You are on page 1of 8

Materials Science, Vol. 36, No.

2, 2000

INFLUENCE OF INCOMPLETE ANNEALING OF TITANIUM ( ~ + ~)-ALLOY


AND ITS WELDED JOINTS ON FATIGUE RESISTANCE
AND CORROSION-FATIGUE RESISTANCE

O. S. Kalakhan and V. I. Pokhmurs'kyi UDC 620.193: 669.295.5:621.791.051

We investigate the influence of incomplete annealing (675~ 10 h) on fatigue, corrosion-fatigue, and


high-temperature (up to 500~ fatigue resistance of titanium (a + [~)-alloyTi-5A1-1.5V- 1Mo with
nominal composition as well as its welded joints made by the electron-beam method. The effect of
annealing depends on a totality of factors which dominate in different ways under fatigue fracture in air,
in a 3% solution of NaC1. and high-temperature fatigue. Possible mechanisms of the influence of in-
complete annealing on fracture of the titanium alloy and its welded joints are analyzed.

Two-phase high-strength heat-treated titanium alloys are used in numerous structures, in particular in welded
ones. Among the alloys of such a type, the most commonly used are T i - A 1 - V and T i - A 1 - M o - V [1, 2]. Their
welded joints are subjected to heat treatment for the removal of residual stresses and stabilization of the structure of
these joints. For this purpose, complete or incomplete annealing is used [2]. The latter (in furnaces with an air
atmosphere) partially realizes these functions. However, it does not require subsequent removal of the oxide film
and gas-saturated surface layer of metal and is simple for realization. Below, we present the results of study of the
influence of incomplete annealing, in particular gas-saturated layers, on fatigue and corrosion fatigue of the base
metal (the alloy including 5% A1, 1.5 V, 1% Mo) and of its welded joints at room temperature as well as on the
fatigue in the 2 0 - 5 0 0 ~ range.

Materials and Methods of Studies

Specimens (17 mm in thickness) were cut from plates parallel to the direction of rolling ( c u = 913 MPa, •o.2
= 882 MPa, ~ = 13%, V = 31%, KCV = 6 4 J / c m 2) and from forged metal with a square cross-section with
sizes 1 6 • 1 6 x 1 0 0 m m ( c u = 1010MPa, ~o.2 = 900MPa, 5 = 2%, V = 35%, KCV = 4 1 J / c m 2) [3].
Welded joints were obtained by electron-beam welding. We tested specimens made of sheet metal in air and in a
3% NaC1 solution, welded joints within the 20-500'(2 temperature range (series I), and specimens made of forged
metal with notch at the center of the weld in air and in a 3% NaC1 solution (series II).
Standard cylindrical specimens with the diameter of the working part of 5 m m with a V-shaped circular notch
(p = 0.5, c% = 1.99) were tested for pure bending with rotation [4]. The base of the experiments was 5 9 107
cycles of loading and it was 2 - 107 cycles in high-temperature studies. The specimens were annealed in the
following regime: heating in a furnace up to 675~ holding for I 0 h, and cooling with the furnace. The
dependences of the growth rate of a fatigue crack (v) on the amplitude of the stress intensity factor (AK) were
constructed according to the known procedure [5]. The conditional growth threshold of a fatigue crack was
estimated as the amplitude of AK for a rate equal to 10-10 m/cycle. Girder-type specimens With rectangular cross-
section (4 • 16 mm) and with a unilateral V-shaped notch 1.6 mm in depth and radius at its top p = 0.1 m m were
rigidly loaded at a frequency of 10Hz with a sign-preserving cycle in air and in a 3% sodium chloride. The depth
of the gas-saturated layer was studied on the basis of determination of the microstructure and microhardness of the
specimens. For statistical treatment of the results of the experiments, we used the method of linear regression [6].
The fatigue curves are presented for the 50% probability of failure.

Karpenko Physicomechanical Institute, Ukrainian Academy of Sciences, L'viv. Translated from Fizyko-Khimichna Mekhanika Materialiv,
Vol. 36, No. 2, pp, 76-82, March-April, 2000. Original article submitted April 9, 1999.

244 1068-820X/00/3602-0244
$25.00 9 2000 Kluwer Academic/PlenumPublishers
INFLUENCE OF INCOMPLETE ANNEALING OF TITANIUM ( ~ + [~)-ALLOY AND ITS WELDED JOINTS ON FATIGUE RESISTANCE 245

Fig. 1. Microstructure of a sheet forge-rolling (a), forged piece (b). welded joint (c), and heat-affected zone (d) of Ti-5AI-1.5V-1Mo
alloy before and after annealing a sheet forge-rolling (675~ 10 h, cooling in a furnace): (e) x200, (f) x500.

Results of Tests

The structure of a sheet alloy along the direction of rolling retains the former [3-grains with fuzzy boundaries
comprising small spherical grains. The arrangement of (x-plates inside the [3-grains resembles "basket wicker."
The average size of (x-colonies varies in the 2 0 - 5 0 g m range, and the thickness of (>plates is 2 p.m (Fig. la).
The forged alloy contains fine-plate toothed grains of irregular form. After recrystallization, high-angle
boundaries and chains of equilibrium grains arise in c~-plates (Fig. lb). For the forged alloy, greater limits of
endurance (by almost 100 MPa) and yield c0. 2 (by 18 MPa) are typically observed. The relative elongation 8
practically remains unchanged, and the relative narrowing ~g increases by 10%. The resilience of the alloy is
18 J/crn 2 KCV less than that for the sheet. In the metal of the weld of a welded joint, a freely dispersed martensite
(x'-phase arises in the small-grain initial matrix of [3-phase due to high rates of cooling (Fig. lc). After the electron-
beam welding, we observe significant porosity of the welded joint. The structure near the weld zone is the
martensite acicular (x'-phase (ct"-phase) against the background of the residual [3-phase, some part of which is in a
metastable state (Fig. ld).
After annealing, the structure of the base metal remained practically without any changes (Fig. 1e, f). Only the
number of small globules on the boundaries of former [3-grains elongated along the direction of forge-rolling
246 O.S. KALAKHANAND V. I. POKIIMURS'KYI

increased. The oxide films 3 - 1 2 gm in depth formed during the incomplete annealing of the alloy at a temperature
of 675~ adhered well to the surface. The base phase of oxide films is rutile TiO 2 [7]. As a result of oxygen
diffusion into the crystal lattice, a solid solution of oxygen in titanium is formed. The depth of the gas-saturated
zone is 6 0 g m , and the surface microhardness is 6.0-6.1 GPa when the microhardness of the bulk of the alloy is
4.8-4.9 GPa.

6
+l
460

380

300

220

140

60
104 105 106 107
N, cycle
Fig. 2. Fatigue strength of Ti-5AI-1.5V-1Mo alloy and of its welded joints in air (solid lines - - smooth specimens, dotted lines - -
specimens with concentrator: (1, I, 3, III) rolled specimens, (2. II, 4, IV) electron-beam welding, (3, 1:I/. 4, IV) annealing at
675~ 10 h.

Specimens made of forge-rolled sheets are durable (eLI = 470 MPa, which amounts to 0.51 cYu) but very
sensitive to the concentration of stresses (Fig. 2). The fatigue resistance of notched specimens decreases down to
150 MPa, and the effective stress Concentration factor is 3.1.
By estimating the operating reliability of the alloy, we have found the following features of the v-AK-curves
(Fig. 3, curve 1): great length of the section of average rates and absence of its abrupt fall at a rate of 10-l0 m/cycle.
The conditional threshold AKth is 3.3 MPa-,/~. In the alloy with the "basket wicker" structtfre, a crack grows with
formation of several holes (microcracks) at some distance from the tip of the crack, and then the material between
these holes and the main crack undergoes breakage. After merging of one of the holes with the main crack, the other
microcracks close.
For welded joints of the forged alloy as well as for the base metal, a great spread of their limited durability is
typically found. Smooth specimens undergo fracture mainly in the metal of a weld, and they undergo fracture far
more rarely in the zone of thermal influence. The most common defect of the metal of a weld of ((z + 13)-alloys is
porosity [8]. Irrespective of the type and regimes of welding as well as of an electrode wire near the welding line,
we observe pores with sizes from 0.05 to 0.8 mm distributed over the entire metal of the weld [9]. The durability
of welds is mostly affected not by the dimension of pores but by their arrangement. Studies of welded joints of
INFLUENCE OF INCOMPLETE ANNEALING OF TITANIUM ( ~ + ~)-ALLOY AND ITS WELDED JOINTS ON FATIGUE RESISTANCE 247

VT14 and VT22 (co + I~)-alloys as well as of Ti-6A1-4V alloy have shown that two to three pores with a diameter
of 0.6-1.0 mm in the central zone of the specimen decrease its durability by 2 to 3 times. Pores in the surface
layers decrease the durability by 5 to 10 times. Such a great effect of porosity on low-cycle fatigue is explained
not only by the concentration of stresses near a pore, but also by the action of hydrogen [9] which favors the
ori~nation of initial cracks in welds and promotes their development. It is found that the concentration of hydrogen
near pores is greater by one order of magnitude than that in other sections of metal. Due to porosity of the weld, the
fatigue resistance of smooth welded specimens is 70% of d_ 1 of the sheet metal. The fatigue resistance of welded
joints is also decreased by tensile stresses. The endurance limit for the specimens with a notch in the middle of the
weld is greater than that for the specimens of sheet metal, and it reaches 230 MPa (see Fig. 2). The ambiguity of the
experimental data concerning the effect of the surface gas-saturated layers on the strength of alloys [10-12] is
attributed to a great extent to temperature-kinetic distinctions of their formation and to their structural-phase
composition.

k.,I

E
>-

I07 2

I0-8
(

I0"9 e~e

3 4 5 6 7 Ax', MPa .q'~

Fig. 3. Kinetic diagrams of fatigue failure of T i - A I - I . 5 V - 1 M o alloy: (1) in air, (2) in a 3% NaC1 solution.

After annealing, the fatigue resistance of smooth specimens made of the base metal decreases by 20%, while
~-1 increases by 40% (210MPa against 150MPa) for those with a gas-saturated layer and with a notch [13]. In
welded specimens with a notch, fatigue resistance essentially worsens after annealing (Fig. 4). Their endurance
limits decrease 2 or 3 times (56% of the value without annealing), while the endurance limits decrease only by
15% (which corresponds to 0.1 and 0.32 of ~u) for smooth welded specimens with a gas-saturated layer. There-
fore, saturation of the surface with gas essentially decreases the fatigue resistance of the base metal of smooth
specimens and specimens with a notch, and the welded joints of the base metal in air.
Thermodiffusion saturation causes an increase in the corrosion resistance of titanium and its alloys in a number
of aggressive media [7, 14]. But whereas the decisive role there belongs to oxide films, in the corrosion-mechanical
failure, this role is played by gas-saturated layers and the structure of the base material. In a 3% solution of sodium
chloride, changes in the endurance limit of the nonannealed specimens of the base material and its welded joints are
insignificant (see Table 1).
248 O.S. KALAKHAN AND V, I. POKHMURS'KYI

6
+1
,1 O-
400

360

280

200
N , N;L,r
120
v\
IIIIIII1 I I I 1 ~ |ll|ltl~
"-,2'I l-.W*,-
|lltl~ lll~t|
40
104 105 106 107
At, cycle

Fig. 4. Corrosion fatigue of T i - 5 A I - 1 . 5 V - 1 M o alloy and its welded joints in a 3% NaCI solution (solid lines-smooth specimens.
dashed lines-specimens with concentrator): (1. I, 3, lJI) rolled products, (2. I/, 4, IV) electron-beam welding. (3, Ill, 4. IV) an-
nealingat 675~ 10h.

Table 1. Influence of Corrosive Medium and Annealing on Endurance of the Alloy

Annealing at 675~ 10 h
Air 3% NaC1
State of the alloy Specimens Air 3% NaC1

~7 1, MPa (~cor
- l , MPa o_ l, MPa cyc~ MPa

After rolling Smooth/with notch 470/150 450/140 370/210 240/190

After electron-beam welding Smooth/with notch 320/230 330/210 270/100 280/60

In a corrosive medium (Fig. 3, curve 2), the growth rate of a fatigue crack on the section of average rates (at
AK > 4.5 MPa-,/m) is appreciably increased. However, upon approaching the threshold values, it abruptly falls due
to intensive microbranching. The threshold AK_lo is higher in the medium than in air.
Annealing somewhat increases the endurance limit of smooth welded joints (Figs. 2 and 4, curves 4) but
decreases the endurance of the base metal for both types of specimens and for the welded joints with notch (see
Table 1). Such a behavior is probably caused by the decay of the martensite c( ( c(')-phase in the weld and in the
zone of thermal influence during thermal treating and by the precipitation of passive dispersed particles of the stable
c~- and [3-phases [15]. As compared with the base metal, this system has a limit of corrosion fatigue that is higher
by 40 MPa.
The abrupt (by 35%) fall of the endurance limit of the smooth specimens cut from forge-rolled sheets after
annealing (Figs. 2 and 4, curves 3) is caused not only by the increase in hardness of the surface but also by the decay
INFLUENCE OF INCOMPLETE ANNEALING OF TITANIUM ( r ~)-ALLOY AND ITS WELDED JOLTS ON FATIGUE RESISTANCE 249

of the (x-phase with formation of precipitations of the (xe-phase [10]. The nuclei of another phase with a
stoichiometric composition of the type of Ti 3 A1 or Ti 2 A1 and with a high electrochemical activity worsen the
corrosion endurance. This is favored also by oxide A12 03 in the surface layers which does not form any stable
compound with TiO2 and makes these layers less corrosion-resistant [ 16]. The molecular volume of this oxide is
twice as large as that in rutile. Therefore, its appearance can be accompanied by loosening of the oxide layer. Here,
as one can see from the presented data, the principal role is played by the medium.

tS_ /e
220 -

I
180

140 Ikx

100

60 i t i 1 i -- ,~
20 100 200 300 400 ~ ~

Fig. 5. Influence of temperature on fatigue resistance o f T i - 5 A i - 1 . 5 V - IMo alloy and its welded joints without and after annealing: (1)
base metal without annealing, (2) welded joints without annealing, (3) base metal, annealing, (4) welded joints, annealing.

High-Temperature Fatigue of Gas-Saturated Alloy and Its Welded Joints

Tests of titanium alloys under elevated temperatures showed [17] that heating appreciably changes their
properties. The authors estimated the influence of gas-saturated surface layers formed in advance during the
incomplete annealing of the alloy on the fatigue of specimens with a constructive concentrator of stresses of the base
metal and its welded joints within the 20-500~ temperature range. As a reference, the authors studied specimens
without previous annealing. It was found (Fig. 5) that the fatigue resistance of the alloy and its welded joints
decreases with increase in temperature; however, the endurance limits of welded joints with a notch in the middle of
the weld are higher as compared with that of the same specimens made of the base metal. This is probably caused
by the fact that the processes of decay of the (x' ((x')-phase with formation of passive dispersed precipitations of the
stable (x- and ~-phases prevail in welded joints during annealing [15].
At 20-500~ temperatures, the annealed welded joints have a higher fatigue limit than the annealed base
metal. On the whole, however, these values are substantially lower than that for specimens without annealing
(Fig. 5, curves 1 and 2). That is, if high-temperature fatigue is important for further operation of welded joints of
T i - 5 A I - 1 . 5 V - 1 M o alloy, it is not expedient to perform preannealing for the removal of residual stresses in
electron-beam welds.
As to the possible causes of the ambiguous influence of incomplete annealing on the fatigue, corrosion-fatigue,
and high-temperature fatigue failure of the (~x + [3)-titanium alloy, we indicate a number of factors which diversely
influence the failure. As a result of preannealing and slow cooling, the (x-phase decays with the formation of
reprecipitations of the tea-phase enriched with aluminum (microzones with the stoichiometric composition of Ti 3 A1
or Ti 2 A1) [10]. Oxygen can also cause precipitations of the ordered cc2-phase earlier than predicted by the state
250 O . S . KALAKHAN AND V. I. POKHMURS'KY~

diagram of the Ti-A1 system [18]. Therefore, one of the causes of a decrease in high-temperature resistance and
corrosion-fatigue resistance (the abrupt fall in the fatigue resistance of smooth specimens of the base metal down to
240 MPa in the medium as compared with 370 MPa in air) can be the reprecipitation and precipitation of the or-
dered ~2-phase.
Oxygen makes the structural inhomogeneity of the surface layer (where A1203 oxide, which do not form
stable compound with TiO 2 , can be present) more pronounced and makes this layer less corrosion-resistant [16].
Oxygen only slightly influences the solubility of hydrogen in titanium at room temperature, but noticeably increases
it at elevated temperatures [21]. From 400~ and higher, titanium absorbs hydrogen with the formation of hydrides
with stoichiometric composition TiHo.83 [20]. Therefore, after cooling of the alloy, all hydrogen will be fixed in
titanium hydride. The density of hydride is essentially less than that of titanium. Due to the gradient of the
concentration of hydrogen, the amount of precipitated hydrides decreases with distance from the surface of the base
metal, and they are localized in the zone of metal with increased stresses near pores in welded joints. In the zones of
the diffusion layer which are richest in oxygen, the precipitatio n of hydride is more intensive [19]. At temperatures
below 300~ (the temperature of beginning of precipitation of hydride), due to the small plasticity of titanium
saturated with oxygen, the possibility of removal of the stresses caused by structural changes and bulk changes in
the surface layer by means of plastic deformation is very limited. Therefore, both during annealing and under long-
term action of high temperatures, brittle films which are not capable of protecting the alloy against intensive
saturation with gases can be formed on the surface. Therefore, along with the structural transformations mentioned
above, the surface embrittlement of the alloy is an important negative factor.
The increase in the endurance limit of the previously annealed alloy at temperatures from 350~ to the room
temperature (Fig. 5) is caused, in our opinion, by dispersion hardening of the alloy. As was stated, a high de~ee of
dispersion of microregions of the me-phase after annealing does not lead to an increase in the strength of the alloy at
low testing temperatures. Only at 350~ does the activation energy become sufficient for growth of nuclei of the
a2-phase to discrete particles which appreciably strengthen the alloy. With further increase in temperature up to
500~ the a2-phase grows rapidly, the alloy becomes "overaged,'" and large particles of the phase enriched with
aluminum increase its structural inhomogeneity. Thus, annealing of the titanium (c~ + [3)-alloy Ti-5AI-1.5V-1Mo
gives rise to the action of various factors, and the effect depends on which of them dominates during fatigue,
corrosion-fatigue, and high-temperature-fatigue failure of smooth specimens, specimens with a notch, and welds.

CONCLUSIONS

Incomplete annealing for the removal of welding residual stresses (heating up to 600-700~ in air, slow
cooling) embrittles the (c~ + 13)-alloy T i - 5 A I - I . 5 V - 1 M o . As a result, the endurance limit for welded joints with
notch decreases by 60% as compared with its value before annealing, for smooth joints by 15%, and for the base
metal by 20%.
In the medium, the influence is diverse. The corrosion-fatigue resistance of the base metal of specimens of two
types decreases abruptly (by up to 40%), which is connected with the formation of reprecipitations of the e~z-phase
in the structure and of oxide A1z 03 in the surface layers. For the welded joint-medium system, annealing does not
worsen its corrosion-fatigue resistance due to the decay of the martensite a ' (a")-phase in the metal of weld and in
the zone of thermal influence with the precipitation of passive dispersed particles of the stable a- and I]-phases.
It is not expedient to perform the preannealing of the alloy and its welded joints which operate at high
temperatures (up to 500~

REFERENCES

1. S.M. Gurevich(editor), ReferenceBook on Welding of Nonferrous Metals [in Russian], NaukovaDumka, Kiev (1990).
2. V.N. Moiseev,F. R. Kulikov,Yu. G. Kirillov,et al., WeldedJoints of Titanium Alloys [in Russian], Metallur~ya,Moscow (1979).
3. O.S. Kalaldaan and N. A. Prevars'ka, "Durabilityof Ti-5AI-I.5V-1Mo type alloy of varyingstructure and its weldedjoints in the
presence of the c~-layerand medium," Fiz.-Khim. Mekh. Mater., 29, No. 6, 115-117 (1993).
INFLUENCE OF INCOMPLETE ANNEALING OF TITANIUM (C~+ ~)-ALLOY AND ITS WELDED JOINTS ON FATIGUE RESISTANCE 251

4 V.I. Pokhmurskii, CorrosionFatigue of Metals [in Russian], Metallurgiya. Moscow (1985).


5 S. Ya. Yarema, "Method for determining the growth rate of cracks and the characteristics of crack resistance under cyclic load,"
Fiz.-Khim. Mekh. Mater.. 30, No. 3, 137-152 (1994); 30, No. 4, 121-136 (1994); 30. No. 6, 101-I 12 (1994); 31, No. 1, 145-157
(1995).
6. B. Ya. Oleksiv, M. I. Frankiv, and O. S. Kalakhan, Regression Analysis in the Study of Fatigue Strength of Metals. Processing of
Experimental Data [in Ukrainian], Preprint No. 27, Karpenko Physicomechanical Institute, Ukrainian Academy of Sciences, L'viv
(1980), pp. 11-14.
7 V.M. Fedirko, O. S. Kalakhan, and I. M. Pogrelyuk, "Influence of the structure of the surface layers of nitrided titanium on its
corrosion-electrochemical properties in acid solutions," Fi.;.-Khim.Mekh. Mater.. 31. No. 6, 6%72 (1995).
8 V.I. Trufyakov (editor), Endurance of Welded Joints under Varying Loads [in Russian]. Naukova Dumka, Kiev (1990).
9 S.I. Kishkina, F. R. Kulikov, L. A. Stronina, et al., "Low-cycle fatigue of welded joints of ( ct + [5) titanium alloys." in: Titanium.
Physical Metallurgy and Technology [in Russian], All-Union Institute of Casting and Alloys, Moscow (1978), Vol. 2. pp. 273-282.
10 B.B. Chechulin and Yu. D. Khesin, Cyclic and Corrosion Durabili~ of Titanium Alloys [in Russian], Metallurgiya, Moscow
(1987).
11 B.A. Drozdovskii, L. V. Prokhodtseva, and N. I. Novosil'tseva, Crack Resistance of Titanium Alloys [in Russian]. Metallurgiya,
Moscow (1983).
12 G.G. Maksimovich. V. N. Fedirko, and A. T. Pichugin, "Influence of thetemperature of annealing in air on the strength properties of
titanium alloys." Fi=.-Khim.Mekh. Mater.. 16, No. 5, 85-87 (1980).
13 O.S. Kalakhan, "Durability of welded joints made of Ti-5 AI-1.5 V-1 Mo titanium alloy of different structures after annealing," in:
Abstracts of 2nd Intern. Syrup. on Mis-Matching of Welds (April 24--26, 1996. Reinstirf-Liineburg, Germanyl (1996).
14 V.M. Fedirko, I.M. Pohreliuk, and O. S. Kalakhan, "Corrosion behavior of surface titanium layers in acid solutions after nitrogen
thermodiffusion saturation," WerkstoffeKorrosion, 49, No. 6, 435-439 (1998).
15 V.N. Moiseev, F. R. Kulikov, Yu. G. Kirillov, et al., WeldedJoints of Titanium Alloys [in Russian], Metallurgiya, Moscow (1979).
16 I.I. Komilov, N. G. Boriskina, E. M. Kenina, and M. N. Zabrodskaya, "Influence of long-term oxidation on corrosion resistance and
mechanical properties of AT3 and AT6 alloys." in: Titanium. Physical Metallurgy and Technology [in Russian], All-Union Institute
of Casting and Alloys, Moscow (1978). Vol. 2. pp. 85-92.
17 O.S. Kalakhan, N. Ya. Yaremchenko. and N. A. Prevarskaya, "'Durability of titanium PT3V alloy at elevated temperatures," Fiz.-
Khim. Mekh. Mater., 21, No. 5, %11 (1985).
18 V.V. Vavilova. "Influence of oxygen on the properties of titanium and its alloys.'" Metalloved. Term. Obrab. Met.. No. 10. 10-12
(1973).
19 B.A. Kolachev, Hydrogen Brittleness of Nonferrous Metals [in Russian], Metailurgiya. Moscow (1979).
20 V.A. Kachuk and M. B. Svetlov, "Study of certain properties of VT5P alloy with additions of rare-earth metals," in: Titanium.
Physical Metallurgy aJu4Technology [in Russian], All-Union Institute of Casting and Alloys, Moscow (1978), Vol. 3, pp. 367-373.

You might also like