Sunteți pe pagina 1din 11

Strength of Materials, Vol. 43, No.

6, November, 2011

EFFECT OF PVD COATINGS ON THE STRAIN AND LOW-CYCLE FATIGUE RESISTANCE


OF STAINLESS STEEL AND TITANIUM ALLOYS

A. P. Gopkalo, N. R. Muzyka, A. V. Rutkovskii, UDC 539.4


and V. P. Shvets

The effect of different PVD coatings and their thickness on the tensile and cyclic elastoplastic strain
resistance of Kh18N10T stainless steel and VT1 and VT20 titanium alloys was experimentally
substantiated. The application of PVD coatings, also after tension or cyclic loading, enhances
tensile and cyclic elastoplastic strain resistance and leads to the change in fracture behavior
(quasistatic to fatigue) and cyclic properties of the material.

Keywords: PVD coating, low-cycle fatigue, stress, strain, damage accumulation, homogeneity parameter.

Introduction. The fracture of structural components is known to be often initiated by the local damage of
surface layers of the material. The mechanical properties of the surface layer are built up in manufacturing the
structural components and affected in service. The major contributors to the fracture of surface layers are their
roughness, stress concentrations, and nonuniform distribution of mechanical properties over the surface and in depth
of the material [1]. Experimental investigations, e.g., [2], demonstrated that in a homogeneous stressed state, the
yield limit of the surface layer of low-carbon steel was 25% lower than that of the base material. As evidenced by
some other findings [3], the yield limit of the thinnest specimens is only 20% of that of the thick ones. Therefore,
plastic flow on the surface sets in earlier than in the middle layers.
Engineering practice employs different treatments, including the application of metallic and nonmetallic
films and coatings, to enhance the strength and protective properties of item surfaces.
As early as 1926, Rosko [4] began studying the effect of surface oxide films on the mechanical properties of
metals in evaluation of the microhardness of cadmium monocrystals. It was revealed that the oxide film less than 20
atoms thick increased the critical stress of cadmium monocrystals by about 50%. It also interferes with plastic strains
in the crystal, which contributes to an increase in the hardness of the surface layer.
Investigations [5–9] showed that not only oxide films but also PVD coatings (physical vapor deposition)
enhanced the tensile and cyclic load resistance, mainly due to the restriction on plastic strains of the base by a hard
surface layer with higher mechanical strength, large residual compression stresses, and good adhesion of the coating
material to the base [7–9]. An increase in the tensile and cyclic load resistance is associated with the thickness and
hardness of PVD coatings and their nitrogen contents [7, 8, 10].
Residual stresses in a PVD-coated material are determined by the difference in the thermal expansion
coefficients of the base and coating materials and the crystal grown in coating application [11]. With thicker coatings
[12], they increase and influence the hardness of the coating [13]. The thicker the coating, the higher the hardness
[8]. However, an increase in the PVD coating thickness results in lower adhesion of coating layers to the base
material, which is caused by large residual compression stresses [14, 15]. Large residual stresses in PVD coatings can
give rise to its local fracture and separation on the component edges if the coating is thicker than the edge radius. The
stressed state at the interface is independent of the coating thickness if it is larger than about three amplitudes of the
base roughness [15].
Institute of Problems of Strength, National Academy of Sciences of Ukraine, Kiev, Ukraine. Translated from
Problemy Prochnosti, No. 6, pp. 23 – 39, November – December, 2011. Original article submitted December 11,
2009.

604 0039–2316/11/4306–0604 © 2011 Springer Science+Business Media, Inc.


Fig. 1. Specimen for monotonic tensile and low-cycle fatigue tests.

It is noted [7, 8, 16] that TiN PVD coatings enhance slightly the tensile strain resistance. Strains exceeding
0.34–0.40% give rise to extensive cracking and separation of the coating material. After the cracking of the coating
layer, strains are passing into the base. Notice that the fatigue life of PVD-coated specimens is extended only in the
range of small stress amplitudes since larger stress amplitudes lead to the coating fracture [7, 16, 17]. Strains on the
base surface cause the coating fracture at early fatigue stages.
The surface treatment of specimens after cyclic loading influences the fatigue strength. The cyclic load
resistance of AISI 4140 steel specimens is enhanced after plasma nitriding of initial specimens, also after their cyclic
loading up to 30% fatigue life as well as after repeated nitriding with preloading up to the same level. Cyclic
preloading up to 60% fatigue life of the initial and nitrided specimens, followed by repeated nitriding, reduced the
cyclic life of this steel [18].
Thus, in tension and under cyclic loading, the strength properties of PVD-coated materials are dependent on
the physicomechanical characteristics of the base material, coating, its thickness, stress and plastic strain levels,
number of loading cycles, etc.
The present experimental investigations were carried out to evaluate the effect of several PVD coatings and
their thickness on the tensile strain resistance of Kh18N10T stainless steel and VT1 and VT20 titanium alloy
specimens. Proceeding from the results of tensile tests, one of the materials, viz a VT20 alloy with a 6-mm TiN PVD
coating, was chosen to establish the basic relationships for the effect of PVD coatings on the low-cycle loading
resistance. The same material was used in experimental studies to evaluate the effect of cyclic loading at different
stress levels, followed by the application of PVD coatings, on the low-cycle loading resistance. For determining the
effect of PVD coatings on damage accumulation in tension and under cyclic loading, a known hardness method was
employed [19]. The scatter in hardness (homogeneity) data, obtained from the test (measurement) results, was taken
as the damage parameter.
Experimental Procedure. Monotonic tensile and low-cycle fatigue tests were performed on flat specimens
cut from Kh18N10T stainless steel and VT1 and VT20 titanium alloy sheets 1.5 and 1.0 mm thick, respectively
(Fig. 1).
The PVD coatings applied to Kh18N10T stainless steel specimens were: TiN 6, 10, 16.5 mm, Cr 4 mm, and
(Cr+Ni) 10 mm thick, the coatings on VT1 titanium alloy specimens were: TiN 6, 10, 16.5 mm thick, and on VT20
titanium alloy specimens they were: TiN 2, 4, 6, 8, 10 mm, (TiC)N 2, 4, 8, 10 mm, and (TiAl)N 2, 4, 6, 8 mm thick.
Before the application of PVD coatings, the specimens were cleaned in an alcohol-benzine mixture. Then
they were placed in the vacuum chamber, which was evacuated down to a 1. 33 × 10 -3 Pa air pressure and filled with
nitrogen up to 2.66 Pa, after that they went trough the cleaning in glow discharge for 15–25 min (600–1200 V,
0.5–1.0 A). Then the chamber was evacuated down to a 1. 33 × 10 -2 Pa nitrogen pressure and filled with argon up to a
6. 66 × 10 -2 Pa pressure, followed by switching the arc evaporator (titanium cathode, 100 A, 30–32 V). When voltage
is applied to the specimen (1100–1300 V, 24 A), ionic etching of the surface is accomplished. The time is 15–25 min
at 510–530°C. After cleaning the coating deposition was started.
The TiN coating was deposited at a pressure of 0.43–0.46 Pa in nitrogen. The current of the are evaporator is
85–95 A, voltage is 30–32 V, the voltage on the specimen is 350–550 V, the current is 3–6 A, the deposition

605
temperature is 510–530°C. The coating thickness was dependent on the deposition time, e.g., it is 40 min to get the
coating 6 mm thick.
After the coating deposition, the specimens with the chamber were cooled down to room temperature (about
1 h at 1. 33 × 10 -3 Pa).
Monotonic tensile and low-cycle fatigue tests were performed in a 3201UE20 standard electrohydraulic
machine at room temperature in air. The specimens were fixed with hydraulically operated grips. Axial forces at the
moment of free specimen end fixing did not exceed 1 kN. The coaxiality of force application was provided in setting
up the load-bearing frame. Additional stresses, arising from bending moments on specimen clamping, did not exceed
1% maximum axial ones. Axial elastoplastic strains of the specimen test portion were measured with a strain
gauge.
Monotonic tension of the specimens was effected at a travel speed of the active grip of ~ 5 mm/min, which
corresponds to a loading rate of 30 m/s (GOST 1497–84). Low-cycle fatigue tests were carried out in axial tension
with the frequency f » 0. 3 Hz. Triangular-cycle loading was conducted under the control of maximum and minimum
stress levels (“soft” loading), with the stress ratio R s » 0.060–0.075. Minimum cyclic stresses were kept constant,
viz 70 MPa (3.5–4.2% maximum stresses). Such loading conditions give rise to large elastoplastic strains, including
cyclic creep. Accumulated cyclic creep strains can cause quasistatic or fatigue fracture.
The three sets of specimens were tested: initial, with the PVD coating 6 mm thick, also after tension or
low-cycle loading.
Monotonic tension of initial VT20 titanium alloy specimens was effected up to 0.4% strains and cyclic
loading up to 53–86% of the number of cycles to fracture. After tension, the specimens were coated with TiN PVD
6 mm thick and put through tension to fracture. After cyclic loading, the specimens were coated with TiN PVD also
6 mm thick, then cyclic loading was continued to fracture at the same cyclic stresses. Low-cycle fatigue tests were
run to failure.
The damage distribution over the TiN-coated specimen surface in tension and under cyclic loading was
determined by a known method of evaluating the “degradation” of the structural state of materials [19]. The scatter in
hardness data was taken as the damage parameter.
Microcracks are often considered the onset of carrying capacity losses of the structure, its residual level is
monitored by parameters characterizing degradation, i.e., the material damage level. The density of bulk and surface
defects, average distances among them, etc. are taken as the simplest damage measures. Internal friction, the elastic
modulus defect or the difference between the moduli in tension and compression, the sound velocity, electrical
resistance, etc. belong to physical parameters correlating with damage.
Hardness is the most readily attainable and straightforward mechanical characteristic whose estimation does
not require either failure of the structure or preparing the specimens. Practically all mechanical properties correlate
with this characteristic to a greater or lesser extent [20]. Methods of estimating the static and dynamic hardness are
widely used for in-service diagnostics of the metal condition in pipe-lines, pressure vessels, carrying metallic
structures, etc. However, despite their obvious merits (convenience and ease of instrument use, especially in case of
dynamic hardness meters), the classical version of the hardness method exhibits low sensitivity to many of the
structural transformations, i.e., to the damage level.
As special experiments demonstrated, a more reliable estimation of the structural state of the material should
be based not on absolute values of a measured physicomechanical characteristic but on their certain derivatives, in
particular the scatter in the data measured on equal specimens under identical conditions [21]. The scatter in the data,
including those of hardness, is inherent in all materials, and their scattering is strongly dependent on their structural
state. Thus, the change in the structural state, i.e., the degradation of the material due to operation under certain
conditions, can be judged by the scatter in the data on their mechanical characteristics.
The homogeneity parameter m characterizes integrally the structural state of the material on processing the
hardness data. The structural inhomogeneity of the material can be evaluated only by statistical methods based on
physically substantiated distribution laws if a sufficiently large body of data on its properties is available. Evidently,
the normal distribution law is unacceptable in this case since it also extends to negative hardness values. At the same

606
time, there is a wealth of experience of using the Weibull distribution in mechanics of materials, particularly in the
construction of the statistical theory of strength [22]

P (s ) = 1 - e - ( s / k ) ,
m
(1)

where m and k are the distribution parameters, m indicates the scattering amplitude for the data on a characteristic
under study, which is determined by the Gumbel formula [23, 24], taking on the form for hardness tests

d ( n)
m= . (2)
2. 30259S (log H )

The d ( n) value is defined from n measurements, with n being no less than 15, and S (log H ) is obtained
from n measurement results
1 n
S (log H ) = å
n -1 i =1
(log H i - log H ) 2 , (3)

1 n
where log H is the mean hardness logarithm value, log H = å log H i .
n i =1
Large m values are consistent with a narrow scatter in the hardness data and, thus, with a higher-ordered
structure and a low damage level, while its smaller values correspond to a high damage level.
The processing of experimental results also made use of the coefficient of variation n, showing the scatter
in the data relative to the mean value [21]

-1/ 2
1 é 1 n ù
n = ê- å
H êë n - 1 i = 1
(H i - H ) 2 ú , (4)
úû

where H i is the hardness in an ith measurement, H is the mean hardness. With larger inhomogeneity, the parameter
values increase.
The damage distribution along the specimen after tension and cyclic loading was evaluated on initial and
TiN-coated specimens, also after cyclic loading. Hardness was determined over the surface in the unloaded specimen
portions (grips), in the initial part of the test portion, between the initial part and the fracture zone, and 5–7 mm from
the fracture edge (fracture zone). Hardness was measured by a Computest-ST hardness meter (Ernst, Switzerland). In
each specimen zone, 35 measurements were carried out.
Results and Discussion. Mechanical Characteristics. The effect of coating thicknesses on the strength in
monotonic tension was experimentally investigated. Data in Table 1 and Fig. 2 demonstrate the effect of ionic
etching and further coating on the yield limit s 0. 2 and tensile strength s u of the material. Thus, ionic etching
makes a major contribution to an increase in the yield limit (12.3 and 25.5% for a VT1 titanium alloy and Kh18N10T
stainless steel, respectively). For examined materials and PVD coatings the yield limit increased from 13.1 to 32.0%
at coating thicknesses of 4–10 mm.
Notice that the application of PVD coatings, also after tension up to 0.4% strain, gives rise to a higher yield
limit, which results in narrowing the stress range of strain hardening s u - s 0. 2 (Table 2, Fig. 3). As is seen, VT1 is
most prone and VT20 is least inclined to strain hardening.
Fatigue Strength and Lifetime. Low-cycle fatigue test results for VT20 titanium alloy specimens in the initial
condition and after application of a 6-mm TiN coating, also after cyclic loading, are shown in Fig. 4. At maximum
cyclic stresses of about 1075 MPa, fatigue curves possess a characteristic kink, caused by the change from quasistatic
to fatigue fracture.

607
TABLE 1. Mechanical Characteristics of Examined Materials in Monotonic Tension
Material, Initial Ionic etching Coating thickness, mm
coating condition (0.25 mm)
2.0 4.0 6.0 8.0 10.0 16.5
VT1 243.0 273.0 287.0 290.0 276.0
– – –
(TiN) 693.6 696.7 688.0 683.0 680.0
Kh18N10T 535.0 6715
. 685.0 706.2 670.0
– – –
(TiN) 949.8 9981. 1026.0 1045.4 1012.0
Kh18N10T 535.0 646.3
– – – – – –
(Cr) 949.8 967.7
Kh18N10T 535.0 693.6
– – – – – –
(Cr+TiN) 949.8 10181
.
VT20 977.0 1097.2 1100.4 1108.3 1108.2 1105.2
– –
(TiN) 1114.2 1095.3 1192.0 11917
. 1183.8 1189.3
VT20 977.0 11521. 1159.8 11511
. 11621.
– – –
(TiC)N 1114.2 11719
. 1169.8 1177.9 11819
.
VT20 977.0 1062.4 10891
. 1105.2 1103.9
– – –
(TiAl)N 1114.2 1130.4 1150.2 1155.7 11816
.

Note. Data for s 0. 2 and for s u are cited over and under the line, respectively.

TABLE 2. Mechanical Characteristics of a VT20 Titanium Alloy


Material su , s0. 2 , su - s0. 2 , d, % y, %
MPa MPa MPa
Initial 1114.2 977.0 137.2 11.50 18.63
TiN coating 1192.0 1108.3 83.7 9.33 13.89
Strain 0.4% + TiN 1309.0 1258.9 50.1 8.00 11.27

s0. 2 , MPa su , MPa

a b
Fig. 2. Yield limit (a) and ultimate strength (b) vs PVD coating thickness for examined materials: (1) VT1
(TiN); (2) VT1 (ionic etching); (3) Kh18N10T (TiN); (4) Kh18N10T (ionic etching); (5) Kh18N10T (Cr);
(6) Kh18N10T (Cr+TiN); (7) VT20 (TiN); (8) VT20 (TiC)N; (9) VT20 (TiAl)N.

Photographs of damaged specimens typical of low-cycle quasistatic and fatigue fracture ranges are presented
in Fig. 5. The application of a 6-mm TiN coating enhances the resistance to low-cycle loading as compared to the

608
su - s0. 2 , MPa
s max , MPa

N f , cycles

Fig. 3 Fig. 4
Fig. 3. Strain hardening vs coating thickness. (Designations are the same as in Fig. 2).
Fig. 4. Low-cycle fatigue curves for a VT20 titanium alloy in the initial condition (1),
after 6-mm TiN coating application (2), after cyclic loading (3), and loading + TiN (4).
(Open symbols – quasistatic fracture and solid symbols – fatigue fracture.)

a b
Fig. 5. General view of the specimen test portion fracture for a VT20 alloy with a TiN coating
after low-cycle fatigue tests: (a) low-cycle quasistatic fracture and (b) low-cycle fatigue fracture.

initial condition of a VT20 alloy in the range of low-cycle quasistatic fracture and does not exert great influence on
cyclic life in the range of low-cycle fatigue fracture (experimental points are within the natural scatter). The
application of a 6-mm TiN coating after cyclic loading to 53–86% of the number of cycles to fracture increases cyclic
lifetime 1.5–3.0 times as compared to the initial condition in the range of low-cycle fatigue fracture at the narrow
scatter of experimental data. Such a combination of cyclic loading and TiN coating application results only in fatigue
fracture (range of quasistatic fracture is absent on the fatigue curve).
Longer lifetime of TiN-coated VT20 specimens after cyclic loading can be explained by a hardening effect
of residual compression stresses arising in the coating precess, which modifies the cyclic properties of the material
and cyclic elastoplastic strain kinetics (Fig. 6). The initial titanium alloy specimen under cyclic loading exhibits
properties typical of a cyclically hardened material: smaller elastoplastic strain, then its jump, caused by compression

609
s max s 0. 2

N f , cycles
Fig. 6 Fig. 7
Fig. 6. Total elastoplastic strain kinetics per cycle for a VT20 titanium alloy: open symbols – cyclic loading of
the initial specimen, solid symbols – cyclic loading after TiN coating application at the same stresses as
preloading [(1) s =1000 MPa; (2) s =1040 MPa; (3) s =1060 MPa; (4) s =1080 MPa; (5) s =1090 MPa].
Fig. 7. Low-cycle fatigue curves for a VT20 titanium alloy in the reduced coordinates: maximum stresses in a
cycle/yield limit vs number of cycles to fracture: s 0. 2 = 977 MPa for initial specimens and s 0. 2 = 1108. 3 MPa
for TiN-coated specimens. (Designations are the same as in Fig. 4.)

stresses as a result of coating application. Further cyclic loading changes the properties of the material that becomes
cyclically softened (growth of elastoplastic strains).
If maximum cyclic stresses on low-cycle fatigue curves (Fig. 4) are related to the yield limit s max s 0. 2
(Fig. 7) for the material in the initial condition and TiN-coated (relative load level), the mutual arrangement of
low-cycle fatigue curves would change. Thus, over the examined range of lifetimes for a VT20 titanium alloy in the
initial condition, maximum cyclic stresses exceed yield limit in contrast to those of the TiN-coated material, also
after cyclic loading, when the stress is lower than the yield limit. Coating application enhances the yield limit and
cyclic lifetime with a decrease in a relative load level.
The curve of cyclic lifetime of a VT20 titanium alloy in the initial condition and TiN-coated, also after
cyclic loading, vs cyclic creep strains accumulated to fracture e f (Fig. 8), can also display quasistatic and fatigue
fracture zones. In the present investigation, fracture is taken to be quasistatic and fatigue at e f above and below 4%,
respectively.
The effect of TiN coating application, also after cyclic loading, on the relation between cyclic creep strains
accumulated to fracture and maximum cyclic stresses is illustrated by Fig. 9. The accumulation of cyclic creep strains
to fracture over 4% is observed at maximum cyclic stresses above 1075 MPa, which corresponds to the range of
quasistatic fracture (Fig. 4). The run of the curves shows that at fatigue fracture e f asymptotically approaches zero
with a decrease in maximum cyclic stresses, while at quasistatic fracture it tends to limiting values of initial plasticity
d, determined in tensile tests, with an increase in maximum stress levels (Table 2).
Effect of PVD Coatings on Damage Distribution in Tension and under Cyclic Loading. The damage
distribution pattern was investigated with VT20 titanium alloy specimens. Measurement results for the distribution of
hardness HRC over the surface of loaded and unloaded specimen portions are presented in Fig. 10 as mean hardness
values from 35 measurements in each examined zone: in the initial condition, after TiN coating application, and
cyclic loading up to 53–86% of the number of cycles to fracture. The damage distribution under cyclic loading was
investigated, choosing a specimen fracture pattern within quasistatic and fatigue ranges. The results point to low
sensitivity of a hardness-based diagnostics method to structural transformations, i.e., the damage level of the
material. The data on initial specimen hardness variations in tension and cyclic loading data at quasistatic and fatigue
fracture for initial specimens and for TiN-coated ones as well as after cyclic loading followed by TiN coating

610
ef ,% s max , MPa

N f , cycles ef ,%
Fig. 8 Fig. 9
Fig. 8. Cyclic life vs cyclic creep strains to fracture for a VT20 titanium alloy. (Designations are the
same as in Fig. 4.)
Fig. 9. Cyclic creep strains to fracture vs maximum cyclic stresses for a VT20 titanium alloy.
(Designations are the same as in Fig. 4.)

Fig. 10. Hardness distribution along the specimen in the initial condition (1–3), after TiN coating
application (4, 5), and cyclic loading + TiN (6, 7) under different conditions (open symbols –
quasistatic fracture, solid symbols – fatigue fracture): tension (1); cyclic loading: s max =1098 MPa (2);
s max =1034 MPa (3); s max =1070 MPa (4); s max = 980 MPa (5); s max =1090 MPa (6); s max =1000
MPa (7).

application show that the hardness of examined materials under the above conditions varies differently: it may
increase or decrease (Fig. 10).
The hardness distribution over the surface of the loaded specimen portion is dependent on the test conditions
(tension or cyclic loading), TiN coating, and loading history (Fig. 10). Test results suggest that in tension and under
cyclic loading of initial specimens, the hardness of the test portion increases slightly, however, as the fracture zone is
approached, it decreases (curves 1–3 in Fig. 10). Hardness along the TiN-coated specimen under cyclic loading
changes to a greater extent than that of the initial specimen, while loading followed by TiN coating application gives
rise to higher hardness values in the test portion of the specimen.

611
Fig. 11 Fig. 12
Fig. 11. Distribution of the homogeneity parameter along the specimen in the initial condition, after TiN
coating application, and after cyclic loading under different conditions. (Designations are the same as in
Fig. 10.)
Fig. 12. Distribution of the variance parameter along the specimen in the initial condition, after TiN coating
application, and cyclic loading under different conditions. (Designations are the same as in Fig. 10.)

The damage distribution over the specimen surface evaluated by the homogeneity and variance parameters
(scatter in the hardness data) is demonstrated in Figs. 11 and 12.
The distribution of the homogeneity parameter along the specimen (gripping zone, initial part of the test
portion, center of the test portion, between its initial part and fracture zone, area adjacent to the fracture zone,
possessing the highest damage level) is illustrated by Fig. 11. The damage level is dependent on the test conditions
(tension or cyclic loading), coating, and stress levels, contributing to a certain fracture behavior. Notice that TiN
coating application reduces the damage level on the specimen surface. Loading followed by TiN coating application
leads to a homogeneity parameter 40–42% higher, i.e., to a lower damage level in the area adjacent to the fracture
zone, as compared to the same parameter in the unloaded specimen portion, while the hardness variation is no more
than 5%.
Variations of the homogeneity parameter are indicative of the fact that the damage accumulation is also
influenced by maximum cyclic stress levels. At their large values (quasistatic fracture range), this process is much
more intensive, though the difference in maximum stress levels for other specimens, examined in the present study,
was 8–9%. The distribution of the homogeneity parameter along the specimen is noteworthy. The fatigue fracture of
cyclically-loaded initial specimens leads to a decrease in the homogeneity parameter in the fracture zone by 33.7% as
compared to their unloaded portion, while at the quasistatic one it is 56%. For comparison: in tension such a decrease
is 52%. It suggests that at fatigue fracture, the material in the fracture zone is less damaged than at quasistatic one,
thus, the damage accumulation rate is lower. Irrespective of a fracture pattern, TiN coating application, also after
cyclic loading, makes the homogeneity parameter in the fracture zone 38.5–41.5% smaller as compared to the
unloaded specimen portion, i.e., coating reduces the damage accumulation rate.
The distribution of the variance parameter (scatter in the hardness data) along the initial specimen, after TiN
coating application, after cyclic loading under different conditions and stress levels is shown in Fig. 12.
The results point to a wider scatter of the hardness data in the fracture zone irrespective of loading and
fracture patterns. Coating application after cyclic loading causes a slightly wider scatter of the hardness data in the
fracture zone.
It is of interest that the same damage accumulation pattern under cyclic loading was observed in low-cycle
tests of 10GN2MFA, 17GS, and 17G1S pipe steels [25], when under loading, hardness varies inconsiderably, and the
scatter in its data (homogeneity parameter) outside and in the fracture zone varies about twofold.

612
CONCLUSIONS

1. The effect of PVD coating application and its thickness on the strength and plasticity of examined
materials in tension was experimentally established. Ionic etching contributes essentially to an increase in the yield
limit. The PVD coating thicknesses of 6–10 mm provide maximum strength values of examined materials.
2. The TiN coating application on the initial material, also after tensile or cyclic elastoplastic strain, leads to
the strength growth both in tension and under cyclic loading. The PVD coating application after elastoplastic strain
in the low-cycle quasistatic fracture range results in the change from the quasistatic to fatigue fracture.
3. The TiN coating application after cyclic loading of the specimen changes the cyclic properties of the
material (cyclically hardened material becomes cyclically softened) and cyclic elastoplastic strain kinetics.
4. The distribution of the homogeneity and variance parameters along the specimen points to the effect of
test conditions (tension or cyclic loading), TiN coating, stress levels on the damage accumulation in the fracture
zone. The TiN coating application brings about a considerably lower damage level on the specimen surface.
The results obtained can be useful for the development of techniques to restore service properties of the
materials and to extend life of components damaged in operation, e.g., for maintenance or repair work.

REFERENCES

1. R. N. Pagborn, S. Weissmann, and J. R. Kramer, “Work hardening in the surface layer and in bulk during
fatigue,” Scr. Met., 12, No. 2, 129–131 (1978).
2. Y. Sato, H. Sasaki, and A. Kumana, “Surface layer yielding of low-carbon steel cylinders,” J. Mater. Sci.
Soc. Jap., 17, No. 3-4, 185–192 (1980).
3. S. Miyazaki, K. Shibata, and H. Fujita, “Effect of specimen thickness on mechanical properties of
polycrystalline aggregates with various grain sizes,” Acta Met., 27, No. 5, 855–862 (1979).
4. Yu. A. Kharlamov and N. A. Budag’yants, Physics, Chemistry, and Mechanics of the Solid Surface [in
Russian], VUGU, Lugansk (2000).
5. V. E. Ivanov, A. I. Somov, and M. A. Tikhonovskii, “Dislocation mechanism of the effect of hard surface
films on the strain and fracture of metals,” in: Protective High-Temperature Coatings [in Russian], Nauka,
Leningrad (1972), pp. 291–305.
6. R. R. Rykalin, M. Kh. Shorshorov, and V. P. Alekhin, “Environmental effects on plastic strain of crystals
(review),” in: Effect of the Physicochemical Medium on the High-Temperature Strength of Metallic
Materials [in Russian], Nauka, Moscow (1974).
7. E. S. Puchi-Cabrera, F. Martinez, I. Herrera, et al., “On the fatigue behavior of an AISI 316L stainless steel
coated with a PVD TiN deposit,” Surf. Coat. Techn., 182, 276–286 (2004).
8. K. R. Kim, C. M. Suhb, R. I. Murakami, and C. W. Chung, “Effect of intrinsic properties of ceramic coatings
on fatigue behaviour of Cr–Mo–V steels,” Surf. Coat. Techn., 171, 15–23 (2003).
9. J. A. Berrios, D. G. Teerb, and E. S. Puchi-Cabrera, “Fatigue properties of a 316L stainless steel coated with
different TiNx deposits,” Surf. Coat. Techn., 148, 179–190 (2001).
10. S. Hotta, Y. Itou, K. Saruki, and T. Arai, “Fatigue strength at a number of cycles of thin hard coated steels
with quench-hardened substrates,” Surf. Coat. Techn., 73, 5–13 (1995).
11. H. Oettel, R. Wiedemann, and S. Preibler, “Residual stresses in nitridå hard coatings prepared by magnetron
sputtering and arc evaporation,” Surf. Coat. Techn., 74-75, 273–278 (1995).
12. M. Gelfi, G. M. La Vecchia, N. Lecis, and S. Troglio, “Relationship between through-thickness residual
stress of CrN-PVD coatings and fatigue nucleation sites,” Surf. Coat. Techn., 192, 263–268 (2005).
13. S. Baragetti, G. M. La Vecchia, and A. Terranova, “Fatigue behaviour and FEM modelling of thin-coated
components,” Int. J. Fatigue, 25, 1229–1238 (2003).
14. M. P. Nascimento, R. C. Souza, W. L. Pigatin, and H. J. C. Voorwald, “Effects of surface treatments on the
fatigue strength of AISI 4340 aeronautical steel,” Int. J. Fatigue, 23, 607–618 (2001).
15. U. Wiklund, J. Gunnars, and S. Hogmark, “Influence of residual stresses on fracture and delamination of thin
hard coatings,” Wear, 232, 262–269 (1999).

613
16. C. J. Villalobos-Gutierrez, G. E. Gedler-Chacon, J. G. La Barbera-Sosa, et al., “Fatigue and corrosion-
fatigue behavior of an AA6063-T6 aluminum alloy coated with a WC–10Co–4Cr alloy deposited by HVOF
thermal spraying,” Surf. Coat. Techn., 202 (18), 4572–4577 (2008).
17. J. A. M. Ferreira, J. D. M. Costa, and V. Lapa, “Fatigue behaviour of 42CrMo4 steel with PVD coatings,”
Int. J. Fatigue, 19, 293–299 (1997).
18. A. Alsaran, I. Kaymaz, A. Celik, et. al., “A repair process for fatigue damage using plasma nitriding,” Surf.
Coat. Techn., 186, 333–338 (2004).
19. A. O. Lebedev, M. R. Muzyka, and N. L. Volchek, Method of Estimation of the Material Degradation as a
Result of In-Service Damage Accumulation. LM Hardness Method [in Ukrainian], Patent of Ukraine No.
52107A. Valid since January 15, 2003.
20. M. S. Drozd, Nondistructive Tests of the Mechanical Characteristics of a Metal [in Russian], Metallurgiya,
Moscow (1965), pp. 147–156.
21. A. A. Lebedev, N. R. Muzyka, and N. L. Volchek, “Determination of damage accumulated in structural
materials by the parameters of scatter of their hardness characteristics,” Strength Mater., 34, No. 4, 317–321
(2002).
22. W. Weibull, “A statistical distribution function of wide applicability,” J. Appl. Mech., 18, No. 3, 293–297
(1951).
23. E. J. Gumbel, Statistical Theory of Extreme Values and Some Practical Applications, National Bureau of
Standards, Washington (1954).
24. N. A. Makhutov, V. V. Zatsarinnyi, Zh. L. Bazaras, et al., Statistical Relationships of Low-Cycle Fracture
[in Russian], Nauka, Moscow (1989).
25. A. O. Lebedev and M. R. Muzyka, “Technical diagnostics of the material by the LM hardness method,” in:
Problems of Lifetime and Safe Operation of Constructions, Structures, and Machines [in Ukrainian], Paton
Institute of Electric Welding, National Academy of Sciences of Ukraine, Kyiv (2006), pp. 97–101.

614

S-ar putea să vă placă și