transfer mechanism during the wear process is
explored. Li et al. [19] studied the fretting wear behavior of bolted joint interfaces under tangential loading through experimental
methods, and revealed the effect of interfacial wear on the dynamic characteristics of structures. Zhang et al. [20] investigated the roles
of thread wear on self-loosening behavior of bolted joints subjected to transverse cyclic loading. They found that the self-loosening can
be caused by the fretting wear occurring at the thread without the rotation of the nut, and the preload plays an important role during
this process. Siegfried’s team developed an energy wear rate method to quantitatively evaluate the interface wear behavior in blade
disk assemblies [21], and provided a general method to study the interfacial wear damage in assembly structures [22,23]. Besides,
some scholars focused on the relationship between the wear and fatigue behaviors and revealed the influence of interfacial wear on the
fatigue life and failure modes [24,25]. It was concluded that the knowledge of the interfacial wear damage is indispensable for un­
derstanding the structure’s reliable service.
Apart from the interfacial wear in the assembly structures reviewed above, the wear characteristics of Al-alloy and Ti-alloy also
received extensive attention [26–29]. Kenneth et al. [30] summarized the tribo-characteristics of grade 2 commercially pure titanium
and the age-hard-enable Ti6Al4V and pointed out that both the alloys have poor abrasion resistance. The effect of groove surface
texture on the fretting wear of Ti6Al4V were investigated by Wang et al. [31], they found that the surface textures would influence the
formation and development of wear debris and then change the wear mechanism. Also, some researches have been conducted to reveal
the effect of element addition on the wear performance of Ti-alloy [32,33]. Similarly, the wear characteristics of Al-alloy were also
studied from two aspects: basic wear mechanism and surface medication[34]. Ghafaripoor et al. [35]and Akbari et al. [36] investi­
gated the morphology and action mechanism of PEO coating on 7075 Al alloy, which proved can improve the corrosion and tribo­
corrosion resistance of the materials. The role of material transfer in fretting wear behavior and mechanism of Alloy 690TT mated with
Type 304 stainless steel was investigated by Long’s article [37], they revealed the relationship between the type of material transfer
and fretting regime. These studies revealed the basic wear mechanism of Ti-alloy and Al-alloy, which is of great help to this research.
However, to the author’s best knowledge, the interfacial wear between Al-alloy and Ti-alloy around overlapping area in bolted
structures are rarely reported, especially in interference-fit bolted joints. In addition, surface treatment is generally introduced to
prevent electrochemical corrosion between these two dissimilar materials, which will make the wear process more complicated. In
order to make the structural designers and users understand the interfacial wear mechanism and damage evolution process in Al-alloy/
Ti-alloy interference-fit bolted joints, the present paper conducted experimental and numerical methods to reveal the interfacial wear
mechanism between Al-alloy and Ti-alloy with different surface coatings. First, different surface treatments were performed on the
materials, and then the surface hardness and morphology of all samples were recorded respectively. After that, wear experiments were
conducted to simulated the interfacial wear in Al-alloy/Ti-alloy interference-fit bolted joints. Subsequently, the friction coefficients
(COFs), surface wear topography and element transfer were characterized to reveal the basic interfacial wear mechanism and the
influence of surface coatings. Finally, the interfacial wear behavior between uncoated Ti-alloy and Al-alloy was simulated by finite
element methods.

2. Experiment campaign

2.1. Materials and specimens preparation

This study mainly investigates the interfacial wear behavior between the connected materials in Al-alloy/Ti-alloy interference-fit
bolted joints. The metal plates are Al-alloy 7075-T6 and Ti6Al4V, which are widely used in the aircraft due to their excellent service

Fig. 1. Interfacial wear in Al/Ti interference-fit joint.

2
H. suo et al. Engineering Failure Analysis 143 (2023) 106931

performance, and the fastener is the resist shear Ti-alloy (Ti6Al4V) bolt. The basic mechanical properties of Ti-alloy are provided by
Baoti Group Co., Ltd, China, and Al-alloy’s properties are referred from [11], as shown in Table 1.
According to the requirements of wear experimental equipment, the upper sample is processed into a φ9.3 × 8 mm cylinder block
with a radius of 4.65 mm hemispherical tip, and the lower sample is processed into a φ10 × 7.9 mm cylinder, as shown in Fig. 2. (b).
Before the wear experiment, the two materials were anodized and sprayed with wear layer, as shown in Fig. 2. (c-d). Anodizing
method is hard-anodized and the wear layer brand is HR-7201, which both show the great anti-corrosion and wear-resistance per­
formance during actual service. For easy recording, the hard-anodized and wear layer are abbreviated as HA and HR, respectively. For
example, the Ti(HR) represents the Ti-alloy sample with wear layer. In addition, the hardness of the uncoated and coated samples were
measured by Vickers digital hardness tester.

2.2. Wear test and analysis experiments

The interfacial wear behavior is mainly affected by the material properties, wear amplitude, interfacial normal pressure, loading
frequency and the lubrication state. As to bolted joints, these parameters are related to the connection parameters and relevant service
environment. The wear amplitude and frequency are selected to be 50 μm and 20 Hz through the analysis of the joint parameters and
service environment. The interfacial normal load is selected to be 100 N, realizing a mean contact pressure of about 300 MPa between
Al-alloy and Ti-alloy according to Hertzian contact theory, which is decided by the tightening torque and the bolt adopted of the joint.
For convenience, the wear pair is labeled as the form A-B, which A is the active (upper) sample and B is the driven (lower) sample. As
shown in Table 2, it’s the examples of the typical wear pairs.
After the wear tests, all the samples were ultrasonically cleaned for 30 mins in ethanol. The worn surface of all the samples are
inspected by a field emission scanning electron microscope (FE-SEM, Mira 3 Xmu Tescan, Czech). Then the distribution of elements is
characterized by an energy disperse spectroscopy (EDS, ZEISS SUPAR 55). Also, the wear volume and 3D surface morphology of
uncoated samples were obtained through 3D surface measurement instrument (NT1100, Veeco Wyko).

3. Energy wear model

In order to better characterize the interfacial wear behavior between Ti-alloy and Al-alloy, a FE wear model was established. The
COFs data was fitted according to the experimental results and inputted to the wear model. Energy consumption model and adaptive
grid technique were adopted to realize the wear process. Then the 3D wear scar was simulated and the wear depth and volume were
predicted.

3.1. Finite element model

For numerical simulation of the interfacial wear, the evolution of COFs is very important. The COFs is constantly changing during
the whole wear process. However, the previous researches mainly adopted constant COFs, which was bound to reduce the accuracy of
simulation. A wear model considering variable COFs was established in this research. Generally, according to the evolution of COF, the
interfacial wear process can be divided into two stage: the running-in stage and steady-state stage. As shown in Fig. 3(a), the first
10,000 wear cycles can be regarded as the running-in stage, which can be further divided into three parts: sharp change, slow growth
and relative stable. Polynomial function and sigmoid function were used to fit the actual COF curve by the actual wear stage, as shown
in Fig. 3(b). Then, the COF can be expressed by:


⎪ 5.959 × 10− 3 N − 3.857 × 10− 5 N 2 + 9.306 × 10− 8 N 3 − 1.131 × 10− 10 N 4 0 ≤ N < 380



⎪ 0.14261

⎨ 0.55551 + 380 ≤ N < 1780
1 + e(900− N)/100


⎪ 2
⎪ 6.762 × 10− 5 (N − 1780) − 1.084 × 10− 8 (N − 1780) − 1.353 × 10− 12 (N − 1780)
3



⎩ 4
− 16
+7.390 × 10 (N − 1780) + 0.6981061780 ≤ N

where N denotes the number of wear cycles.


As shown in Fig. 4, a ball-on-flat model of the interfacial wear is established in ABAQUS. The input model details and loading
history are illustrated in Fig. 4 (b) and Fig. 4(c) respectively. The surrounding of the lower sample is fully constrained and the top
surface of the hemi-sphere sample is restrained to prevent it from tilting. Normal load P is applied on the top surface of hemi-sphere
sample to bring two parts into full contact. Also, periodic tangential displacement S is applied to the hemi-sphere sample to generate
relative sliding between the two parts. In order to improve the accuracy and efficiency of the calculation, adaptive grid technique is

Table 1
Basic mechanical properties of Al-alloy and Ti-alloy.
E (GPa) ν Hardness (HB) Anodizing Wear layer

Al-alloy(7075-T6) 72 0.33 150 Hard-anodized HR-7201


Ti-alloy(Ti6Al4V) 111 0.30 280 Hard-anodized HR-7201

3
H. suo et al. Engineering Failure Analysis 143 (2023) 106931

Fig. 2. Wear test and the detail of the samples: (a) wear apparatus; (b) geometry of the samples; (c-d) coating types of the samples.

Table 2
Examples of typical wear pairs.
Wear pair Active sample (upper) Driven sample (lower)

Al-Ti Al-alloy Ti-alloy


Ti-Al Ti-alloy Al-alloy
Ti(HR)-Al Ti-alloy with wear layer Al-alloy
Al(HR)-Ti(HA) Al-alloy with wear layer Hard-anodized Ti-alloy

Fig. 3. Actual coefficient of friction curve obtained from wear experiment and the fitted curve.

introduced in this study. Hence, during the calculation process, the density of grid points at the wear contact area is not unchanged, but
will be constantly adjusted in the iterative process, and the grid will be refined to adapt to the current state. As a result, even when the
mesh size is large, the calculation accuracy can be guaranteed [38]. Referring to the results of previous studies [39,40], the initial mesh
size of the contact area is refined, the refined grid size is 10 μm and the eight-node linear hexahedron element (C3D8R) is used in the FE
model. Master-slave interaction contact is adopted in the ball-on-flat FE model. Hemi-sphere sample is defined as master and lower
sample is defined as slave. Tangential property is isotropic coulomb friction and normal property is hard contact. In addition, inter­
facial normal pressure and tangential displacement loading process are illustrated in Fig. 4(c).

3.2. Energy wear approach

In this study, energy consumption model is adopted to simulate the wear process. During the wear experiments, the interfacial
normal force between Ti-alloy and Al-alloy (Fint), tangential displacement (δ) and tangential force (T) are recorded separately. Previous
studies have suggested that the single wear cycle can be described by a wear loop, such as Fig. 5. Thus the work done by the tangential
friction can be equivalent to the area of the closed area formed by the loop, and this value can be calculated by:

4
H. suo et al. Engineering Failure Analysis 143 (2023) 106931

Fig. 4. Illustration of the wear FE model: (a) 3D ball-on-flat model; (b) model details; (c) loading history.

ET(j) ≈ 4δmax T = 4δmax Fint μj (1)

where ET(j) , δmax, Fint, T and μj represent the wear energy (J), friction amplitude (μm), interfacial normal force (N), tangential force (N)
and the coefficient of friction in jth friction cycle. Then, the total wear energy during the whole process can be expressed by:

N ∑N
ET = ET(j) = 4δmax Fint μ
j=1 j
(2)
j=1

The wear volume of Ti-alloy (VTi) and Al-alloy (VAl) are determined through 3D surface measurement instrument hence total wear
volume is: V = VAl + VTi. Finally, the wear rate (α) of the whole wear process is determined by:
V
α= (3)
ET
Therefore, for the FE model of interfacial wear, the wear depth at position x of jth wear cycle is:
hi (x) = αqi (x)Δsi (x) (4)

where hi(x), qi(x), ΔSi(x) represent the wear depth, shear force and relative slip distance at position x of jth wear cycle, respectively.

Fig. 5. Energy consumption analysis during wear process.

5
H. suo et al. Engineering Failure Analysis 143 (2023) 106931

Then the total wear depth and total wear volume at position x during whole wear process can be calculated by eq.(5) and eq.(6)
respectively:

N
h(x) = α qi (x)Δsi (x) (5)
i=1

V(x) = ΔA(x)h(x) (6)

where ΔA(x) is the micro-area at position x.

4. Result and discussion

4.1. Hardness and morphology of original specimens

In order to better reveal the wear characteristics of the materials, the hardness and surface microstructure of all the samples were
measured before the wear experiment. The vickers hardness is adopted in this research, which the 5 kg force is applied for 10 s during
the measure process. As shown in Fig. 6, for both Ti-alloy and Al-alloy, the hardness change greatly after surface treatment. Specially,
the hard-anodized film is harder than the original material, while the surface hardness decreases greatly after adding the wear layer.
This hardness change is determined by the performance of the two coatings. The hard-anodized film general composed of barrier layer
(inner layer) and oxide layer (outer layer). The inner layer is harder than the outer layer, which generally plays the role of protecting
the original materials. While some loose pores distributes at the outer layer which can absorb a variety of lubricants and then improve
the wear resistance and lubricating property of the coating [17]. It can also be observed from the microstructure, that a certain number
of loose pores distribute on the surface of the materials (Fig. 7(a-2) and Fig. 7(b-2)). While for the wear layer, it consists of lubrication
component and bonding components. The lubrication component is the highly dispersed colloidal molybdenum disulfide, and the
bonding components is the epoxy resin [13]. Therefore the hardness of wear layer is smaller (mainly decided by the bonding com­
ponents), and its surface microstructure appears more loosened (Fig. 7(a-3) and Fig. 7(b-3)).

4.2. Basic interfacial wear mechanism

As shown in Fig. 8, the evolution of friction coefficient (COFs) between uncoated Ti-alloy and Al-alloy during wear process are
presented. According to the variation of COFs between two contact surfaces, the entire wear process can be divided into two stages, the
running-in stage and steady-state stage [39]. The first 10,000 cycles is called the running-in stage, and the evolution details are
particularly shown in Fig. 8(b). For two wear pairs, the COFs both experiences three periods: sharp rise, violent fluctuation and relative
stability. In this typical running-in stage, obvious mechanical adjustment occurs between the wear interfaces. As for the asperity
contact of the samples under dry condition, this mechanical adjustment process is very complex, which closely related to the hardness
and the surface morphology of the samples [41]. For the Ti-Al wear pair, because Ti-alloy is harder than Al-alloy, it will produce a
plowing effect in the early wear stage. That is, the Al-alloy surface is continuously damaged by the Ti-alloy sample. As for the Al-Ti
wear pair, due to the hardness of Al-alloy is relatively small, the adhesion effect is more obvious in the early wear stage, which
also explains why the COFs of this wear pair fluctuate greatly than Ti-Al pair in this stage.
After 10,000 cycles (running-in stage), both wear pairs enter the steady-state stage, the evolution of COFs for these two wear pair is
similar, and the COFs fluctuate around 0.8 within a small range. In order to better explain this phenomenon, the element distribution
and microscopic morphology in the wear area are analyzed. For Ti-Al wear pair (Fig. 9 (a)), only a small amount of Ti element is
detected in the lower sample (less than 1%), indicating that only a very small amount of Ti is transferred from the Ti-alloy sample to the

Fig. 6. The Vickers hardness of the specimens.

6
H. suo et al. Engineering Failure Analysis 143 (2023) 106931

Fig. 7. Microstructure of Al-alloy and Ti-alloy before wear test.

Fig. 8. The evolution of COFs between uncoated Al-alloy and Ti-alloy.

lower sample during the wear process. As for the Al-Ti wear pair, a large amount of Al element is detected in the lower sample (more
than 50%), indicating that Al is continuously transferred to the surface of Ti samples during the wear process. This phenomenon can
also be observed from the microscopic morphology of the wear interface (Fig. 10). For both wear pairs, mainly the Al is transferred to
the surface of Ti-alloy samples, which is related to the plowing (Ti-Al) and the adhering (Al-Ti) effect between the contact surfaces. A
certain degree of oxygen element is detected on the lower specimen surface for both wear pairs, which is related to the oxidation of
metal materials (mainly aluminum oxidation). For Ti-Al wear pair, the oxygen element is more distributed at the edge of the wear area
(Fig. 9(a)), indicating that mainly the aluminum in the wear edge area is oxidized during the wear process, which can also be observed
from the micro-morphology (Fig. 10(a)). This is due to the material in the central area is peeled off and then piled up at the edge area,
and this part of the material is gradually oxidized after contact with the air. However, at the central area, a certain closed cavity forms,
thus the oxidation rate in this area is far lower than edge area. As to the Al-Ti wear pair, the oxygen element mainly distributes in the
central wear area (Fig. 9(b)), and roughly coincides with the distribution of the Al element, which indicates that aluminum transferred
to the Ti-alloy sample is oxidized during the wear process.

4.3. Influence of surface coating

Fig. 11 shows the COFs evolution curve between the coated sample and original material. It can be seen that the surface coatings
has a great influence on the wear characteristics between the Ti-alloy and Al-alloy. The wear layer and hard-anodized film on the
samples’ surface can both reduce the COFs to certain extent. Moreover, the COFs of the wear pair including wear layer sample is
smaller, and the evolution process can be obviously divided into three stages: running-in stage, stable growth stage and drastic
adjustment stage. As mentioned above, the early wear stage belongs to the typical running-in stage, that is, the mechanical adjustment

7
H. suo et al. Engineering Failure Analysis 143 (2023) 106931

Fig. 9. The distribution of elements of wear area on lower specimen: (a) Ti-Al; (b) Al-Ti.

Fig. 10. Microstructure of the specimens: (a) Ti-Al; (b) Al-Ti.

8
H. suo et al. Engineering Failure Analysis 143 (2023) 106931

stage. Compared with other wear pairs, the mechanical adjustment process which wear layer involved is more stable, and the
adjustment process is shorter (before 1000 cycles), mainly due to its good cushioning effect. After 1000 wear cycles, the COFs is in a
relatively stable growth process for a long time, which shows the excellent lubrication characteristics of the wear layer. In this paper,
the solid-film lubricant of grade HR7201 is selected, which adopts highly dispersed colloidal molybdenum disulfide as the lubrication
component. At the same time, the resin components are added as bonding components in the synthesis process of solid lubricants to
improve their adhesion. During the wear process, the lubrication properties of molybdenum disulfide are better played due to the great
bonding properties of the resin. It should be noted that the COFs of the wear pair containing wear layer sample has a large fluctuation
in the later wear period, and this fluctuation is related to the fact that the wear layer sample is active or driven part in the wear pair.
When the wear layer sample is the active one, the fluctuation of the COFs in the later period shows a downward trend while the driven
one shows an upward trend. This downward or upward trend is related to the compaction and peeling of the wear layer. As shown in
Fig. 12(b), when the wear layer sample is the driven one, the wear layer on the Ti-alloy is obviously peeled off by Al-alloy sample and
the original Ti-alloy exposes. However, due to the transfer of a certain amount of lubricating components on the Al-alloy grinding head,
the COFs after drastic adjustment remains at a relatively small level (about 0.65). While when the wear layer sample is the active one,
the wear layer is continuously compacted during the wear process, and transfers to the surface of the lower sample (Fig. 12(d)), which
makes the lubrication and protection function of the wear layer are better played, that is, molybdenum disulfide lubricant can play a
more lasting role at the wear interface. This also can be confirmed by the element analysis at the wear interface, when the wear layer
sample is the active part, a large amount of S element is detected in the lower sample, and its content in the wear center area even
exceeds 50%.
Compared with the other wear pairs (Fig. 11), the tribological characteristics of the wear pair containing hard-anodized sample are
quite special. The evolution of COFs is also related to whether the hard-anodized sample is the active one or the driven one. When the
hard-anodized sample is active part, the COFs between the wear interfaces shows dramatic fluctuations, and the fluctuation range

Fig. 11. The evolution of COFs between: (a) coated Ti-uncoated Al; (b) uncoated Al-coated Ti; (c) coated Al-uncoated Ti.

9
H. suo et al. Engineering Failure Analysis 143 (2023) 106931

Fig. 12. Microstructure of the specimens: (a) Al-Ti(HA); (b) Al-Ti(HR); (c) Al(HA)-Ti; (d) Al(HR)-Ti.

reaches about 0.2. While when the hard-anodized sample is the driven one, the COFs evolution process is similar to the trends between
uncoated samples, which can be divided into two stage: mechanical adjustment stage and steady-state stage. This COFs difference is
mainly related to the hardness and microstructure of the hard-anodized films. The hard-anodized film is generally composed by two
layers: the barrier layer (inner layer) and the oxide layer (outer layer), especially the outer layer (oxide layer) distributed with loose

Fig. 13. The distribution of elements of wear area on lower specimen: (a) Ti(HA)-Al; (b) Ti(HR)-Al; (c) Al(HA)-Ti; (d) Al(HR)-Ti.

10
H. suo et al. Engineering Failure Analysis 143 (2023) 106931

pores can absorb a variety of lubricants and then improves the wear resistance and lubricating property of the film [17]. During the
wear process, the debris caused by the interfacial damage will continue to fill the loose pores in the oxide layer. When the hard-
anodized sample is the active one (upper specimen), the filled debris will be continuously peeled off due to the effect of gravity
and friction force, and such filling and peeling action will occur alternately during the whole wear process, thus the COFs shows a large
fluctuation [18]. However, when the hard-anodized sample act is the driven one (lower specimen), the wear debris filled in the pores of
the oxide layer is hard to be peeled off, which makes it adhere to the wear surface for a long time, so the COFs is relatively stable during
the whole wear process. Also, the material transfer at the wear interface is concerned. As mentioned above, when the softer material
wear with the harder one, the softer material will transfer to the harder one by adhesion and ploughing effect [37]. This situation also
occurs between the hard-anodized sample and the uncoated one. As shown in Fig. 12(a) and Fig. 13(a), the integrity of the hard-
anodized film on the Ti-alloy is not compromised (it also because the film on Ti-alloy is more compact than that of Al-alloy), and
more Al elements are detected in the center wear area at the lower specimen. However, for the Al(HA)-Ti wear pair (Fig. 12(c) and
Fig. 13(c)), the hard-anodized film on the surface of the Al-alloy is destroyed totally, and a considerable amount of Al element is
detected on the surface of the Ti-alloy.
As shown in Fig. 14, the COFs evolution curves between the coated samples are illustrated. It should be focused that the wear pair
containing wear layer has lower and more stable COFs during the whole wear process, especially when the wear layer sample is the
active one, its lubrication effect is more prominent. For Al(HR)-Ti(HR) wear pair, it presents better wear resistance and has lower COFs
which keeps at a lower level (about 0.45) during the whole wear process. Compared with other wear pairs, the interface wear damage
for Al(HR)-Ti(HR) is relatively smaller (Fig. 15(d)), and there is no material transfer during the wear process (no obvious Al and Ti
elements are detected on the wear surface)(Fig. 16(c)). The distribution of S element on the surface is also uniform (Fig. 16(c)), which
shows that the wear layer has not been substantively damaged. However, for Al(HA)-Ti(HR) and Al(HR)-Ti(HA) wear pairs, due to the
hardness of hard-anodized film is much higher than wear layer, serious coating peeling occurs during the wear process (Fig. 15(b)&
(c)). For these two wear pairs, the lubricating components in wear layer are constantly transferred to hard-anodized surface and
adsorbed in the pore structure, thus the COFs of these two wear pairs remain stable during the whole wear process [37,42]. For the Al
(HA)-Ti(HR) wear pair, only a small amount of wear layer materials are transferred to the Al-alloy surface, and the cracking of the wear
layer on Ti-alloy surface is obvious(Fig. 15(b)). While for the Al(HR)-Ti(HA) wear pair, the wear layer on the Al-alloy surface is
severely damaged and adhered to the lower sample. For the hard-anodized sample, due to its high hardness and great adhesion (the
film directly forms on the body surface), it maintains great surface integrity but seriously damages the wear layer during wear process.
And for Al(HA)-Ti(HA) wear pair, because the hardness of the Al-alloy anodized film is lower than that of Ti-alloy, the Al-alloy surface
is destroyed and transferred to the Ti-alloy surface (Fig. 15(a)).
According to the above analysis, both the hard-anodized film and wear layer can effectively improve the surface wear resistance of
Ti-alloy and Al-alloy to a certain extent, reducing the COFs and the material surface damage during wear process. Specifically, because
of its excellent cushioning and lubricating property, wear layer has better tribological stability, but is easy to be damaged at the later
wear stage by the harder material surface due to its small hardness and poor surface adhesion. Compared with the wear layer, the hard-
anodized has higher hardness and better adhesion performance, it can maintain great surface integrity during wear process, thus has
better long-term service performance.

4.4. Analysis of numerical simulation results

In order to better characterize the wear mechanism between Ti-alloy and Al-alloy, the wear behavior of Ti-Al wear pair was
simulated, and the wear depth and volume were predicted respectively. As shown in Fig. 17, the experimental and numerical wear
results of Ti-Al wear pair are compared. It can be seen that the wear scar for the experiment and simulation are both elliptic shape,
which mainly depends on the hemispheric grinding head and the wear amplitude. The predicted maximum wear depth is 13.41 μm,
while the measured maximum wear depth is 13.06 μm. The predicted result is slightly larger than the measurement result. During the
wear process, part of the removed materials will break away from the wear surface, while the other part will adhere to and accumulate
around the wear area under the extrusion pressure. This part of the accumulated material will further affect the wear process, and slow
down the material removal rate to a certain extent. The measured maximum wear depth is the difference between the highest point and
the lowest point, which consider both the material removal and material accumulation. However, in simulation process the occurrence
of wear process is realized by mesh ablation [38], and the material accumulation is not considered, thus the predicted wear depth is
larger. The 2D profiles at the wear center area is shown in Fig. 17(c). The results also show that material accumulation has a great
influence on the formation of wear morphology, and the simulated results is larger than experimental. This has also been shown in
other studies [23,39].
On the basis of 3D wear profile, the 3D reconstruction was carried out to obtain the wear volume. It should be noted that the
measured wear volume consists of two parts: accumulation volume and removal volume, and the sum of these two parts is the total
wear volume. The predicated wear volume is 2.15 × 10-3 mm3, compared with the measured wear volume 1.97 × 10-3 mm3, the error is
about 8.37%, which proves the validity of the wear prediction model. Similar to the wear depth, the source of the error is mainly the
influence of the material accumulation. In other studies using adaptive grid technology and improved Archard method to predict wear
volume and depth (ball-on-flat model), the error between experimental and simulation results is also about 10% when the maximum
wear depth is less than 15 μm [38,40].

11
H. suo et al. Engineering Failure Analysis 143 (2023) 106931

Fig. 14. The evolution of COFs between coated Al and coated Ti.

Fig. 15. Microstructure of the specimens: (a) Al(HA)-Ti(HA); (b) Al(HA)-Ti(HR); (c) Al(HR)-Ti(HA); (d) Al(HR)-Ti(HR).

5. Conclusion

The interfacial wear damage mechanism between Ti-alloy and Al-alloy in interference-fit joint and the influence of surface coatings
(hard-anodized film and wear layer) were researched in this paper. The evolution of COFs, surface morphology, and material transfer
situation were adopted as the characterization methods. Especially, a wear model (Ti-Al pair) considering variable COFs is established
to simulate the wear scar and predict the wear depth and wear volume. The following conclusions can be drown:
(1) The wear process between Al-alloy and Ti-alloy can be divided into running-in stage and steady-state stage. The wear behavior
of Ti-Al wear pair is different from that of Al-Ti wear pair. For Ti-Al wear pair, the ploughing effect is more obvious, while the adhering
effect dominates the whole wear process for Al-Ti wear pair.
(2) Both hard-anodized film and wear layer can improve the wear resistance and protect the original material. Hard-anodized film
has more stable COFs and the film integrity is well maintained during wear process. Wear layer has lower COFs but is easier to be
peeled off due to its better lubrication and poor adhesion characteristics.
(3) During the wear process, the softer material will be transferred to harder sample’s surface, and when the harder sample is the
active one, this material transfer is more obvious. When the coated sample is the active one, its lubrication and wear resistance
characteristics are more prominent, which is mainly due to the adjustment effect of the coatings can be better worked out.
(4) The 3D wear scar was simulated, and the maximum wear depth and total wear volume were predicted. The predication results
are in good agreement with the experimental results, and the errors are less than 10%. The influence of material accumulation during
wear experiment is the main reason for this error.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

12
H. suo et al. Engineering Failure Analysis 143 (2023) 106931

Fig. 16. The distribution of elements of wear area on lower specimen: (a) Al(HA)-Ti(HA); (b) Al(HA)-Ti(HR); (c) Al(HR)-Ti(HR).

Fig. 17. Comparison between experimental and numerical results of Ti-Al wear pair: (a) simulated and (b) real 3D surface morphology of lower
sample; (c) 2D wear profiles at typical wear position; (d) wear volume and depth.

13
H. suo et al. Engineering Failure Analysis 143 (2023) 106931

Data availability

Data will be made available on request.

Acknowledgements

This work was supported by the National Natural Science Foundation of China [No. 52035011] and the Natural Science Basic
Research Program of Shaanxi (No. 2022KJXX-74).

References

[1] Y. Cao, Z. Cao, Y. Zuo, L. Huo, J. Qiu, D. Zuo, Numerical and experimental investigation of fitting tolerance effects on damage and failure of CFRP/Ti double-lap
single-bolt joints, Aerosp. Sci. Technol. 78 (2018) 461–470, https://doi.org/10.1016/j.ast.2018.04.042.
[2] F. Bayata, C. Yildiz, The effects of design parameters on mechanical failure of Ti-6Al-4V implants using finite element analysis, Eng. Fail. Anal. 110 (2020),
104445, https://doi.org/10.1016/j.engfailanal.2020.104445.
[3] R. Masoudi Nejad, F. Berto, M. Tohidi, Fatigue performance prediction of Al-alloy 2024 plates in riveted joint structure, Eng. Fail. Anal. 126 (2021) 105439.
10.1016/j.engfailanal.2021.105439.
[4] J.P. Davim, Mechanical and Industrial Engineering: Historical Aspects and Future Directions, Cham Switzerland: Springer, 2022.
[5] J.P. Davim, Wear of Advanced Materials, Wiley-ISTE, London, 2012.
[6] H. Suo, H. Cheng, B. Liang, K. Deng, B. Luo, K. Zhang, H. Chen, The mechanical degradation mechanism of CFRP/Al double-lap bolted joints (with and without
corrosion protections) after seawater ageing, Compos. Struct. 276 (2021), 114561, https://doi.org/10.1016/j.compstruct.2021.114561.
[7] A.K. Abdul Jawwad, N. ALShabatat, M. Mahdi, The effects of joint design, bolting procedure and load eccentricity on fatigue failure characteristics of high-
strength steel bolts, Eng. Fail. Anal. 122 (2021) 105279. 10.1016/j.engfailanal.2021.105279.
[8] M. Samaei, M. Zehsaz, T.N. Chakherlou, Experimental and numerical study of fatigue crack growth of aluminum alloy 2024–T3 single lap simple bolted and
hybrid (adhesive/bolted) joints, Eng. Fail. Anal. 59 (2016) 253–268, https://doi.org/10.1016/j.engfailanal.2015.10.013.
[9] R. Starikov, J. Schön, Quasi-static behaviour of composite joints with protruding-head bolts, Compos. Struct. 51 (2001) 411–425, https://doi.org/10.1016/
S0263-8223(00)00157-4.
[10] G. Restivo, G. Marannano, G.A. Isaicu, Three-dimensional strain analysis of single-lap bolted joints in thick composites using fibre-optic gauges and the finite-
element method, J. Strain Anal. Eng. Des. 45 (2010) 523–534, https://doi.org/10.1243/03093247JSA599.
[11] H. Li, K. Zhang, H. Cheng, H. Suo, Y. Cheng, J. Hu, Multi-stage mechanical behavior and failure mechanism analysis of CFRP/Al single-lap bolted joints with
different seawater ageing conditions, Compos. Struct. 208 (2019) 634–645, https://doi.org/10.1016/j.compstruct.2018.10.044.
[12] T.N. Chakherlou, M. Mirzajanzadeh, B. Abazadeh, K. Saeedi, An investigation about interference fit effect on improving fatigue life of a holed single plate in
joints, Eur. J. Mech. - A/Solids. 29 (2010) 675–682, https://doi.org/10.1016/j.euromechsol.2009.12.009.
[13] J. Hu, K. Zhang, Q. Yang, H. Cheng, S. Liu, Y. Yang, Fretting behaviors of interface between CFRP and coated titanium alloy in composite interference-fit joints
under service condition, Mater. Des. 134 (2017) 91–102, https://doi.org/10.1016/j.matdes.2017.08.018.
[14] J.P. Davim, Tribology for Engineers: A Practical Guide, Woodhead Publishing, Cambridge, 2011.
[15] J.P. Davim, Progress in Green Tribology: Green and Conventional Techniques, De Gruyter, Berlin, 2017.
[16] Z. Cao, M. Cardew-Hall, Interference-fit riveting technique in fiber composite laminates, Aerosp. Sci. Technol. 10 (2006) 327–330, https://doi.org/10.1016/j.
ast.2005.11.003.
[17] H. Suo, Z. Wei, K. Zhang, K. Deng, H. Cheng, B. Luo, H. Li, L. Wang, B. Liang, Interfacial wear damage of CFRP/Ti-alloy single-lap bolted joint after long-term
seawater aging, Eng. Fail. Anal. 139 (2022), 106464, https://doi.org/10.1016/j.engfailanal.2022.106464.
[18] S.R. Soria, A. Tolley, A. Yawny, A study of debris and wear damage resulting from fretting of Incoloy 800 steam generator tubes against AISI Type 304 stainless
steel, Wear. 368–369 (2016) 219–229, https://doi.org/10.1016/j.wear.2016.09.022.
[19] D. Li, D. Botto, C. Xu, M. Gola, Fretting wear of bolted joint interfaces, Wear 458–459 (2020), 203411, https://doi.org/10.1016/j.wear.2020.203411.
[20] M. Zhang, L. Lu, W. Wang, D. Zeng, The roles of thread wear on self-loosening behavior of bolted joints under transverse cyclic loading, Wear 394–395 (2018)
30–39, https://doi.org/10.1016/j.wear.2017.10.006.
[21] S. Fouvry, P. Kapsa, L. Vincent, Quantification of fretting damage, Wear 200 (1996) 186–205, https://doi.org/10.1016/S0043-1648(96)07306-1.
[22] S. Fouvry, P. Arnaud, A. Mignot, P. Neubauer, Contact size, frequency and cyclic normal force effects on Ti–6Al–4V fretting wear processes: an approach
combining friction power and contact oxygenation, Tribol. Int. 113 (2017) 460–473, https://doi.org/10.1016/j.triboint.2016.12.049.
[23] S. Baydoun, S. Fouvry, S. Descartes, P. Arnaud, Fretting wear rate evolution of a flat-on-flat low alloyed steel contact: a weighted friction energy formulation,
Wear 426–427 (2019) 676–693, https://doi.org/10.1016/j.wear.2018.12.022.
[24] L. Zhao, X. He, B. Xing, X. Zhang, Q. Cheng, F. Gu, A. Ball, Fretting behavior of self-piercing riveted joints in titanium sheet materials, J. Mater. Process. Technol.
249 (2017) 246–254, https://doi.org/10.1016/j.jmatprotec.2017.06.016.
[25] R. Sun, Z. Che, Z. Cao, H. Zhang, S. Zou, J. Wu, W. Guo, Effect of laser shock peening on high cycle fatigue failure of bolt connected AA2024-T351 hole
structures, Eng. Fail. Anal. 141 (2022), 106625, https://doi.org/10.1016/j.engfailanal.2022.106625.
[26] M.X. Shen, M.H. Zhu, Z.B. Cai, X.Y. Xie, K.C. Zuo, Dual-rotary fretting wear behavior of 7075 aluminum alloy, Tribol. Int. 48 (2012) 162–171, https://doi.org/
10.1016/j.triboint.2011.11.024.
[27] J.L. Mo, M.H. Zhu, J.F. Zheng, J. Luo, Z.R. Zhou, Study on rotational fretting wear of 7075 aluminum alloy, Tribol. Int. 43 (2010) 912–917, https://doi.org/
10.1016/j.triboint.2009.12.032.
[28] J. Peng, X. Jin, Z. Xu, J. Zhang, Z. Cai, Z. Luo, M. Zhu, Study on the damage evolution of torsional fretting fatigue in a 7075 aluminum alloy, Wear. 402–403
(2018) 160–168, https://doi.org/10.1016/j.wear.2018.02.008.
[29] H. Attia, M. Meshreki, A. Korashy, V. Thomson, V. Chung, Fretting wear characteristics of cold gas-dynamic sprayed aluminum alloys, Tribol. Int. 44 (2011)
1407–1416, https://doi.org/10.1016/j.triboint.2011.05.006.
[30] K.G. Budinski, Tribological properties of titanium alloys, Wear. 151 (1991) 203–217, https://doi.org/10.1016/0043-1648(91)90249-T.
[31] J. Wang, W. Xue, S. Gao, S. Li, D. Duan, Effect of groove surface texture on the fretting wear of Ti–6Al–4V alloy, Wear 486–487 (2021), 204079, https://doi.org/
10.1016/j.wear.2021.204079.
[32] B.N. Mordyuk, V.V. Silberschmidt, G.I. Prokopenko, Y.V. Nesterenko, M.O. Iefimov, Ti particle-reinforced surface layers in Al: effect of particle size on
microstructure, hardness and wear, Mater. Charact. 61 (2010) 1126–1134, https://doi.org/10.1016/j.matchar.2010.07.007.
[33] P.K. Verma, S. Warghane, U. Nichul, P. Kumar, A. Dhole, V. Hiwarkar, Effect of boron addition on microstructure, hardness and wear performance of Ti-6Al-4 V
alloy manufactured by laser powder bed fusion additive manufacturing, Mater. Charact. 172 (2021), 110848, https://doi.org/10.1016/j.matchar.2020.110848.
[34] N. Arun Prakash, R. Gnanamoorthy, M. Kamaraj, Fretting wear behavior of fine grain structured aluminium alloy formed by oil jet peening process under dry
sliding condition, Wear. 294–295 (2012) 427–437, https://doi.org/10.1016/j.wear.2012.07.026.
[35] M. Ghafaripoor, K. Raeissi, M. Santamaria, A. Hakimizad, The corrosion and tribocorrosion resistance of PEO composite coatings containing α-Al2O3 particles
on 7075 Al alloy, Surf. Coatings Technol. 349 (2018) 470–479, https://doi.org/10.1016/j.surfcoat.2018.06.027.
[36] E. Akbari, F. Di Franco, P. Ceraolo, K. Raeissi, M. Santamaria, A. Hakimizad, Electrochemically-induced TiO2 incorporation for enhancing corrosion and
tribocorrosion resistance of PEO coating on 7075 Al alloy, Corros. Sci. 143 (2018) 314–328, https://doi.org/10.1016/j.corsci.2018.08.037.

14
H. suo et al. Engineering Failure Analysis 143 (2023) 106931

[37] L. Xin, Y. Lu, T. Shoji, The role of material transfer in fretting wear behavior and mechanism of Alloy 690TT mated with Type 304 stainless steel, Mater. Charact.
130 (2017) 250–259, https://doi.org/10.1016/j.matchar.2017.06.020.
[38] A. Bastola, D. Stewart, D. Dini, Three-dimensional finite element simulation and experimental validation of sliding wear, Wear. 504–505 (2022), 204402,
https://doi.org/10.1016/j.wear.2022.204402.
[39] L. Li, L. Kang, S. Ma, Z. Li, X. Ruan, A. Cai, Finite element analysis of fretting wear considering variable coefficient of friction, Proc. Inst. Mech. Eng. Part J J.
Eng. Tribol. 233 (2019) 758–768, https://doi.org/10.1177/1350650118800615.
[40] K. Fallahnezhad, S. Liu, O. Brinji, M. Marker, P.A. Meehan, Monitoring and modelling of false brinelling for railway bearings, Wear. 424–425 (2019) 151–164,
https://doi.org/10.1016/j.wear.2019.02.004.
[41] J. Fei, H.-J. Li, J.-F. Huang, Y.-W. Fu, Study on the friction and wear performance of carbon fabric/phenolic composites under oil lubricated conditions, Tribol.
Int. 56 (2012) 30–37, https://doi.org/10.1016/j.triboint.2012.06.022.
[42] I. Rustamov, F. Guo, Z. Wang, Experimental investigations into fretting wear and damage mechanisms of Inconel X-750 alloy, J. Mech. Sci. Technol. 33 (2019)
4701–4713, https://doi.org/10.1007/s12206-019-0818-8.

15

You might also like