Lychagin 2020

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Tribology International 147 (2020) 106284

Contents lists available at ScienceDirect

Tribology International
journal homepage: http://www.elsevier.com/locate/triboint

Deformation of hadfield steel single crystals by dry sliding friction with the
normal load/friction force orientations ½110�/½110� and ½110�/[001]
D.V. Lychagin a, A.V. Filippov b, *, O.S. Novitskaya b, Y.I. Chumlyakov a, E.A. Kolubaev b, L.
L. Lychagina c, d
a
National Research Tomsk State University, 36, Lenin Ave., Tomsk, 634050, Russia
b
Institute of Strength Physics and Material Science of SB RAS, 2/4 Akademichesky Ave., Tomsk, 634055, Russia
c
Tomsk State University of Architecture and Building, 2, Solyanaya Sq., Tomsk, 634003, Russia
d
Tomsk State University of Control Systems and Radioelectronics, 40, Lenin Ave., Tomsk, 634050, Russia

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

Keywords: This paper studies the friction and wear of Hadfield steel single crystals with the normal load/friction force
Hadfield steel orientations ½110�/½110� and ½110�/[001]. Due to the different friction force orientation, deformation occurs by
Sliding friction
twinning in the first case and by slip in the second case. The shear stresses were estimated and correlated with the
Wear
observed slip band systems. Changes in the orientation of the worn surface were examined by EBSD analysis.
Deformation
TEM studies were performed to investigate the dislocation structure evolution near the worn surface.

1. Introduction It was found that the hardening of Hadfield steel involves a whole
range of mechanisms, but the role of each mechanism depends on the
Iron-manganese steels are very attractive for industrial applications specific deformation conditions [3]. The types of dislocation structure
due to their mechanical characteristics. Hadfield steel, known for its determine the stress-strain curve stages and the work hardening coeffi­
excellent wear resistance under high loading conditions, is widely used cient [21]. Depending on the dominant deformation mechanism, the
in mining, railway industry, in the manufacture of caterpillar tracks and role of substructural hardening [22] and hardening due to internal stress
excavator buckets [1,2]. Owing to its high strength characteristics, fields [23,24] is different. In Hadfield steel, a change in the type of
Hadfield steel is a very popular material for tribological applications. dislocation structure leads to a change in the work hardening coefficient
The boundary between contacting bodies is the place where various and determines the stress-strain curve shape for single crystals with
processes occur simultaneously, such as deformation, heating, wear, and different orientations [5]. The hardening characteristics of Hadfield
so on. All of them influence the deformation structure evolution in the steel largely depend on the crystallographic orientation for single crys­
contact zone. tals and texture for polycrystal due to the anisotropy of slip systems for
Analysis of a large body of literature data on austenitic steels with slip and twinning, as well as on the stress sign in the slip plane during
high manganese content shows that slip and twinning induced plasticity twinning (tension or compression). Non-textured polycrystals provide
are the main deformation mechanisms in this type of steels [3–10]. The averaged information because of the statistical scatter of grain orienta­
dominant deformation mechanism and possible types of interaction tions in a polycrystal and heterogeneous stress fields in the deformed
between structural defects can be determined by such parameters as material at different scale levels.
crystallographic orientation [4–7], stacking fault energy [12], strain According to Ref. [25], while interacting with dislocations, carbon
rate [13–17], and temperature [3,11,18,19]. The high degree of work atoms in C–Mn clusters produce a high dislocation density in tangles,
hardening of Hadfield steel is associated with the formation of defor­ giving rise to short-range stress fields. The authors of Ref. [26] proposed
mation micro- and nanotwins [12]. Deformation twins can also be a model to describe the work hardening of Hadfield steel. The model is
formed due to loading rate increase [13]. It was clearly demonstrated in based on the assessment of the interaction between carbon atoms in
impact load tests that a higher strain rate leads to a higher activity of interstitial sites and nearest neighboring metal atoms. The stress for the
twins [14,20]. motion of subsequent dislocations is lower than that required to trigger

* Corresponding author.
E-mail address: avf@ispms.ru (A.V. Filippov).

https://doi.org/10.1016/j.triboint.2020.106284
Received 7 December 2019; Received in revised form 31 January 2020; Accepted 27 February 2020
Available online 2 March 2020
0301-679X/© 2020 Elsevier Ltd. All rights reserved.
D.V. Lychagin et al. Tribology International 147 (2020) 106284

dislocation climbing or glide in less stressed slip planes. In Hadfield steel


of conventional composition, the order will be restored during defor­
mation, suppressing planar glide. In Ref. [27,28], an elasto-viscoplastic
polycrystalline model was proposed which includes slip and twinning in
the form of deformation modes. The model was applied to describe the
behavior of Fe–Mn–C austenitic steel with low stacking fault energy. It
was suggested that work hardening results from the competition be­
tween hardening and softening associated with twinning.
Earlier, the strain hardening and wear mechanisms under various
friction conditions were studied for polycrystalline Hadfield steel
[29–38] and other high manganese steels [39], including those alloyed
with N, Cr [40], Al [41,42] and reinforced by TiC [43] and V2C particles
[44]. These studies were carried out in order to clarify the influence of
Fig. 1. Test configuration and position of ½110�/[001] (a) and ½110�/½110�
friction conditions and its intensity on the main wear mechanisms. In
Hadfield steel single crystals (b) in dry sliding friction.
most cases, the main wear mechanism of materials not reinforced by
particles at low loads is the oxidative wear; as the load increases, other
determining the average wear rate for the entire test period. The test
types of wear become more pronounced, such as microcutting, lamina­
termination criterion was met when the wear volume reached 3 mm3.
tion, and plastic deformation. The addition of particles (their incorpo­
This criterion was selected on the basis of a large number of experi­
ration into the austenitic matrix) changes the wear resistance and wear
mental studies, because upon reaching this value the wear rate of the
mechanisms, but a clear dependence of the wear (mechanisms and wear
samples stabilizes regardless of the considered crystallographic
rate) on the loading (deformation) conditions of the material remains.
orientation.
In Ref. [29], the authors observed a relationship between the crys­
The orientation of microregions was determined on a TESCAN VEGA
tallographic orientation of grains in polycrystalline Hadfield steel and
II LMU scanning electron microscope equipped with an EBSD detector.
the deformation behavior of the material when scratched by a single
The minimum size of misoriented regions was 1 μm. The deformation
fixed abrasive. The degree of damage to the surface of grains depended
structure of surface and subsurface layers was examined in cross section
on their orientation. The wear mechanisms and coefficient of friction
after tribological tests by transmission electron microscopy (JEOL JEM-
also varied depending on the grain orientation. The main wear mecha­
2100F).
nism of grains with the (001) surface was microploughing, and for grains
with the (111) surface it was microcutting. The results of Ref. [29]
clearly show that by comprehensively examining the crystallographic 3. Results
orientation effect it is possible to simulate a texture in a polycrystalline
material that would be most suitable for the required friction conditions. 3.1. Analysis of the friction process: wear and coefficient of friction
Hadfield steel single crystals under sliding friction are insufficiently
investigated. Earlier, we studied the deformation relief evolution during For the ½110�/½110� sample, the values of the coefficient of friction
sliding friction of Hadfield steel single crystals with the normal pressure (CoF) vary in the range 0.41–0.53 within the first 48 h of friction, after
axis ½118� and the friction axis ½441� [45]. The main deformation which they sharply increase to 0.63–0.69 (Fig. 2a). For the ½110�/[001]
mechanism for the selected orientations was sliding. It was found that sample, the friction coefficient values are also low (0.48–0.58) within
during wear the deformation relief on the sample faces evolves with the the first 39 h of friction, but later they increase to 0.58–0.67.
sequential involvement of different slip systems depending on the cur­ The sample wear is nonuniform. The ½110�/½110� single crystal wears
rent stress. Then, we performed experimental studies for single crystals out slightly within the first 48 h of friction, while within 48–56 h the
with the normal pressure axis orientation ½1071� and the friction axis wear increases significantly (Fig. 2b). The wear of the ½110�/[001] single
directed along ½342� [46]. In this case, possible deformation mechanisms crystal from the start of friction is much greater than that of ½110�/½110�.
with regard to the crystallographic orientation were both slip and The next time interval (27–31 h) is characterized by a sharp increase in
twinning. In addition to analyzing the deformation relief evolution and the wear rate. In both cases, the wear rate first reaches its maximum,
the stress-strain state, we established a correlation between the wear and then decreases and stabilizes at a level of ~0.03–0.04 mm3/h. The ob­
the parameters of acoustic emission signals. tained data indicate that the average wear rate of ½110�/½110� single
This paper is aimed to consider the deformation near the worn sur­ crystals is 0.021 mm3/h, while for ½110�/[001] samples it is 0.07 mm3/h.
face of Hadfield steel single crystals and to correlate it with the calcu­
lated slip and twin stresses induced by different orientations of the
friction force at a constant normal load. The reorientation behavior of 3.2. Stress-strain analysis
the worn surface will be studied for regions with different orientations of
the principal stresses. Changes in the dislocation structure with distance 3.2.1. Sample with ½110�/½110� orientations
from the worn surface will be analyzed. In the considered case, the normal load is oriented along the ½110�
direction, and the friction force acts along ½110�. The normal load di­
2. Materials and methods rection is perpendicular to the ð111Þ and ð111Þ planes; the friction force
direction is normal to the (111) and ð111Þ planes. This means that
In the present paper, the deformation behavior during friction was deformation in these planes under these forces is impossible and the
studied using Hadfield steel single crystals. The preparation technique nominal stresses are zero. The close-packed direction ½110� in the (111)
and properties of the samples are described in previous works [46,47].
and ð111Þ planes is perpendicular to the applied normal load. The close-
The size of single crystal samples was 3.6 � 3.6 � 10 mm3. A disc made
packed direction ½110� in the ð111Þ and ð111Þ planes is normal to the
of 105WCr6 steel was used as a counterbody. Friction tests (Fig. 1) were
applied friction force. Therefore, the nominal stresses will be zero in
conducted by dry sliding at a load of 21 N on samples with the normal
these directions under the action of the applied forces. The nominal
load/friction force orientations ½110�/½110� and ½110�/[001]. Friction
twinning stresses in the ð111Þ and ð111Þ planes will also be zero under
was carried out at a slip velocity of 0.1 m/s.
The wear resistance of the studied samples was compared by the normal load, while in the (111) and ð111Þ planes the nominal
stresses will be zero under the friction force.

2
D.V. Lychagin et al. Tribology International 147 (2020) 106284

Fig. 2. Time variation of the friction coefficient (a) and wear rate (b) for Hadfield steel single crystals with the friction force orientation ½110� and [001].

The same maximum values of nominal slip stresses (0.66 MPa) were dislocation glide and mechanical twinning. However, the acting forces
determined for four slip systems: (111)½101�, (111)½011�, ð111Þ½011�, and are oriented perpendicular to the close-packed direction ½110�. The
ð111Þ½101�. The only acting force in all these systems is the normal vector of the friction force is also perpendicular to the close-packed
pressure force. The maximum nominal stresses under the friction force direction [110].
in the ð111Þ and ð111Þ planes are much lower (0.37 MPa) than those Hence the stresses in these directions are zero. Unlike the ½110�/½110�
under the normal load in the (111) and ð111Þ planes. Consequently, the single crystal, the normal pressure and friction forces in the (111) plane
normal pressure force governs shear deformation in the single crystal for act in two common directions ½101� and ½011�, yielding a larger total
the selected friction force direction. nominal stress (1.03 MPa). The opposite situation is observed in the
The estimated twinning stresses in the considered case exceed the ð111Þ plane: the forces are directed in opposite directions. They suppress
nominal slip stresses. The maximum total values of 0.76 MPa are ach­ each other and reduce the total nominal stress to low values (0.29 MPa).
ieved in the systems (111)½112� and ð111Þ½112�. For other systems they Thus, we have two possible slip systems (111)½101� and (111)½011� with
are lower: 0.43, 0.38, and 0.21. Note that the stress caused by the normal the same stress level. The activity of these slip systems should result in
pressure force activates one pair of octahedral slip planes, and the the appearance of non-intersecting slip bands on the (001) face of single
friction force activates the other one. With the chosen calculation crystals (Figs. 3 and 4).
method, the nominal twinning stress due to the friction force is lower The twinning directions are not perpendicular to the applied forces,
because it is proportional to the friction coefficient value. The friction which is favorable for mechanical twinning during sample deformation.
force value decreases exponentially with distance from the worn surface. For the considered active slip systems, preferred are the following
It can be assumed that using the coefficient of friction we can derive twinning directions: ½112� and ½121� for the ½011� direction in the (111)
information only about an averaged value for the subsurface region. plane, ½112� and ½211� for the ½101� direction in the (111) plane (Fig. 3).
A comparison of the nominal twinning and slip stresses shows the The nominal twinning stresses in this case exceed the nominal slip
probability of one or the other mechanism in competing twinning and stresses. The maximum value (1.19 MPa) is achieved in the system
slip systems. The comparison results are given in the form of ratios in (111)½112�. The ratio τtw/τsl ¼ 1.05 (Table 2) indicates that slip and
Table 1. The ratio τtw/τsl ¼ 1.15 in the considered case indicates that slip mechanical twinning may occur simultaneously in the given case.
and mechanical twinning probably occur simultaneously. However, the value of τtw/τsl ¼ 1.05 is much lower than that for the
½110�/½110� sample (τtw/τsl ¼ 1.15, Table 2). Hence the probability of
3.2.2. Sample with ½110�/½001� orientations
mechanical twinning in the ½110�/[001] crystal is lower.
Unlike the ½110�/½110� sample, the friction force vector in ½110�/½001� The theoretically calculated nominal shear stresses should activate
is not perpendicular to the slip planes. Therefore, the friction force can particular slip systems and give a surface deformation pattern consisting
affect the deformation processes in all four planes by the mechanism of of sets of slip bands, as shown in Fig. 3. The verification of the theoretical
shear pattern analysis will be done in the next section.
Since the nominal acting stresses in the (111) and ð111Þ planes are
Table 1
equal, shear can occur in both planes, which should be observed
Maximum nominal stresses for the normal load/friction force orientations
experimentally on the side faces of single crystals (Figs. 3 and 4). Thus,
½110�/½110�
for the considered active slip systems, preferred are the following
Slip Close- Total Partial Total Twinning-
twinning directions: ½112� and ½121� for the ½011� direction in the (111)
plane packed nominal dislocation nominal to-slip
direction slip stress Burgers twinning stress ratio plane, ½112� and ½211� for the ½101� direction in the (111) plane, ½112�
(τsl, MPa) vector stress (τtw, (τtw/τsl) and ½121� for the ½011� direction in the ð111Þ plane, ½112� and ½211� for
MPa)
the ½101� direction in the ð111Þ plane (Fig. 5).
(111) ½101�½011� 0.66 ½112� 0.76 1.15
0.66
ð111Þ ½011�½101� 0.66 ½112� 0.76 1.15
0.66

3
D.V. Lychagin et al. Tribology International 147 (2020) 106284

Fig. 3. Schematic view of shear directions and slip planes in Hadfield steel single crystals with different crystallographic orientations of the normal load/friction
force ½110�/½110� and ½110�/½001�

Table 2
Maximum nominal stresses for the normal load/friction force orientations
½110�/½001�
Slip Close- Total Partial Total Twinning-
plane packed nominal dislocation nominal to-slip
direction slip stress Burgers twinning stress ratio
(τsl, MPa) vector stress (τtw, (τtw/τsl)
MPa)

(111) ½011�½101� 1.03 ½112� 1.19 1.05

Fig. 4. Schematic view of slip band patterns. Fig. 5. Shear directions in slip and twinning systems with the normal load/
friction force orientations ½110�/½110� and ½110�/½001�

4
D.V. Lychagin et al. Tribology International 147 (2020) 106284

3.3. Surface deformation pattern area is under nominal compressive/tensile stresses. The orientation here
changes from <110> to <211>. This agrees well with the knowledge
An experimental study of the revealed sets of slip bands on the faces about FCC crystal reorientation. The orientation of tensile FCC single
of single crystals is shown in Figs. 6 and 7. The ½110�/½110� sample has crystals changes similarly [50]. A similar friction-induced change in the
two intersecting sets of slip bands on the ½110� face and one set (parallel orientation of copper single crystals was observed in a detailed study of
to the plane of friction) on the ½001� and [001] faces (Fig. 6). According reorientation in the wear lip zone [49].
to the theoretical analysis (see Section 3.2.1), these bands belong to the EBSD analysis revealed a simultaneous rotation of both the friction
and normal load axes. In so doing, individual microregions rotate in two
(111) and ð111Þ planes. They are induced by the normal pressure force.
different directions towards ½321� and ½521�. The left side of the EBSD
The ½110�/½001� sample exhibits one system of slip bands on the
pattern is subjected to normal pressure and is rotated by 20� towards
[001] and ½110� faces and two systems on the ½110� face (Fig. 7). Ac­
½321�. Microregions in the center and on the right side are rotated by
cording to the theoretical analysis (see Section 3.2.1), these bands
belong to the (111) plane. They are formed due to the combined action 25.4� towards ½521�. The reorientation of these areas indicates that the
of normal load and friction force. normal load axis rotates together with the friction axis (Table 3).
Changes in the crystallographic orientation of microregions on the
worn surface make the sample deformation nonuniform. Thus, the
3.4. EBSD analysis of reorientations on the worn surface of ½110�/½110� reorientation of local microregions can change the stress-strain state
and ½110�/[001] samples near the worn surface. This will affect the ability to implement one or the
other deformation mechanism (slip or twining). Changes in the stress-
Friction is accompanied by inhomogeneous stress distribution in the strain state were assessed by determining acting stresses in slip sys­
friction zone [48,49]. The leading face of the test body (the pin in the tems with account for the reorientation of the considered microregions.
pin-on-disk configuration) is subjected to nonuniform multiaxial In the case when the axes of the applied forces rotate towards the
compression. The region near the trailing face is under compressive/­ friction force/normal load directions ½103�/½321�, shear can occur only
tensile stresses. A similar pattern is observed near the side faces, but in the (111) plane by slip in the ½101� direction. Twinning is suppressed
with different principal stress components. The nominal stress caused by because the nominal twinning stress is lower than the nominal slip
the friction force rapidly decreases with distance from the worn surface. stress. When the axes of the applied forces rotate towards the friction
Therefore, we may say that the component of the nominal force/normal load directions ½104�/½521�, shear can occur in two ð111Þ
friction-induced stress dominates in the immediate vicinity of the worn and (111) planes by slip in the [101] and ½101� directions, respectively.
surface, while that of the nominal normal pressure stress is dominant far In this case, twinning is suppressed because the twinning stress is lower
away from it. The intermediate region is under the conditions presented than the slip stress (Table 4).
in the main calculations by the combined action of normal load and EBSD analysis shows the simultaneous rotation of both the friction
friction force. and normal load axes. Unlike the previous case, here we observe a well-
The consideration for the contact friction force and its interaction defined macroshear due to which the normal load axis rotates by 30�
with the sliding friction force in different parts of the contact surface towards the ½211� direction. According to the worn surface reorientation
[49] may contribute to a more rigorous validation of nonuniform wear analysis, the friction axis rotates by 18.5� towards the ½120� direction
lip formation in the sample. The occurrence of one or the other defor­ (Table 5).
mation mechanism depends on the crystallographic orientation and In the given case, the reorientation close to the worn surface is more
stress in slip systems and differs for compression and tension. Conse­ significant and affects a large part of the sample surface. The change in
quently, deformation on the worn surface will occur by different the single crystal orientation strongly affects the nominal shear stress
mechanisms depending on the principal stress orientation in local re­ redistribution in active slip systems in the friction contact zone. In the
gions of the crystal and will produce different nominal stress levels. case when the applied force axes rotate towards the friction force/
The above assumption was verified for both investigated crystallo­
normal load directions ½120�/½211�, shear can occur in two (111) and
graphic orientations of single crystals. The orientation of microregions
ð111Þ planes by slip in the ½101� and ½101� directions, respectively.
was determined by EBSD scanning of the local areas shown in Fig. 8 with
Analysis of nominal stresses in slip and twinning systems (Table 6) in­
different orientation of the principal stresses.
dicates that the priority of the active slip system changes compared to
Figs. 9–12 show the reorientation after friction in the scan area.
the initial orientation. Twinning in this case is suppressed because the
According to the presented results, the ½110�/[001] samples have a
nominal twinning stress is either lower or slightly exceeds the nominal
greater resistance to reorientation compared with the ½110�/½110� sam­ slip stress.
ples. In the second case, the reorientation in the scan area is most The obtained results suggest that the conditions for slip and twinning
intensive at the trailing face (relative to the friction force direction). This

Fig. 6. Slip bands on the faces ½001� (a), ½110� (b), and [001] (c) of Hadfield steel single crystals with the friction force orientation.½110�

5
D.V. Lychagin et al. Tribology International 147 (2020) 106284

Fig. 7. Slip bands on the faces ½110� (a), ½001� (b), and ½110� (c) of Hadfield steel single crystals with the friction force orientation [001].

Fig. 8. Schematic illustration for EBSD scanning of the worn surface showing the reorientation of single crystals depending on the friction force direction: ½110� – a,
[001] – b.

Fig. 9. Orientation maps and inverse pole figures of the worn surface of ½110�/[001] Hadfield steel single crystal. Orientation of crystallographic axes relative to the
sample and the scan areas on the end face (the images highlight boundaries with a misorientation angle of more than 2� ).

deformation immediately under the worn surface and in the friction features that allow them to be distinguished as separate layers and
contact area are noticeably different due to different stress tensors. The interpreted as separate types of dislocation substructures. The main
reorientation of local micro- and macroregions near the worn surface types of substructures formed during severe plastic deformation are
mechanically suppresses twinning in both samples. discussed, e.g., in Ref. [51,52].
The observation will be performed in the direction towards the worn
3.5. TEM study of subsurface deformation surface. The substructure evolution will be similar to its evolution with
increasing strain, when the material is deformed with a constant rate.
While discussing the deformation mechanisms, we should consider We will determine the stages of dislocation substructure evolution
structural changes in the vicinity of the worn surface. The dislocation which take place in the immediate vicinity of the surface during the
structure of the sample with the normal load/friction force orientation initial stage of friction.
½110�/½110� is represented by successive images of the worn face to a The dislocation density far from the worn surface is low. The sub­
depth of 20 μm (Fig. 13a). Its fragments were also studied at a greater structure is represented by single and intersecting dislocations. Inter­
distance from the surface. secting dislocations quickly form dislocation pile-ups. Closer to the worn
The substructures in Fig. 13 demonstrate a set of morphological surface, the average dislocation density increases and dislocation cells

6
D.V. Lychagin et al. Tribology International 147 (2020) 106284

Fig. 10. Schematic illustration for the reorientation of microregions of Hadfield


steel single crystal near the worn surface after friction for the normal load/
friction force orientations ½110�/[001]. Fig. 12. Schematic illustration for the reorientation of a macroregion of Had­
field steel single crystal near the worn surface after friction for the normal load/
friction force orientations ½110�/½110�
are formed (Fig. 13b). Then their size decreases, and the dislocation
density in the cell walls increases. These changes in the formed cellular
dislocation substructure upon approaching the worn surface are Table 3
consistent with its changes with increasing plastic strain [53], indicating To the reorientation analysis of ½110�/[001] single crystal.
that deformation develops in the direction towards the surface. Acting force Initial direction After rotation Rotation angle
Upon further approaching the worn surface, the cellular substructure
exhibits single elongated deformation twins and then twin systems of Rotation towards the ½321� direction

small width. This layer with a thickness of approximately 9 μm can be Normal load ½110� ½321� 20�

called a layer of inactive twinning (Fig. 13a). The considered layers or Friction [001] ½103� 18.4�
zones characterize the slip and twinning deformation. Their occurrence Rotation towards the ½521� direction
in the direction from the worn surface, as follows from the results given Normal load ½110� ½521� 25.4�
above and obtained for other single crystal orientations [45,46], de­ Friction [001] ½104� 14�
pends on the crystallographic orientation of the compression and fric­
tion axes, the local stress tensor, and the test time. The latter indicates
the cyclic development of the hardening and wear processes. of size 0.2 μm, which are destroyed by newly formed subboundaries. The
The next layer of thickness 4 μm is represented first by one and then next substructure is represented by groups of microband systems with
by two intersecting twin systems against the background of the cellular broken and closed subboundaries, as well as by a substructure with a
dislocation substructure. A 3 μm thick layer of two twin systems is chaotic subboundary distribution and bending extinction contours. The
bounded by bundles of one twin system 1 μm wide (Fig. 13c). It can be thickness of the layer occupied by the given substructure is approxi­
assumed that the number of twin systems is determined by the orien­ mately 1.5 μm. The subboundaries in this layer are oriented at angles of
tation and reorientation of the considered local region. 15, 30, and 60� to the worn surface (Fig. 13d). The penultimate sub­
Then, the substructure of deformation microtwins changes for the surface layer is a layer of nanobands and nanotwins (Fig. 13d). It has the
microbanded substructure. It is represented mainly by one system of minimum width of 0.25 μm. The microdiffraction pattern from the
microbands 1.5 … 2.0 μm wide, which are oriented at a small angle to substructure with more ordered subboundaries differs from that with
the friction surface. The distance between the subboundaries in this chaotic discrete subboundaries and bending extinction contours. In the
substructure is 0.2 … 0.4 μm. The microbands contain dislocation cells first case, reflections in the diffraction patterns are split. In the second,

Fig. 11. Orientation maps and inverse pole figures of the worn surface of ½110�/½110� Hadfield steel single crystal. Orientation of crystallographic axes relative to the
sample and the scan areas on the end face (the images highlight boundaries with a misorientation angle of more than 2� ).

7
D.V. Lychagin et al. Tribology International 147 (2020) 106284

Table 4
the reflection is extended (inserts in Fig. 13d).
Maximum nominal stresses after reorientation towards the ½321� and ½521�
The identified bending extinction contours point to the presence of
directions. long-range stress fields and the insignificant role of relaxation processes.
A small fraction of closed subboundaries and the absence of sub­
Slip Close- Total Partial Total Twinning-
plane packed nominal dislocation nominal to-slip stress
boundaries with a characteristic banded contrast indicate that there is
direction slip stress Burgers twinning ratio (τtw/ no dynamic recrystallization. A similar structure was observed in a
(τsl, MPa) vector stress (τtw, τsl) Fe–Ni alloy compressed by 60 … 70% (1.0 … 1.2 logarithmic strain)
MPa) [52]. The dislocation structure between dislocation walls is represented
Friction force/normal load orientations ½103�/½321� by an inhomogeneous cellular or mesh-cellular substructure, which is
(111) ½101� 0.9 ½112� 0.84 0.93 formed as a result of failure of the cellular dislocation substructure.
Friction force/normal load orientations ½104�/½521�
In the layer with the microbanded substructure, twinning occurs
½101� 1.16 0.56 0.48
within bands. They can be defined as deformation nanotwins by their
ð111Þ ½121�
size. This process is most intense in a layer of thickness 0.2 … 0.4 μm
(111) ½101� 1.07 ½112� 0.92 0.86
consisting of one or two deformation microbands. The width of nano­
twins is 10 … 15 nm. The inclination angle with respect to the sub­
boundary is 30–45� .
Table 5 In the vicinity of the surface, there is a 0.5 μm thick layer consisting
To the reorientation analysis of ½110�/½110� single crystal. of 10–20 nm nanograins. The nanograins comprise parallel nanotwins
spaced at several nanometers (Fig. 13e). The microdiffraction pattern
Acting force Initial direction After rotation Rotation angle
from the nanograins is almost circular (inserts in Fig. 13e).
Normal load ½110� ½211� 30� The surface exhibits the zone of interaction between the sample
Friction ½110� ½120� 18.5� material and the disk (mixing zone).

4. Discussion

The study of dry sliding friction of Hadfield steel single crystals


Table 6
showed that the values of the coefficient of friction (CoF) for the
Maximum nominal stresses for the friction force/normal load orientations
considered samples are on average close. However, the coefficient
½120�/½211�
changes in two stages. At the beginning of friction, the average CoF
Slip Close- Total Partial Total Twinning-
values for the ½110�/½110� and ½110�/[001] samples are equal to ~
plane packed nominal dislocation nominal to-slip stress
direction slip stress Burgers twinning ratio (τtw/
0.42–0.53. Then they increase to ~0.63–0.67 in both cases. The increase
(τsl, MPa) vector stress (τtw, τsl) in CoF may be due to sample hardening during deformation as well as
MPa) due to reorientation in the subsurface region.
(111) ½110� 0.7 ½211� 0.71 1.01 Higher hardening at the second stage of friction (after the CoF in­
0.75 0.71 0.95
crease) is caused by the accumulation of defects, achievement of the
ð111Þ ½101� ½211�
maximum density of dislocations and twins due to intense plastic
deformation, fragmentation, and subsurface region reorientation.
Higher CoF values (0.67 versus 0.63 for the ½110�/[001] sample) for the

Fig. 13. Dislocation structure changes from the worn surface. General view (a). Dislocation cell structure (b). Microtwin system and dislocation cell structure (c).
Microbands and nanobands with nanotwins (d). Nanograins with nanotwins (e).

8
D.V. Lychagin et al. Tribology International 147 (2020) 106284

½110�/½110� sample at the second stage of friction are consistent with our and induce local fracture, leading to delamination of the material and
assumption that its work hardening is more intense due to slip in the wear. EBSD data confirm the different nature of the principal stress
(111) and ð111Þ planes as well as twinning. Twinning in the ½110�/½110� tensor in different parts of the surface. Consequently, there can arise
sample occurs mainly at the initial stage of friction, though it may also tensile stresses that will contribute to crack opening.
occur after reorientation of the subsurface region, as evidenced by the With further friction the gradient accumulation of deformation will
analysis of the τtw/τsl ratio (see Sections 3.2 and 3.4.). Another indica­ continue, leading to hardening. This explains the cyclic nature of wear
tion of its higher work hardening is the higher wear resistance as and hardening characteristic of the studied samples during friction. The
compared with the ½110�/[001] sample. periodicity of the process depends on the number of active slip systems
The wear rate stabilization at the level of ~ (0.03–0.04) mm3/h at present in single crystals with different crystallographic orientations.
the second stage of friction (after the CoF increase) is related to the After reorientation, the conditions for slip and twinning in the vi­
combined action of two factors. Firstly by the formation of a nano­ cinity of the worn surface and in the friction contact area differ
crystalline surface layer, which is a mechanical mixture of metal and noticeably due to the change in the Schmid factor for these systems. In
oxide wear particles, and secondly by the single crystal reorientation both cases there will be two active slip systems. These are (111)½101� and
near the worn surface. This assumption is confirmed by EBSD data on ð111Þ½101� for the ½110�/½110� sample, and (111)½110� and ð111Þ ½101�
the reorientation of micro- and macroscopic subsurface regions of single for the ½110�/[001] sample. The reorientation of local micro- and mac­
crystals under friction. roregions near the worn surface contributes to the suppression of
At the initial stage of friction, the ½110�/½110� sample is deformed by twinning in both samples. However, the conditions for twinning can be
slip along the (111) and ð111Þ planes. Four slip systems have the same favorable in a local layer. This is indirectly confirmed by TEM which
high slip stresses. The action of equally loaded slip systems in adjacent detects twins in the subsurface layer (Fig. 13).
planes highly promotes the formation of barriers and hence rapid The observed changes in the dislocation structure depending on the
hardening, leading to an even higher critical twinning stress for systems distance to the worn surface indicate that the deformation substructure
in these planes. In each plane, only one twinning system can be active. parameters vary widely already at a 20 μm distance (Fig. 14). A com­
Therefore, it is very probable that twin systems will interact with each parison with severe plastic deformation [5,6] shows that upon
other and dislocations in slip systems will interact with the twin approaching the worn surface the types of substructures change simi­
boundaries. The ½110�/[001] sample has a high slip stress in two systems larly to their evolution with increasing plastic strain. The principal stress
in the (111) plane and an even higher twinning stress for one system in components under uniaxial compression change at 30–40% strain when
this plane. The slip and twinning stresses in other systems are lower. As a the sample height to width ratio is equal to 2. In this case, other stress
result, we should expect less hardening of this sample than for the tensor components become nonzero (Fig. 15a).
The substructure refinement is promoted by the change in the stress
½110�/½110� sample.
tensor during deformation. Such substructures are formed under com­
There is a significant dislocation density gradient near the worn
plex loading conditions, e.g., high-speed pounding [19]. Consequently,
surface (Figs. 13 and 14). Hardening due to subsurface layers reduces
the processes in the region near the worn surface are determined by all
the tendency of the subsurface layer to fatigue failure. Above the
nonzero principal stress components.
nanograin layer there is a structure of deformation micro/nanobands
The next region is subjected to the compressive and frictional force
and twins which additionally contribute to hardening and prevent the
components acting in the direction from the worn surface. The influence
propagation of deformation deep into the material. At the same time,
of the friction force decreases sharply with distance from the friction
deformation is quite intense only in a strongly deformed narrow sub­
contact area.
surface region. At a sufficient stress level, cracks will initiate in this zone
Later the role of the friction force decreases so much that the
deformation can be considered uniaxial. Fig. 15 schematically shows the
behavior of the stress and substructure parameters under compression
and friction. The figure bears witness to the fact that strong structural
changes can occur only if all principal stress components are different
from zero. Moreover, the stress tensor must have the maximum intensity
for the occurrence of misorientations in local regions. As follows from
the expression for this quantity, the greater is the difference in the values
of the principal stress components, the higher is the stress intensity. The
difference in the principal stress components is most pronounced in the
immediate vicinity of the worn surface.
According to the literature data [48] as well as our EBSD results for
the subsurface orientation, the stress tensor on the worn surface also
varies in different surface areas of Hadfield steel single crystals.
All data indicate that the stress tensor difference in different surface
and bulk regions will have a different effect on slip and twinning in
various local regions of samples with the same crystallographic orien­
tation. Consequently, the difference will have a different effect on the
orientation change in these areas.
To conclude, the change in the crystallographic orientation of the
side faces relative to the friction force led to shear stress redistribution.
Fig. 14. Schematic view of changes in the types and parameters of dislocation
As a result, the orientation of different surface areas changed in different
substructures depending on the distance from the worn surface. The arrows crystallographic directions and with different rate. According to our
point to the regions with characteristic types of dislocation structure: I – calculations, the most active systems in the ½110�/½110� single crystals
nanograins with nanotwins, II – nanobanded with nanotwins, III – fragmented were (111)½101� and ð111Þ½101�, while in the ½110�/[001] crystals these
microbanded, IV – microbanded, V – two systems of microtwins, VI – one were (111)½101� and ð111Þ½101�.
system of microtwins, VII – cellular with twins, VIII – cellular, IX– mesh-
cellular. The dotted lines indicate arbitrary boundaries of regions with dislo­
cation structures.

9
D.V. Lychagin et al. Tribology International 147 (2020) 106284

Fig. 15. Deformation behavior of the material under quasi-static plastic deformation (a) and in friction (b): σi are the principal stress components, ρ is the scalar
dislocation density, d is the dislocation cell size, H is the distance from the worn surface.

5. Conclusions Kolubaev: Writing - review & editing. L.L. Lychagina: Investigation,


Writing - review & editing.
We have studied the physical deformation mechanisms of Hadfield
steel single crystals under dry sliding friction by comparing the wear and Acknowledgements
deformation behavior of samples with the same normal load direction
but different friction force directions. According to the obtained results: EBSD investigations were performed using the equipment of the
Analytical Center of Geochemistry of Natural Systems of Tomsk State
1. The coefficient of friction changes nonmonotonically in both types of University.
single crystals. At the first stage of friction, the average CoF values TEM studies were carried out using the equipment of the Nano-
for the ½110�/½110� and ½110�/[001] samples are ~ 0.42–0.53. Then, Center of Tomsk Polytechnic University.
these values increase to ~0.63–0.67 in both cases. Friction tests were conducted with the financial support of the
2. The average wear rate of the ½110�/½110� single crystals is 0.021 Fundamental Research Program of the State Academies of Sciences for
mm3/h, while that for the ½110�/[001] samples is 0.07 mm3/h. The 2013–2020 (Project No. III.23.2).
higher wear resistance is due both to a lower slip stress in the
½110�/½110� sample and to a different reorientation of the subsurface References
region.
[1] Ge S, Wang Q, Wang J. The impact wear-resistance enhancement mechanism of
3. The conditions for slip and twinning deformation in the vicinity of
medium manganese steel and its applications in mining machines. Wear 2017;
the worn surface differ noticeably due to different values of the stress 376–377:1097–104. https://doi.org/10.1016/j.wear.2017.01.015.
tensor. The reorientation of local micro- and macroregions near the [2] Lencina R, Caletti C. Assessing wear performance of two high-carbon Hadfield
steels through field tests in the mining industry. Proc Mater Sci 2015;9:358–66.
worn surface promotes the mechanical suppression of twinning in
https://doi.org/10.1016/j.mspro.2015.05.005.
both samples. [3] Chowdhury P, Canadinc D, Sehitoglu H. On deformation behavior of Fe-Mn based
4. Dislocation structure analysis showed that the dominant deforma­ structural alloys. Mater Sci Eng R 2017;122:1–28. https://doi.org/10.1016/j.
tion mechanisms are slip and twinning. The priority of these mech­ mser.2017.09.002.
[4] Karaman I, Sehitoglu H, Gall K, Chumlyakov YI. On the deformation mechanisms in
anisms in the discussed gradient structure is determined by the single crystal Hadfield manganese steels. Scripta Mater 1998;38:1009–15. https://
proximity of the considered deformation zones to the friction doi.org/10.1016/S1359-6462(97)00581-2.
surface. [5] Karaman I, Sehitoglu H, Gall K, Chumlyakov YI, Maier HJ. Deformation of single
crystal Hadfield steel by twinning and slip. Acta Mater 2000;48:1345–59. https://
5. A complex analysis of sample wear intensity and active deformation doi.org/10.1016/S1359-6454(99)00383-3.
mechanisms suggests that the combined action of slip and twinning [6] Canadinc D, Sehitoglu H, Maier HJ, Niklasch D, Chumlyakov YI. Orientation
enhances the wear resistance of samples. evolution in Hadfield steel single crystals under combined slip and twinning. Int J
Solid Struct 2007;44:34–50. https://doi.org/10.1016/j.ijsolstr.2006.04.011.
6. The substructural evolution indicates that the plastic strain increases [7] Astafurova EG, Chumlyakov YI. Strain hardening upon twinning of [-111].[-144]
upon approaching the worn surface. The difference in local stresses and [011] single crystals of Hadfield steel. Phys Met Metallogr 2009;108:510–8.
contributes to the misorientation of microregions and formation of https://doi.org/10.1134/S0031918X09110118.
[8] Astafurova EG, Tukeeva MS, Zakharova GG, Melnikov EV, Maier HJ. The role of
misoriented substructures.
twinning on microstructure and mechanical response of severely deformed single
crystals of high-manganese austenitic steel. Mater Char 2011;62:588–92. https://
Declaration of competing interest doi.org/10.1016/j.matchar.2011.04.010.
[9] Efstathiou C, Sehitoglu H. Strain hardening and heterogeneous deformation during
twinning in Hadfield steel. Acta Mater 2010;58:1479–88. https://doi.org/
The authors declare that they have no known competing financial 10.1016/j.actamat.2009.10.054.
interests or personal relationships that could have appeared to influence [10] Adler PH, Olson GB, OwenWS. Strain hardening of Hadfield manganese steel.
the work reported in this paper. Metall Mat Trans A 1986;17(10):1725–37. https://doi.org/10.1007/BF02817271.
[11] Niendorf T, Rüsing CJ, Frehn A, Chumlyakov YI, Maier HJ. Deformation
mechanisms in high-manganese steels showing twinning-induced plasticity: fine-
CRediT authorship contribution statement grained material and single crystals at ambient and cryogenic temperatures.
Scripta Mater 2012;67:875–8. https://doi.org/10.1016/j.scriptamat.2012.08.011.
[12] Kim J, De Cooman BC. On the stacking fault energy of Fe-18 pct Mn-0.6 pct C-1.5
D.V. Lychagin: Conceptualization, Writing - review & editing. A.V. pct Al twinning-induced plasticity steel. Metatall Mater Trans A-Phys Metall Mater
Filippov: Investigation, Writing - review & editing. O.S. Novitskaya: Sci 2011;42A(4):932–6. https://doi.org/10.1007/s11661-011-0610-6.
Investigation. Y.I. Chumlyakov: Conceptualization, Data curation. E.A. [13] Lee W-S, Chen T-H. Plastic deformation and fracture characteristics of Hadfeld
steel subjected to high-velocity impact loading. Proc Inst Mech Eng Part C: J

10
D.V. Lychagin et al. Tribology International 147 (2020) 106284

Mechanical Engineering Science 2002;216:971–82. https://doi.org/10.1243/ [35] Yan W, Fang L, Zheng Z, Sun K, Xu Y. Effect of surface nanocrystallization on
095440602760400940. abrasive wear properties in Hadfield steel. Tribol Int 2009;42:634–41. https://doi.
[14] Bal B, Gumus B, Gerstein G, Canadinc D, Maier HJ. On the micro-deformation org/10.1016/j.triboint.2008.08.012.
mechanisms active in high-manganese austenitic steels under impact loading. [36] Gnyusov SF, Tarassov SY. Friction and the development of hard alloy surface
Mater Sci Eng, A 2015;632:29–34. https://doi.org/10.1016/j.msea.2015.02.054. microstructures during wear. Conf J Mater Eng Perf 1996;6:737–42. https://doi.
[15] Gumus B, Bal B, Gerstein G, Canadinc D, Maier HJ, Guner F, et al. Twinning org/10.1007/s11665-997-0075-3.
activities in high-Mn austenitic steels under high-velocity compressive loading. [37] Atabaki MM, Jafari S, Abdollah-pour H. Abrasive wear behavior of high chromium
Mater Sci Eng A 2015;648:104–12. https://doi.org/10.1016/j.msea.2015.09.045. cast iron and Hadfield steel-A comparison. J Iron Steel Res Int 2012;19(4):43–50.
[16] Berns H, Gavriljuk VG, Petrov YN, et al. Microstructural changes in the wear https://doi.org/10.1016/S1006-706X(12)60086-7.
surface of high strength stainless austenitic steels. Mater Werkst 2003;34(10–11): [38] Abbasia M, Kheirandish S, Kharrazi Y, Hejazi J. On the comparison of the abrasive
960–5. https://doi.org/10.1002/mawe.200300686. wear behavior of aluminum alloyed and standard Hadfield steels. Wear 2010;268:
[17] Lindroos M, Laukkanen A, Cailletaud G, KuokkalaV -T. Microstructure based 202–7. https://doi.org/10.1016/j.wear.2009.07.010.
modeling of the strain rate history effect in wear resistant Hadfield steels. Wear [39] Dalai R, Das S, Das K. Effect of thermo-mechanical processing on the low impact
2018;396–397:56–66. https://doi.org/10.1016/j.wear.2017.11.007. abrasion and low stress sliding wear resistance of austenitic high manganese steels.
[18] Korshunov LG, Sagaradze VV, ChernenkoNL. Structural and phase transformations Wear 2019;420–421:176–83. https://doi.org/10.1016/j.wear.2018.10.013.
in Hadfield steel upon frictional loading in liquid nitrogen. Phys Met Metallogr [40] Chen C, Lv B, Ma H, Sun D, Zhang F. Wear behavior and the corresponding work
2016;117(8):828–33. https://doi.org/10.1134/S0031918X16080068. hardening characteristics of Hadfield steel. Tribol Int 2018;121:389–99. https://
[19] Feng XY, Zhang FC, Yang ZN, Zhang M. Wear behaviour of nanocrystallised doi.org/10.1016/j.triboint.2018.01.044.
Hadfield steel. Wear 2013;305:299–304. https://doi.org/10.1016/j. [41] Feng Y, Song R, Peng S, Pei Z, Song R. Microstructures and impact wear behavior of
wear.2012.11.038. Al-alloyed high-Mn austenitic cast steel after aging treatment. J Mater Eng Perform
[20] Tukeeva MS, Melnikov EV, Maier HJ, Astafurova EG. Structural features and 2019;28(8):4845–55. https://doi.org/10.1007/s11665-019-04265-y.
mechanical properties of austenitic Hadfield steel after high-pressure torsion and [42] Zambrano OA, Valdes J, Rodriguez LA, Reyes D, Snoeck E, Rodriguez SA,
subsequent high-temperature annealing//the Physics of Metals and Metallography. Coronado JJ. Elucidating the role of κ-carbides in Fe-Mn-Al-C alloys on abrasion
113, 6: 612-662, https://doi.org/10.1134/S0031918X12060129; 2012. wear. Tribol Int 2019;135:421–31. https://doi.org/10.1016/j.
[21] Koneva NA, Kozlov EV. Physical nature of stages in plastic deformation. Russ Phys triboint.2019.03.002.
J 1990;33(2):165–79. https://doi.org/10.1007/BF00894514. [43] Luo ZC, Ning JP, Wang J, Zheng KH. Microstructure and wear properties of TiC-
[22] Koneva NA, Kozlov EV. Nature of substructural hardening. Russ Phys J 1983;25(8): strengthened high-manganese steel matrix composites fabricated by hypereutectic
681–91. https://doi.org/10.1007/BF00895238. solidification. Wear 2019;432–433:202970. https://doi.org/10.1016/j.
[23] Kozlov EV, Koneva NA. Internal fields and other contribution to flow stress. Mater wear.2019.202970.
Sci Eng 1997;234–236:982–5. https://doi.org/10.1016/S0921-5093(97)00381-X. [44] Zhou Z, Shan Q, Jiang Y, Li Z, Zhang ZX. Effect of nanoscale V2C precipitates on
[24] Koneva NA, Kozlov EV, Trishkina LI. Internal field sources, their screening and the the three-body abrasive wear behavior of high-Mn austenitic steel. Wear 2019;
flow stress. Mater Sci Eng A 2001;319–321:156–9. https://doi.org/10.1016/ 436–437:203009. https://doi.org/10.1016/j.wear.2019.203009.
S0921-5093(01)00945-5.17. [45] Lychagin DV, Filippov AV, Novitskaia OS, Chumlyakov YI, Kolubaev EA,
[25] Dastur YN, Leslie WC. Mechanism of work hardening in Hadfield manganese steel. Sizova OV. Friction-induced slip band relief of -Hadfield steel single crystal
Metall Trans A 1981;12:749–59. https://doi.org/10.1007/BF02648339. oriented for multiple slip deformation. Wear 2017:374–5. https://doi.org/
[26] Owen WS, Grujicic M. Strain aging of austenitic Hadfield manganese steel. Acta 10.1016/j.wear.2016.12.028.
Mater 1998;47:111–26. https://doi.org/10.1016/S1359-6454(98)00347-4. [46] Lychagin DV, Filippov AV, Kolubaev EA, Novitskaia OS, Chumlyakov YI,
[27] Shiekhelsouk MN, Favier V, Inal K, Cherkaoui M. Modelling the behaviour of Kolubaev AV. Dry sliding of Hadfield steel single crystal oriented to deformation by
polycrystalline austenitic steel with twinning-induced plasticity effect. Int J Plast slip and twining: deformation, wear, and acoustic emission characterization. Tribol
2009;25:105–33. https://doi.org/10.1016/j.ijplas.2007.11.004. Int 2018;119:1–18. https://doi.org/10.1016/j.triboint.2017.10.027.
[28] Ye DY, Matsuoka S, Nagashima N, Suzuki N. The low-cycle fatigue, deformation [47] Johnson KL. Contact mechanics. Cambridge University Press; 1985.
and final fracture behaviour of an austenitic stainless steel. Mater Sci Eng, A 2006; [48] Popov VL, Heß M. Method of dimensionality reduction in contact Mechanics and
415(1–2):104–17. https://doi.org/10.1016/j.msea.2005.09.081. friction, vol. 265. Springer-Verlag Berlin Heidelberg; 2015. https://doi.org/
[29] Machado PC, Pereira JI, Penagos JJ, Yonamine T, Sinatora A. The effect of in- 10.1007/978-3-642-53876-6.
service work hardening and crystallographic orientation on the micro-scratch wear [49] Tarasov SYu, Chumaevskii AV, Lychagin DV, Nikonov AY, Dmitriev AI.
of Hadfield steel. Wear; 2017. p. 376–7. https://doi.org/10.1016/j. Inhomogeneous strain development and problem of wear lip formation on copper
wear.2016.12.057. 1064–1073. single crystals in dry sliding. Wear 2018;410–411:210–21. https://doi.org/
[30] Dobrynin SA, Kolubaev EA, AYu Smolin, et al. Time-frequency analysis of acoustic 10.1016/j.wear.2018.07.004.
signals in the audio-frequency range generated during Hadfield’s steel friction. [50] Honeycombe RWK. The plastic deformation of metals. London: Edward Arnold &
Tech Phys Lett 2010;36(7):606–9. https://doi.org/10.1134/S1063785010070072. American Society of Metals; 1984.
[31] Aleshina EA, Ivanov YuF, Kolubaev AV, Konovalov SV, Gromov VE. Gradient [51] Koneva NA, Lychagin DV, Trishkina LI, Kozlov EV. Types of dislocation
structure-phase states formed in Hadfield steel during dry sliding wear. Russ Phys J substructures and stages of stress-strain curves of FCC alloys. Int Conf Proc 1985;1:
2008;51(11):1168–73. https://doi.org/10.1007/s11182-009-9151-5. 21–6. https://doi.org/10.1016/B978-0-08-031642-0.50011-8.
[32] Kolubaev AV, Kolubaev EA, Sizova OV. Nanoindentation of the surface layer of [52] Teplyakova LA, Koneva NA, Lychagin DV, Trishkina LI, Kozlov EV. Evolution of the
Hadfield’s steel after sliding friction. Tech Phys Lett 2007;33(12):1038–42. dislocation structure and the stage of work-hardening of Ni3Fe alloy with [001]
https://doi.org/10.1134/S1063785007120164. orientation. Russ Phys J 1988;31(2):99–103. https://doi.org/10.1007/
[33] Kolubaev AV, Ivanov YuF, Sizova OV, Kolubaev EA, Aleshina EA, Gromov VE. BF00896528.
Effect of elastic excitations on the surface structure of Hadfield steel under friction. [53] Koneva NA, Starenchenko VA, Lychagin DV, Trishkina LI, Popova NA, Kozlov EV.
Tech Phys 2008;53(2):204–10. https://doi.org/10.1134/S1063784208020096. Formation of dislocation cell substructure in face-centred cubic metallic solid
[34] Yan W, Fang L, Sun K, Xu Y. Effect of surface work hardening on wear behaviour of solutions. Mater Sci Eng, A 2008;483–484:179–83. https://doi.org/10.1016/j.
Hadfield steel. Mater Sci Eng, A 2007;460–461:542–9. https://doi.org/10.1016/j. msea.2006.08.140.
msea.2007.02.094.

11

You might also like