STP 612-1976
STP 612-1976
MATERIALS AND
COMPONENTS
A symposium
sponsored by ASTM
Committee E-9 on Thermal Fatigue of
Materials and Components
AMERICAN SOCIETY FOR
TESTING AND MATERIALS
Philadelphia, Pa., 17-18 Nov. 1975
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
© by AMERICAN SOCIETY FOR TESTING AND MATERIALS 1976
Library of Congress Catalog Card Number: 76-21535
NOTE
The Society is not responsible, as a body,
for the statements and opinions
advanced in this publication.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
Foreword
The Symposium on Thermal Fatigue of Materials and Components was
presented at the meeting held in New Orleans, 17-18 Nov. 1975. The
symposium was sponsored by The American Society for Testing and
Materials through its Committee E-9 on Fatigue. D. A. Spera, National
Aeronautics and Space Administration Lewis Research Center, and D. F.
Mowbray, General Electric Co., presided as symposium co-chairmen.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
Related
ASTM Publications
Fatigue Crack Growth Under Spectrum Loads, STP 595 (1976), $34.50
(04-595000-30)
Manual on Statistical Planning and Analysis for Fatigue Experiments,
STP 588 (1976), $15.00 (04-588000-30)
Handbook on Fatigue Testing, STP 566 (1974), $17.25 (04-566000-30)
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
A Note of Appreciation
to Reviewers
This publication is made possible by the authors and, also, the un-
heralded efforts of the reviewers. This body of technical experts whose
dedication, sacrifice of time and effort, and collective wisdom in review-
ing the papers must be acknowledged. The quality level of ASTM publica-
tions is a direct function of their respected opinions. On behalf of ASTM
we acknowledge with appreciation their contribution.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
Editorial Staff
Jane B. Wheeler, Managing Editor
Helen M. Hoersch, Associate Editor
Ellen J. McGlinchey, Assistant Editor
Kathleen P. Turner, Assistant Editor
Sheila G. Pulver, Editorial Assistant
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
Contents
Introduction
What is Thermal Fatigue?— D. A. SPERA 3
Nonlinear Analysis of a Tapered Disk Thermal Fatigue Specimen—
D. F. MOWBRAY AND J. E. MCCONNELEE 10
Summary 255
Index 260
Copyright by
Downloaded/printed by
University of
STP612-EB/NOV.1976
Introduction
D. F. Mowbray
Mechanics of Materials Unit, Materials and
Processes Laboratory, General Electric
Company, Schenectady, N.Y.; symposium
cochairman and also coeditor of this
publication.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
D. A. Spera^
ABSTRACT: Definitions are suggested for terms such as thermal fatigue, thermal-
mechanical fatigue, and thermal-stress fatigue. Historical developments in the field of
thermal fatigue from 1838 to the present are reviewed. It is hypothesized that
we have advanced from qualitative to quantitative understanding of the thermal
fatigue process and can now make life predictions with factor-of-two reliability
for conventional metals.
The purpose of the symposium which produced this volume was to ex-
change ideas on methods for determining and improving the resistance of
materials and mechanical components to that combination of failure
mechanisms which we know as thermal fatigue. To various people at
various times, thermal fatigue is also known as thermal stress fatigue,
thermal strain fatigue, creep fatigue, thermal cracking, thermal shock,
thermal rupture, thermal endurance, low-cycle thermal fatigue, heat
checking, craze cracking, fire cracking, and just plain high-temperature
fatigue. Considering this assortment of names, it is appropriate for us to
discuss the terminology we will need in order to communicate effectively
with one another on this complicated subject.
First, a few simple definitions will be suggested which may be useful,
particularly to those who have had Uttle contact with the study of thermal
fatigue. If these definitions help us, we may wish to continue using them.
If not, let us try to improve them. I believe they are relevant to present
studies of thermal fatigue phenomena. And what is just as important,
they contain a certain measure of respect for the historical development
of our field. After giving the definitions, I will briefly review the de-
velopment of terms during the past 137 years of our technical history.
' Consultant, Fatigue Branch, Lewis Research Center, Cleveland, Ohio 44135.
Suggested Definitions
First, let me suggest this definition for thermal fatigue: "Thermal
fatigue is the gradual deterioration and eventual cracking of a material
by alternate heating and cooling during which free thermal expansion is
partially or completely constrained." Constraint of thermal expansion
causes thermal stresses which may eventually initiate and propagate
fatigue cracks. Thermal fatigue may be classified under the more general
heading of low-cycle fatigue, because thermal fatigue cracks usually start
in less than 50 000 cycles. In addition, a thermal fatigue cycle usually con-
tains significant inelastic strain. Thus, I suggest we start defining our
terms by dividing the subject of low-cycle fatigue into two branches, de-
fining one as "thermal fatigue" if the temperature is not constant with
time, and the other as "isothermal fatigue" if it is. As an example,
thermal fatigue might result from the starting and stopping of a piece of
high-temperature equipment, while isothermal fatigue might be the con-
sequence of vibration during steady-state operation.
Constraint of free thermal expansion and contraction is a necessary
ingredient in the thermal fatigue process. For convenience, constraints
may be grouped into two general categories: external and internal. Ex-
ternal constraints are those provided by boundary forces applied to the
surfaces of the body which is being heated and cooled. This type of con-
straint is more typical of specimen testing than it is of components in
service. Designers of high-temperature equipment usually take consider-
able care to provide for overall thermal expansion and contraction
through the use of clearances, sliding supports, bellows, and other
mechanical devices. Thus, thermal fatigue with external constraint is
primarily a laboratory testing practice in which external forces on a test
specimen are used to simulate internal thermal stresses in an actual com-
ponent.
In a typical thermal fatigue test with external constraint, a uniaxial
specimen—usually tubular or hourglass in shape—is heated and cooled
uniformly across its test section while constraining forces are applied
through end grips. This test technique is referred to as "thermal-mechani-
cal fatigue testing." End constraints may be as simple as a semirigid
frame which only allows the specimen to lengthen and shorten in ac-
cordance with the elastic compliance of the frame. This simple device is
often called a Coffin-type thermal fatigue apparatus, after its developer.
Dr. L. F. Coffin, Jr., [1].^ Today, the simple constraining frame has
given way to the modern servocontroUed fatigue machine in which
temperature and strain cycles may be applied independently according to
predetermined programs. Whether the apparatus is simple or complex,
thermal-mechanical fatigue is always characterized by external constraint
' The italic numbers in brackets refer to the list of references appended to this paper.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
SPERA ON THERMAL FATIGUE 5
LOW-CYCLE FATIGUE:
LIFE LESS THAN 50 000 CYCLES
SIGNinCANT INELASTIC STRAIN
THERMAL-MECHANICAL THERMAL-STRESS
FATIGUE: FATIGUE:
EXTERNAL CONSTRAINT INTERNAL CONSTRAINT
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
6 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Historical Review
Having suggested some definitions, let us briefly review the develop-
ment of the field of thermal fatigue and the origin of some of the terms
just defined. Much of the information which follows was taken from a
1956 literature survey prepared by Harry Majors, Jr., of the University of
Alabama 12]. The study of material failure as a result of thermal cycling
goes back to 1838, at least. In that year, Duhamel published the first
equations for calculating the thermal stresses in a body subjected to non-
uniform heating [3]. However, it was not until 1894 that fracture was
combined with thermal stress in a quantitative way. This was done in
Germany by Winkelmann and Schott who studied what we now call
thermal shock of ceramic materials, using only rapid heating [4]. Eight
years later, Hovestadt and Everhart of the United States published an
analysis for thermal shock of glass through rapid cooling [5]. And that
was the trend of the thermal stress failure literature for the next 30 or
more years: thermal shock of materials by either rapid heating alone or
rapid cooling alone. Most—but not all—of the materials studied were
ceramics and glass. The earliest reported study of the thermal shock be-
havior of ductile metals was probably that by BoUenrath and co-workers
in 1938, again in Germany [6].
It was not until 1935, almost a century after Duhamel's publication on
thermal stress, that alternate heating and cooling were discussed in the
literature as a cause of cyclic thermal strains. This discussion appeared in
a text on plasticity in crystals, written by Erich Schmid and Walter
Boas [7]. These authors noted that, when a noncubic crystal is alternately
heated and cooled, it gradually increases in size, provided no phase
changes take place. Theoretically, a high-temperature component made of
such an anisotropic material would produce within itself ever-increasing
thermal stresses and eventual cracking. And this cracking would occur
even if the component were heated and cooled slowly, without tempera-
ture gradients and without external constraints. Nine years later. Boas
and a co-worker by the name of R. W. Honeycombe showed that this
was indeed the case for bearings made from a particular tin alloy. Their
paper, published in 1944, was given the rather liberal title of "Thermal
Fatigue of Metals" [8]. This was probably the first time this phrase ap-
peared in the literature.
As we have seen, then, the original definition of thermal fatigue,
credited to Walter Boas in 1944, was very restrictive. It applied only to
failure in a noncubic lattice material slowly heated and cooled without
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions autho
SPERA ON THERMAL FATIGUE 7
any external constraint. During the next decade, this Hmited definition
would be gradually expanded because engineers had a need for a term to
describe high-temperature material failures that were cyclic in nature.
We can guess that it was difficult at first to combine two concepts that
had developed separately: the concept of thermal stress and the concept
of fatigue failure. As an example, let me quote from an old report from
the archives of my laboratory at the Lewis Research Center of the Na-
tional Aeronautics and Space Administration. At the time this report was
written in 1947, our laboratory was called the Aircraft Engine Research
Laboratory of the National Advisory Committee for Aeronautics
(NACA). The title of this report is "Investigation of Rim Cracking in
Turbine Wheels with Welded Blades," and the principal author was
Morton Millenson. To explain the cause of cracking in certain gas turbine
disks, Millenson pointed to cyclic plastic strains, as follows: "Subsequent
cycles of starting and stopping will cause alternate compressive and tensile
flow and progressively weaken the material until cracking occurs in a
manner somewhat similar to the fatigue of metals " I can guess that
Millenson's coauthor would have used a more definite phrase than
"somewhat similar to the fatigue of metals" had it been solely up to him.
The coauthor of this 1947 report [9] was a young man by the name of
S. S. Manson. It would not be long before Manson conceived and di-
rected fatigue research programs at our laboratory and established an
international reputation in the field of high-temperature material be-
havior.
The terminology was changing and developing, however, as it always
does to fill needs. The earUest mention of thermal fatigue in our files at
the Lewis Research Center was in 1949, when John Weeton, a colleague
of Millenson and Manson, used the term to describe one of the failure
mechanisms he observed in combuster liners for gas turbines [10]. In
1950, the term thermal cracking was used by Harry Wetenkamp and his
co-workers at the University of Illinois in their study of the failure of
railway car wheels caused by the heating action of brakes [77]. In 1952,
Helmut Thielsch wrote a survey article for the Welding Research Council
on the subjects of thermal fatigue and thermal shock [72]. Later that
same year, in a lecture deUvered at the University of Michigan, Manson
spoke on the behavior of materials under conditions of thermal stress
[75]. He stated that thermal shock is the failure of a material after a
single cycle of thermal stress while thermal fatigue requires repeated
thermal stress cycles.
Today, that lecture at Michigan is remembered, not for the thermal
stress information it contained, but for Manson's proposal that low-cycle
fatigue Ufe depends primarily on the imposed range of inelastic strain. He
suggested a power law which is known throughout the world today as the
Manson-Coffin equation.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authori
8 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Conclusions
Now another decade has passed, the second since Manson and Cof-
fin introduced life analysis to thermal fatigue. To conclude, let me offer
the following hypothesis. In this second decade, we have advanced from
qualitative to quantitative understanding, and from order-of-magnitude
life estimates to factor-of-two life predictions, at least for conventional
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authoriz
SPERA ON THERMAL FATIGUE
References
[/] Coffin, L. F., Jr. and Wesley, R. P., Transactions, American Society of Mechanical
Engineers, Vol. 76, Aug. 1954, pp. 923-930.
[2] Majors, H., Jr., "Thermal Shock and Fatigue—A Literature Survey," Technical
Report No. 1, Office of Ordnance Research Project No. 1230, Sept. 1956.
[3] Duhamel, J. M. C. in Memoires de I'Institut deFrance, Vol. 5, 1838, p. 440.
[4] Winklemann, A. and Schott, O., Annalen der Physikalischen Chemie, Vol. 51, 1894,
p. 730.
[J] Hovestadt, H. and Everhart, J. D. in Jena Glass, Macmillan, New York, 1902, p. 228.
[6\ BoUenrath, F., Cornelius, H., and Bungardt, W., Jahrbuch der Deutschen Luft-
fahrtforschung, 1938, Vol. 11, pp. 326-338.
[7] Schmid, E. and Boas, W. in Kristallplastizitaet, Verlag Julius Springer, Berlin, 1935,
p. 202.
[S] Boas, W. and Honeycombe, R. W. K., Nature, Vol. 153, April 1944, pp. 494-495.
[P] Millenson, M. B. and Manson, S. S., "Investigation of Rim Cracking in Turbine
Wheels with Welded Blades," NACA RM E6L 17, National Advisory Committee for
Aeronautics, Feb. 1947.
[10] Weeton, J. W., "Mechanisms of Failure of High Nickel-Alloy Turbojet Combustion
Liners," NACA TN 1938, National Advisory Committee for /\eronautics, Oct. 1949.
[77] Wetenkamp, H. R., Sidebottom, O. M., and Schrader, H. J., "The Effect of Brake
Shoe Action on Thermal Cracking and on Failure of Wrought Steel Railway Car
Wheels," Bulletin 387, Engineering Experiment Station, University of lUinois, Urbana,
Vol. 47, No. 77, June 1950.
[72] Thielsch, H., "Thermal Fatigue and Thermal Shock," Welding Research Council
Bulletin, Series No. 10, April 1952.
[13] Manson, S. S., "Behavior of Materials Under Conditions of Thermal Stress," NACA
TN-2933, National Advisory Committee for Aeronautics, July 1953.
[14] Coffin, L. F., Jr., Transactions, American Society of Mechanical Engineers, Vol. 76,
Aug. 1954, pp. 931-950.
[75] Coffin, L. F., Jr. in Symposium on Effect of Cyclic Heating and Stressing on Metals
at Elevated Temperatures, ASTM STP 165, American Society for Testing Materials,
June 1954, pp. 31-52.
[16] Baron, H. G. in Thermal Stress, P. P. Benham and R. Hoyle, Eds., Pitman and
Sons, London, 1964, p. 204.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
D. F. Mowbray^ and J. E. McConnelee^
KEY WORDS: thermal fatigue, fatigue (materials), disks, thermal analysis, stress
analysis
10
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
12 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
The hot bath or maximum temperature (Jmax) was varied between 1500
and 1900°F, with exposure times (4) of 1 or 4 min.
The test material analyzed is the cast cobalt-base superalloy, FSX414.
This alloy has been described in Ref 6. Table 1 gives a summary of static
properties over the temperature range of interest.
Tensile Rupture
80 106 60 18
800 86 41 18
1000 79 36 19
1200 74 33 21
1400 65 31 23 21.0 2150 39
1600 45 30 28 10.0 1150 12
1800 4.5 1870 4
Analysis Methods
Both analysis methods started with the transient temperature fields de-
termined by FE computation. Temperature-dependent values of thermal
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
MOWBRAY AND MC CONNELEE ON TAPERED DISK SPECIMEN 13
conductivity and specific heat were utilized, and the heating and cooling
shocks were simulated by applying time-independent values of convective
heat transfer coefficient on all exterior surfaces. The brief transfer time
from the hot to cold bath was included by specifying a convective heat
transfer coefficient corresponding to still air during the first 3.5 s of the
cooling shock. The thermal properties employed are given in Table 2.
(The heat transfer coefficients are those measured in the baths.) Axisym-
TABLE 2—Thermal and physical properties for FSX414.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions
14 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
/ <
/
/ C B
ISOOr- HEATING
1600- '"^^
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
MOWBRAY AND MC CONNELEE ON TAPERED DISK SPECIMEN 15
Te-Tovq.^F
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
16 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
A = P - Q/(T + 460)
(2)
G = l/(U - VT)
UESI Method
In this method, the total strain-time-temperature histories were ob-
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions author
MOWBRAY AND MC CONNELEE ON TAPERED DISK SPECIMEN 17
Rp=O.OIin.
Tmo»=l688''F
STRAIN,
% 0
The method of solution was patterned after that given by Spera [4].
Total strain is considered to be the sum of three components; that is
(3)
where
£, = instantaneous total strain,
£, = elastic component.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
18 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
o = Ee^ (4)
02 04 06 08 1.0
STRAIN,%
£p = f(a,t„T) (5)
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
MOWBRAY AND MC CONNELEE ON TAPERED DISK SPECIMEN 19
tion technique was applied. Following Spera's [4] lead, the method of
Runge and Kutta [12] was utilized to set up a time incremental calculation
procedure.
Axisymmetric FE Method
The finite element computer program used in this analysis is based on
incremental plasticity theory and the von Mises yield condition emd asso-
ciated flow rule. Although capable of treating combined isotropic and
kinematic hardening, the present analysis was restricted to purely iso-
tropic hardening. The same creep relationships were used as with the
UESI method. Creep stress transfer was assumed in accordance with the
strain hardening rule. The program is currently restricted to the use of bi-
linear stress-strain relationships. The bilinear approximations employed in
the analysis are indicated by dotted lines in Fig. 7.
As discussed earlier, the finite element model used in the stress-stredn
calculations was a very much simplified model with only 16 elements and
33 node points. This simplification was used to reduce the cost of the in-
elastic analysis after preliminary comparisons with results from the com-
plete FE model indicated that this reduced model was adequate to accu-
rately calculate the stress-strain response at the periphery of the disk.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
20 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Tmox=l688''F,Rp=O.OIin.
NUMBERS IN PARENTHESES F(800)
INDICATE TEMPERATURE
IN'F
the stress continues to relax to Point E after the strain achieves a zero
value (Point D).
The cooling shock introduces tensile strain and intially increasing ten-
sile stress. Since the cooling shock starts at a high temperature and the
stress is increasing, creep occurs initially in the cooling shock. At some
point, designated E' in Fig. 8, the temperature dips below where creep
and stress relaxation will occur. The stress and strain continue to increase
to Point F, which is the maximum positive strain generated by the cooling
shock. The peculiar curvature obtained in loading from Points £ to F is a
result of the initial stress relaxation and rapid increases in the flow stress
with the decreasing temperature. At Point F, the strain begins to de-
crease, and the fiber unloads elastically. The unloading extends into the
compressive stress region to the zero strain position. Point G. Close to the
zero strain point, compressive plastic flow initiates. Upon reintroduction
of the heating shock, time-independent plastic flow continues to the mini-
mum strain position (Point B).
A solution from the FE method was obtained for the test conditions
discussed in Figs. 6 and 8. This result is compared with the corresponding
UESI method solution in Fig. 9. Figure 9a shows strain-temperature his-
tories, and Fig. 9b, hysteretic loops. For consistent comparison, the hyster-
etic loop calculated by the UESI method employed isotropic hardening
and the bilinear stress-strain curves.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
MOWBRAY AND MC CONNELEE ON TAPERED DISK SPECIMEN 21
0.2
STRAIN,
1800
— UESI METHOD
o FE METHOD
-0.5 -0.4
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
22 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
0.4 0.5
FIG. 10—Hysteretic loops calculated with differing assumptions on the Bauschinger effect
{UESI method).
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproduct
MOWBRAY AND MC CONNELEE ON TAPERED DISK SPECIMEN 23
Tm0K = l688*F,Rp=0.0liii.
• STRAIN HARDENING
-TIME HARDENING
FIG. 11—Hysteretic loops calculated with the strain hardening and time hardening rules
for creep-stress transfer (UESI method).
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized
24 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Tmox = l688''F
Rp=O.OI in
Rp=0.02
Rp=0.04
-0.5 - 0 4
(a) Effect of « , .
FIG. 12—Calculated hysteretic loops for various test conditions (UESI method).
Rp=0.0lin
-0.6 -0.5
(6)Effectof r^ax.
FIG. n—Continued.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further
M O W B R A Y A N D M C C O N N E L E E ON TAPERED DISK S P E C I M E N 25
Tmo« = l778°F,Rp=0.0lin
•th=4MIN
• th^l MIN
0.3 0.4
STRAIN, %
(c) Effect of h •
FIG. n—Continued.
fatigue life predictions. Although it is not the purpose of the present pa-
per to examine fatigue life of the disk specimens, it is pertinent to exam-
ine the analysis results in light of potential application of current high
temperature life prediction concepts. The following three damage con-
cepts will be considered: (a) creep damage (Spera [4\, (b) strain range
partitioning (Manson et al [13]), and (c) frequency modified fatigue life
(Coffin [14]).
The creep damage hypothesis assumes that high-temperature fatigue
damage is solely the result of creep damage. In his application of it, Spera
proposes summing the damage via the linear life fraction rule, with the as-
sumption that creep in compression is additive to that in tension. For
varying temperature cycles, the creep damage per cycle, <t>, is given by
(6)
0 tr
where
A^ = time increment and
tr = creep-rupture time for the average stress and temperature within
the time increment.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
26 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
N^ = l.O (7)
M pp Fpp " =
^
A
•^^
(8)
' ' cp ^cp -*-'
etc., and
1 1 1 1 1
+ -1- + N,, = Nf (9)
A'.. N,.
In these relations, Npp, N^p, etc., are lives corresponding to the properly
subscripted strain components £pp, £,.p, etc.; Ny is the life when more than
one component is acting.
For thermal loading of the type examined in this paper, only the £pp and
£,p components are significant. These correspond to the Cp and £,. compo-
nents calculated. To use the concept, one would then have to know the
isothermal fatigue properties, a, ft, A, and B. It has been proposed, on
the basis of some initial experimental evidence, that these are independent
of temperature [15]; that is, damage from each type of deformation is in-
dependent of temperature. This possibility would greatly simplify a dam-
age analysis.
The frequency modified fatigue life approach is based on the hypothesis
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
MOWBRAY AND MC CONNELEE ON TAPERED DISK SPECIMEN 27
v^tf = c, (10)
where
V = frequency of cycling,
tf = time to failure, and
k, c = material constants.
It is further assumed that Ci is a constant for each plastic strain range
and that the following relationship exists
C/AE,^ = C2 (11)
where
Lt,p = the total plastic strain range (Atp + AeJ and
ji, C2 = material constants.
Taking vtf = Nf, the following modified Coffin-Manson form can be
arrived at
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
28 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
References
[/] Glenny, E., et al, "A Technique for Thermal Shock and Thermal Fatigue Testing Based
on the Use of Fluidized Solids," Journal of the Institute of Metals, 1958, Vol. 87, pp.
294-302.
[2] Howes, M. A. H. in Fatigue at Elevated Temperatures, ASTM STP 520, American So-
ciety for Testing and Materials, 1973, pp. 242-254.
[3] Howe, P. W. H. in Thermal and High-Strain Fatigue, The Metals and Metallurgy
Trust, 1967, pp. 122-141.
[4] Spera, D. A., "The Calculation of Thermal Fatigue Life Based on Accumulated Creep
Damage," NASA TMX-52558, National Aeronautics and Space Administration, 1969.
[5] Mowbray, D. F. and Woodford, D. A., "Observations and Interpretation of Crack
Propagation Under Conditions of Transient Thermal Strain," International Conference
on Creep and Fatigue, Institute of Mechanical Engineers, 1973, pp. 179.I-179.il.
[6] Foster, A. D. and Sims, C. T., Metal Progress, 1969, pp. 83-95.
[7] Mendelsohn, P/aif/c/y.' Theory and Application, MacMillan, New York, 1968.
[51 Hunsaker, B., et al, "A Comparison of the Capability of Four Hardening Rules to Pre-
dict a Material's Plastic Behavior," Report #75-51. Texas Engineering Experiment
Station, June 1975.
[9] Penney, R. K. and Marriott, D. L., Design for Creep, McGraw-Hill, London, 1971.
110] Manson, S. S., Thermal Stress and Low-Cycle Fatigue, McGraw-Hill, New York, 1966.
[11] Topper, T. H. et al. Journal of Materials, Vol. 4, No. I, 1969, pp. 200-209.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
MOWBRAY AND MC CONNELEE ON TAPERED DISK SPECIMEN 29
[12] Hildebrand, F. B., Advanced Calculus for Engineers, Prentice-Hall, Englewood Cliffs,
N.J., 1949.
[13] Manson, S. S., Halford, G. R., and Hirschberg, M. H., "Creep-Fatigue Analysis By
Strain Range Partioning," NASA TMX-67838, National Aeronautics and Space Ad-
ministration, May 1971.
[14] Coffin, L. F., Jr., Journal of Materials, 1971, Vol. 6, pp. 388-402.
[15] Halford, G. R., Hirschberg, M. H., and Manson, S. S. in Fatigue at Elevated Tempera-
tures, ASTM STP 520, American Society for Testing and Materials, pp. 658-669.
[16] Coffin, L. F., Jr., Proceedings, Institute for Mechanical Engineers, 1974, Vol. 188, pp.
109-127.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
D. C. Gonyea'
KEY WORDS: thermal fatigue, thermal stress, stress analysis, stress concentration
30
Modeling Technique
Figure 1 depicts the three general types of geometries that have been
studied, and Table 1 lists the actual dimensions utilized in this study. A
total of 15 geometries has been investigated which encompasses the range
of parameters typically encountered in actual turbine geometries of the
author's company.
'tufytff
% ROTOR
FIG. 1—Sketch of three types of geometries studied and definition of geometric param-
eters.
TABLE 1—Geometry description and thermal stress concentration factors found using
finite element approach. All dimensions are given in units of inches. See Fig. I for definition
of symbols.
Case D DS D\ R L TSCF
1 20 42 3 3.72
2 20 30 3 2.66
3 20 30 1 3.57
4 20 22 1 1.77
5 20 24 1 2.33
6 20 20.5 0.25 1.74
7 12.3 24 4 2.60
8 22 22 39 0.5625 5 2.03
9 29.25 29.25 39.6 1.25 6 .75 1.80
10 22 22 28.25 0.4375 2 .25 1.74
11 20 30 40 4 4.06
12 20 26 30 4 3.39
13 20 23 30 A 2.56
14 20 23 26 4 2.67
15 20 26 30 6 3.19
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
32 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
SURFACE
TEMPERATURE r
• THERMALLY INSULATED
• AXIAL DISPL. EQUAL
SAME AS • NO SHEAR LOADS
OPPOSITE • NET AXIAL LOAD ZERO
END
• SOLID, AXISYMMETRIC
• UNIFORM INITIAL TEMPERATURE
• NO EXTERNAL OR BODY LOADS
FIG. 2—Thermal and mechanical boundary conditions utilized in finite element study.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
GONYEA ON LARGE SHAFTS 33
Temperatures/Stress Response
Figure 3 illustrates the temperature and stress response that is found to
result from the surface ramp temperature change. The interior tempera-
tures lag the surface temperature on heating and reach a quasi-steady
OUTER / ^
SURFACE / ^
TEMPER- \ / /
ATURE / / \cENTER OF
SHAFT
/ /
L^"
MAX. FILLET
' STRESS, <r'
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
34 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Oa
TSCF = (1)
where
TSCF = thermal stress concentration factor,
oa = maximum Von Mises equivalent stress, and
Ooo = Von Mises equivalent stress removed from the fillet.
In Eq 1, the values of stresses are taken after a quasi-steady-state
temperature field develops. During the initial heating portion, the magni-
tude of the thermal stresses are less than those developed during quasi-
steady state. Figure 4 illustrates this initial time dependent nature of the
thermal stresses.
_ K..1 '<, = ^ A T T , M E ,
K, = — AT QUASI-STEADY STATE
a = THERMAL DIFFUSIVITY
FIG. 4—Time dependence of thermal stress concentration during initial portion of heat-
ing. Note that the range illustrated includes wheel, step, and combined geometries.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized
GONYEA ON LARGE SHAFTS 35
FIG. 5—Normalized distribution of surf ace stress in region of fillet. All cases investigated
included in scatterband.
= -v^[^°->f
TSCF = I +\/ - \- ^ - 1r (2)
'Peterson, R. E., Stress Concentration Design Factors, Wiley, New York, 1953.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
36 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
where
DEQ DS/D for step geometry
D minimum [Dl/D, DS/D + 0.35 L/DS] for wheel and com-
bined geometry
It should be noted that TSCF is defined in terms of the nominal equiv-
alent stress at diameter, D, in all cases.
Figure 6 illustrates that Eq 2 describes the results of the finite element
studies connected with the step geometries in the range of parameters
studied. Figure 7 is a direct comparison of TSCF found by the finite ele-
ment technique and by the proposed analytical relationship, Eq 2 for all
cases studied.
10
R ^
X'P^-^
TSCF-1
1
. ^ '^X
y. ^
FIG. 6—Comparison of finite element results for step geometry and equation developed
to define thermal stress concentration behavior.
1 2 3 4
TSCF FROM EQUATION
FIG. 7—Comparison of TSCF calculated using the finite element technique from Eq 2.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authori
GONYEA ON LARGE SHAFTS 37
0.0^^^'
• TSCF STUDIES
MECHANICAL BENDING FROM PETERSON
THERMAL STRESS CONCENTRATION EQUATION
0 1 1 LXJ I I l_l_L -L. -L.
0.01 0.1 10
f-'
FIG. 8—Comparison of thermal and mechanical stress concentration factors for shaft
with step change in diameter.
cantly larger than the corresponding mechanical value. This effect is due
principally to a thermal mismatch mechanism that gives rise to additional
stress in the case of the thermal loading, particularly for large diameter
ratios. The difference in average temperatures between the large and small
sections of shaft causes this thermal mismatch. A similar mechanism does
not exist for the case of mechanical loading.
Summary
The finite element technique has been used to investigate the nature
of thermal stress concentrations in large turbine rotors during tempera-
ture transients. On the basis of this analysis, it is tentatively concluded
that the thermal stress concentration factor, TSCF, is independent of time
and heating rate under conditions of slow uniform heating in which a
quasi-steady state temperature field exists. An approximate relationship
is given, Eq 2, which defines TSCF in terms of the geometric parameters
of the rotor. A normalized plot of the distribution of thermal stress in
the vicinity of the fillet region is also given. A comparison of TSCF
with the mechanical stress concentration factor due to bending illustrates
that there may be a significant difference in numerical values of these
two factors. For large diameter ratios, the value of TSCF is larger than
that due to bending, owing to an additional thermal stress component
which is due to a thermal mismatch mechanism.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
S. Tepper^
REFERENCE: Tepper, S., "Low-Cycle Fatigue Analysis of the Turbine Disli for
the National Aeronautics and Space Administration High-Temperature Turbine
Rig," Thermal Fatigue of Materials and Components, ASTM STP 612, D. A. Spera
and D. F. Mowbray, Eds., American Society for Testing and Materials, 1976, pp.
38-54.
ABSTRACT: The high-pressure, high-temperature turbine test rig was designed for
the National Aeronautics and Space Administration to test turbine components in an
environment that simulates current and future gas turbines operating at 41 atm
(600 psi) and 2204 °C (4000 °F). This paper illustrates the design approach that led
to selection of a typical test rig component: the turbine disk. The final configuration
of the rotor assembly and the turbine disk was determined by analysis of low-cycle
fatigue and creep.
KEY WORDS: thermal fatigue, creep (materials), disks, thermal cycling tests, plastic
analysis, substructures, turbines
38
by the universal slopes method according to Manson and Halford [7]' and
considering the studies of Morrow, Wetzel, and Topper [2,3]. The inter-
action between creep and LCF, and the strain range partitioning approach
are also discussed [4-6].
Initial Guidelines
Step 1—Establish all specified and tabulated operating conditions. If
each contributory action for a given operating condition could be linearly
superposed, the following expression could be used during the preUminary
design
^ = i i'^Sai
Sa
^<l
' The italic numbers in brackets refer to the list of references appended to this paper.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized
40 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
TEPPER ON LOW-CYCLE FATIGUE 41
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
42 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
where
S = any design condition, for example, stress and hours of life,
a = allowable value,
e = equivalent total, and
/ = corresponding contributory condition.
In this analysis, the only linear contributory load, neglecting differen-
tial stiffening effects, is the pressure. For a given temperature distribution
and rotational speed, only two conditions were analyzed: zero and maxi-
mum pressure. The other two contributory actions were nonlinear, that
is, temperature and rotational speed.
Step 2—Establish intermediate operating conditions if the acting loads
cannot be linearly superposed. In the present study, seven speeds (six
operational and one for overspeed design), four pressure values (two at
each side of the disk), and six temperature maps existed. These variables
produce a total of 192 combinations. However, not all of these combi-
nations are operational.
Step 3—Discard nonoperational and minor operating conditions. Once
the obviously insignificant conditions are eliminated, the discarding of
some intermediate operating conditions can be very challenging. An
example to be considered would be, which temperature maps are the most
significant? Here one must choose between high-temperature conditions
with lower temperature gradients, and low-temperature conditions with
large temperature gradients. The best method is to keep both extreme
conditions and at least one intermediate condition.
Step 4—Retain conditions that define a cyclic operation. Basically, one
tends to study or select only the highly loaded conditions because of their
relation to ultimate or yield strength and to creep life. This approach
is valid when the representative mission profile indicates few cycles.
Frequently, however, that is not the case.
The mission profile analyzed here comprises many long and short tests
in which the corresponding transients affect each of the contributory
parameters—speed, pressure, temperature—and their mutual relationships.
Cyclic life is a fundamental condition to satisfy.
Step 5—If conditions allow, start using an elastic model. Once the con-
tributory loads are defined (in this case, speed, pressure, and temperature
maps), an elastic finite-element model may be used. This model may be
used even if the operating conditions are not Unear. Use of an elastic
finite-element model assumes that: (o) the possible plastic areas are con-
fined to local small spots and (b) these spots will not create a mechanism
between two or more parts of the model.
Elastic models are more economical than plastic models because they
run faster in the computer. Also, there are many large computer pro-
grams able to accept elastic models. In many cases, an elastic-plastic
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions aut
TEPPER ON LOW-CYCLE FATIGUE 43
960 IN.-LB a
(108 Nm) vj,
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authori
44 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
+ 0.0022860
+0.0003566 r—1-^
+ 0.0020674
mj
8° ^^^^^^ \ + 0.0018034
60 Tr"^'/^
\
SO 1 *^
(8.701
L
r ^
+ O.OO13970
+0.0004572 1
' li=L,4+0.000«096
100 1===^
no L + 0.0006080
130 (18.861
+0.0003912-
w
b.i
a
side plate and establish the matching and boundary loads for a sub-
structure surrounding the area under study.
Also, two other actions (bolt loosening and tightening) were simul-
taneously analyzed. By using a bolt made of the same material as the
disk, the thermal expansion was the same for both. However, the cen-
trifugal field acting on the disk produced two principal stresses (radial Op
and hoop Oe) of the same order, while the bolt is mainly under axial
loading. Because the Poisson's ratio effect narrows the disk farther than
the bolt, the bolt could become loose if sufficient preloading is not
provided.
Once preloading is defined, the effect of the following must be investi-
gated: (a) the different working conditions at room temperature and at
idle and maximum operating speeds, {b) the creep associated with the
stress and temperature field, and (c) the maximum manufacturing toler-
ances allowed.
Figure 6 depicts the load distribution at the disk-sideplate interface.
The contact pressure loads are transferred through the bolt. Matching
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions aut
TEPPER ON LOW-CYCLE FATIGUE 45
5027 HR
LEGEND:
AT 17,500 RPM
^ AT21.S00RPM
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions autho
46 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
P, + P2~0
OPtiPi.Pt.H.n) — Op,(^p,^p,,H,n)
where
subscripts = degrees of freedom associated with the
loads,
variables between parenthesis = nonlinear dependencies,
superscripts = loading configurations,
" = free centrifugal field, (17 500 rpm),
R = total resultant displacement along the
radial direction, and
Pi, Pi, and Hi = loading resultants from the contract
pressures PRS, PRD, and PH respec-
tively (see Fig. 6).
The solution of the system of equations gave the shear stresses at the
boh section and the axial preload required to keep the sideplate attached
to the disk. Considering the additional preload required to control the
thermal expansion, the Poisson's ratio effect, the manufacturing toler-
ance, and the creep relaxation, the maximum axial stress in the bolt
is 75 to 80 ksi. The lower axial stress is about 40 ksi for the worst com-
bination of operating conditions. Bolt retightening time is defined by both
creep relaxation and cyclic operation.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
TEPPER ON LOW-CYCLE FATIGUE 47
TEMPERATURE, ° F r C I
These life zones are depicted in Fig. 9 for both 17 500 rpm and 21 800
rpm and the established creep values.
If the LMP curves reflect minimum values, a safety factor of four is
common practice. However, it should be noted that the expected Ufe
was based on an elastic-plastic solution where no stress relaxation was
considered.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
48 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
ioo\ \
11 \ K"
J \ioo
(14.50)y \ (17.40)YV 4ioo
V120 >*
100
/4^
nt
'120
0*0 \^
0 ,
/100
.
yi2o
^
120
100 /
(14.501
100
M /
1 ^
(a) At 17 500 rpm, ksi (MPa).
(b) At 21 800 rpm, ksi (MPa).
FIG. i—Effective stresses.
in the substructure. After the runs, recheck the interface and boundary
agreements.
Step <S—When plasticity is expected, run the full elastic-plastic analysis
for the critical substructure or for the total structure, if required. A
first assessment of possible plasticity should compare the effective stress
with the corresponding temperature and yield stress at a given point.
An effective stress below the yield stress does not guarantee elastic values.
However, the analyst will be able to check the different types of stresses
involved and the particular area under consideration. If a point shows
plasticity, all the substructures (or the structure) should be run in a plastic
solution.
Step P—For the specified creep value (1 percent, 2 percent, rupture,
etc.), define the creep Ufe map. Since no stress relaxation is considered
at this time, the LMP approach should suffice. A typical LMP expression
is
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
TEPPER ON LOW-CYCLE FATIGUE 49
where
T = absolute temperature (temperature °F + 460),
t = expected life in hours, and
log = logarithm to base 10.
Most of the materials have curves relating stresses to LMP
LMP=/{(o)K
Step 10—Clearly mark the plastic areas with special emphasis on the
following three aspects: (a) plastic areas on the surface, (b) if there is
more than one area, mark possible connectivity, and (c) plastic areas
that are large in proportion to their surrounding elastic areas.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
50 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Peterson [7] has shown that notch size can affect the fatigue life and
that a fatigue concentration factor, K,, should be used
where
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized
TEPPER ON LOW-CYCLE FATIGUE 51
(3)
\ ASAE /
Kf {LSbeE)"^ = (AoAeE)'
where E is the Young's modulus.
If strain, e, is elastic, then
For a given load and geometry, the left side of Eq 4 is a constant. The
set of solutions of Ao and Ac, therefore, form a hyperbola as shown in
Fig. 10. The intersection at A in Fig. 10 is the solution if the stress and
strain are entirely elastic. Points B and C are the location of true stress
and strain for two different materials that have the same elastic modulus
but different strengths.
Crews and Hardrath [8\ experimentally demonstrated the validity of
using these control hyperbolas to determine the residual stresses and
strains after a complete cycle. This work was carried farther by Wetzel
[9] in demonstrating the use of control hyperbolas to determine actual
stress and strain throughout the load history.
Figure 11 shows how the actual stresses and strains may be followed
throughout the load history by returning to the control hyperbolas after
each load change. It may also be seen how the hysteresis loop stabilizes
and strain hardening has taken place for the particular material.
Manson and Halford [1\ developed an empirical relationship between
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authoriz
52 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
1ST HYPERBOLA
3R0 HYPERBOLA
2ND HYPERBOLA'
lOTH CYCLE
60TH CYCLE
600TH CYCLE
total strain and the number of cycles required to cause fatigue failure
by using conventional tensile properties, as follows
. , ^ UTS.,. + Dr'^-^Nf
(6)
Ac = 3.5 —^Nf
where
AE, = stabilized or final total true strain range,
UTS = ultimate tensile strength,
D, = tensile ductility, and
Nf= cycles to failure.
This formula has been shown to fit fatigue data for many materials. It
is adequate for predicting pure fatigue life where temperature and loading
frequencies are such that cyclic creep damage is not a factor. Many in-
vestigations have been performed using Eq 6 with acceptable correlation
between fatigue life (N/) and total strain (Ac,).
By using this formula, the value AE, can be obtained directly from the
plastic analysis without going through the Neuber's hyperbola approach.
In analyzing the HTTR, both Neuber's hyperbola and the plastic solution
were used. During the selective analysis, however, the analysis was limited
to the nominal or equivalent elastic stresses and the Neuber's hyperbola
approach.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
TEPPER ON LOW-CYCLE FATIGUE 53
where cp is the life fraction and subscripts C and LCF correspond re-
spectively to creep and low-cycle fatigue
n m
qoc = J. t^(pc, cpLCF = Z ACPLCFM (8)
A = 1 Af = I
A ^* A ^^" /m
Acpc, = — AcpLCF„ = T T - (9)
where
hk = hours spent at condition k,
hi, = total hours possible at condition k,
CY„ = number of cycles used at condition M, and
Cy,„ = total number of cycles possible at condition M.
During the initial phases of the analysis, this method is sufficient for a
comparison of the design alternatives. As can be observed, the results are
the same if most of the cycles are spent at the beginning or at the end of
the creep life. This method does not consider the cycle frequency, nor
does it consider if the mechanical properties of the material related to the
creep can be affected by the cyclic operation.
In Manson and Halford's relationship [7], the number of cycles as
calculated by the Eq 6 was adjusted for creep phenomena. This was done
by using values related to the effective fraction of each cycle for which
the material may be considered to be subjected to the maximum stress,
frequency of stress application, and material coefficients.
In recent years, new techniques have been discussed; the most promising
among them is strainrange partitioning [4-6\. By this technique, any
cyclic behavior can be represented as a combination of four basic types
of inelastic strainranges defined as plastic-plastic, plastic-creep, creep-
creep, and creep-plastic. However, for a given cycle, not more than three
of these types can exist. Each one of these basic types must be tested for
a given material. To take full advantage of this new technique, considera-
ble effort must be expended to compile a library of materials for the
basic cycle types.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
54 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
References
[/] Manson, S. S. and Halford, G., "A Method of Estimating High Temperature Low-
Cycle Fatigue Behavior of Materials," NASA Report TMX 52357, Lewis Research
Center, National Aeronautics and Space Administration, Cleveland, Ohio, Jan. 1967.
[2] Morrow, T., Wetzel, R. M., and Topper, T. H. in Effects of Environment and Complex
Load History on Fatigue Life, ASTM STP 462, American Society for Testing and
Materials, 1970, pp. 74-91.
[3] Morrow, T., Wetzel, R. M., and Topper, T. H., Journal of Materials, Vol. 4, No. 1,
March 1969, pp. 200-209.
[4] Spera, D. A. in Fatigue at Elevated Temperatures, ASTM STP 520, American Society
for Testing and Materials, 1973, pp. 648-657.
[5] Manson, S. S. in Fatigue at Elevated Temperatures, ASTM STP 520, American Society
for Testing and Materials, 1973, pp. 744-782.
[6\ Halford, G. R., Hirschberg, M. H., and Manson, S. S. in Fatigue at Elevated Temper-
atures, ASTM STP 520, American Society for Testing and Materials, 1973, pp. 658-669.
[7] Peterson, R. E., Material Research and Standards, Vol. 3, No. 2, Feb. 1963, pp. 122-
139.
[8] Crews, J. H. and Hardrath, H. F., Experimental Mechanics, Vol. 6, No. 6, June 1966,
pp. 313-320.
[9] Wetzel, R. M., Journal of Materials, Vol. 3, No. 3, Sept. 1968, pp. 646-657.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
D. P. H. Hasselman,' R. Badaliance,' andE. P. Chen^
KEY WORDS: thermal fatigue, crack propagation, silicon nitrides, silica glass, frac-
ture, fatigue failure.
55
Materials
Soda-Lime-Silica Glass
Soda-lime-silica' glass in the form of rods with radius of 0.236 cm,
identical to the glass studied previously [10], was selected for the present
program. Table 1 lists the appropriate physical properties of this glass.
Experimental data for slow crack growth were obtained from the literature
TABLE l~Datafor physical properties, crack propagation behavior, and heat transfer required
for calculations of thermal fatigue resistance of soda-lime-silica glass rods subjected to a
water quench.
PHYSICAL PROPERTIES
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
HASSELMAN ET AL ON BRITTLE CERAMICS 57
where
V = rate of slow crack growth,
Vo = preexponential factor,
b = a constant,
Q = activation energy,
R = Boltzmann constant, and
T = absolute temperature.
Values for b and Vo corresponding to the high and low Ki are given in
Table 1. The experimental data for crack growth suggests that no crack
growth occurred below the fatigue limit (A"o).
Strength of the glass rods was measured at liquid nitrogen {N2) temper-
ature in four-point bending with a span between the central loading points
of 2.4 cm. This gave an average value of strength of 2.38 X 10* Nm'^
ranging from a minimum value of 1.94 X 10' to a maximum value of
2.76 X 10' Nm"^ for a total of ten specimens. A statistical analysis of the
data gave a value for the Weibull parameter, q ~ 8," also listed in Table 1.
Silicon Nitride
Thermal fatigue data for hot-pressed silicon nitride' were obtained by
Ammann et al [12] using a specimen geometry which approximated the
shape of a vane in a turbine engine. As shown by Evans et al [13], this
rate of slow crack growth in this material could be expressed by
V = AKrexp(-Q/RT) (2)
where
V = crack-velocity in ms"',
A, n = constants, and
Ki, Q. R, and J = as defined in Eq 1.
Table 2 lists the appropriate values for the constants in Eq 2, together
with the values for the physical properties pertinent to the thermal stress
calculations and fatigue predictions. Tensile strength at room temperature
as reported by Nessler [14] ranged from 2.41 to 3.71 x 10' Nm"^ with an
average value of 3.33 x 10' Nm"^
Thermal Environment
Soda-Lime-Silica Glass
The glass rods were subjected to thermal fatigue by repeated quenching
'This symbol corresponds to the letter m in the original Weibull theory.
'HS-130, Norton Company.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
58 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
PHYSICAL PROPERTIES"
Constant(y4) 5 x 10'''mkgsunits(m/s)
Constant (n) 6
Activation energy (Q) 170 kcal • mol"'
Critical stress intensity factor (A'lc)
Room temperature 4.7 MNm""^
1000°C S.OMNm-'"
" Representative values at 1000 °C, see text.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
HASSELMAN ET AL ON BRITTLE CERAMICS 59
Silicon Nitride
The thermal fatigue resistance of the silicon nitride, as determined by
Ammann et al [12], consisted of cycling the specimens between an elec-
trically heated laboratory oven and fluidized bed at 40°C. The specimen
was specifically selected such that, for the heat transfer characteristics of
the fluidized bed, the temperature and thermal stress history of the speci-
men was similar to that experienced by a vane in a turbine engine under-
going rapid shutdown. As calculated by finite element analysis, the vane,
originally at 1200°C and suddenly subjected to compressor air at 410°C,
experiences a transient thermal stress at the forward edge as shown in Fig.
la with a peak value of approximately 3.0 x 10' Nm"^, also included in
Fig. la. The transient temperature of the specimen at the forward edge
is shown in Fig. lb.
FATIGUE SPECIMEN
TIME (SECONDS)
FIG. la—Transient thermal stress history of silicon nitride fatigue specimen quenched into
fluidized bed and vane in turbine engine subjected to rapid shutdown.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
60 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
iZ 500-
TIME (SECONDS)
mal quench. The crack depth was assumed to be much smaller than the
specimen size, such that the crack configuration is closely approximated
by an edge crack in a semi-infinite solid. Also, it was assumed that the
presence of the crack did not affect the transient temperature field. Fi-
nally, in view of the shallow crack depth relative to the specimen size, it
was assumed that the crack "sees" a uniform temperature and stress field
calculated to exist at the surface of the rod. This latter assumption was
shown [15] to be valid for the ratios of flaw depth to specimen size of the
present study. For the glass rods, the temperature and stresses were cal-
culated directly from general equations given in the literature for infinitely
long rods of circular cross section. For the silicon nitride specimens, the
analytical expressions for the temperature and stresses in a cylindrical rod
were used by adjusting the appropriate parameters such that the actual
temperatures and stresses were closely matched to those shown in Fig. 1.
The use of analytical expressions for the transient temperatures and ther-
mal stresses considerably simplifies the calculations of slow crack growth
for the silicon nitride specimens.
The transient surface temperatures, T{R,t) of the rod under condition
of convective heat transfer is given by [16]
2pJ,(pn)
T{R, t) = To + (T, - To) I txp{-p„'xt/R') (3)
n^iim' + /3„Vo(^n)
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement.
HASSELMAN ET AL ON BRITTLE CERAMICS 61
where
R = cylinder radius,
To = temperature of quenching medium,
r, = initial uniform temperature of the cyUnder,
m = Biot's modulus with m = Rh/k,
h = heat transfer coefficient,
k - thermal conductivity,
K = therm2il diffusivity with JC = k/qc,
Q = density,
c = specific heat,
t = time,
Jo and /i = Bessel functions of the zero and first order,
^„ = the root of
From the distribution of the transient temperatures in the rod, the tran-
sient thermal stresses at the surface of the rod for the state of generalized
plane strain are [17]
OriR, 0 = 0
J.iP.) (,^
(1 - V) (. /.= ! (m^ + P^')Jo'Wn)
where
E = Young's modulus of elasticity,
V = Poisson's ratio, and
a = coefficient of thermal expansion.
For the water quench, the heat transfer coefficient, h~0.lcal- cm"^• s"' °C"'
was estabUshed by Evems et al [75]. The corresponding value of the Biot
modulus for this value of h and rod radius of 0.236 cm is m ~ 10. For the
silicon nitride subjected to thermal fatigue in the fluidized bed, a peak
value of thermal stress of 3.0 x 10» Nrn'^ for a value of AT = 1260°C
corresponds to an effective Biot modulus, Weff - 0.65. Similarly, the time
dependence of the transient temperature and thermal stress corresponds to
that of a circular cylinder with effective radii of /?eff = 0.60 and 1.135 cm,
respectively. These values of ntetf and Rett are consistent with the average
specimen dimensions and the heat transfer coefficient of the fluidized bed
with/i = 850Wm-^°C-'.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
62 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
In general, the crack depth (a) at any time t can be obtained by integra-
tion of Eq 1 or 2, that is
Copyright by ASTM
Downloaded/printed by
University of Washington
HASSELMAN ET AL ON BRITTLE CERAMICS 63
For o, = 2.38 X 10» Nm"^ and a value of Ki, = 8.7 x 10' Nm''''^ at
liquid Ni temperature as reported by Wiederhorn [20], an initial flaw
depth,flo~ 8.6 ^m can be calculated from Eq 5. This value of initial flaw
depth should correspond approximately to the flaw depth of the fifth
specimen out of nine to fail by thermal fatigue.
No such statistical analysis was carried out for the strength data of the
silicon nitride since the number of data points available was insufficient
to obtain a reliable value of the Weibull parameter q. For a critical stress
intensity factor, Kic = 4.7 MN-m"''^ [27] at room temperature, the strength
values reported earlier correspond to flaw depths ranging from about 40
to 95 nm, with a value of flaw depth of approximately 50 /im for the aver-
age value of strength.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized
64 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
EXPERIMENTAL
FIRST SPECIMEN OF 9
FIFTH SPECIMEN OF 9
PREDICTED
FATIGUE QQ • 10/i.m
LIMITS
8/i.m lorc.
Oum 9I»C.
CYCLES
EXPERIMENTAL
BATH TEMP. I FIRST SPECIMEN OF 9
65»C. i FIFTH SPECIMEN OF 9
• PREDICTED
QQ ' 8/im
FIG. 2—Predicted and experimental number of cycles-to-failure of glass rods with initial
flaw-depth = 8.6 jjm subjected to thermal fatigue into water at (a) 33 °C and (b) 65 °C.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions aut
HASSELMAN ET AL ON BRITTLE CERAMICS 65
10 100 1000
CYCLES TO FAILURE
FIG. 3—Experimental and predicted thermal fatigue behavior of silicon nitride subjected
to thermal cycling from high temperature into afluidized bed at 40°C.
depth much higher than the flaw depth corresponding to the average
strength of the set of specimens discussed earlier. Again, this discrepancy
is likely to be due to the statistical nature of brittle fracture referred to
earlier. In Fig. 3, as expected, the slope of the predicted curves corre-
sponding to the lower values of initial crack depth and higher tempera-
tures are steeper than those for the larger values of initial crack-depth.
Again, this is due to the simultaneous decrease in temperature as well as
level of thermal stress with decreasing temperature difference. For initial
crack depths of 80, 90, and 100 ptm, the temperature levels and thermal
stresses are sufficiently low such that the crack growth per cycle is less
than 10"' [oa. This implies that for these values of a^, the specimen stress
is just below the fracture stress, even at the first cycle. For these flaw
sizes, even a very small decrease in AT can make the specimens last almost
indefinitely. However, the low slopes of thermal fatigue curves imply
that, for accurate thermal fatigue predictions, thermal stresses, tempera-
tures, crack depths, and crack-growth behavior need to be established pre-
cisely.
The results for the glass and silicon nitride as just presented indicate
that reasonably accurate estimates of thermal fatigue resistance can be
made. However, the relative complexity of making such failure predic-
tions is clearly indicated from Table 1, which, including the initial temper-
ature of the specimen, lists a total of 17 values of physical properties,
crack propagation behavior, and thermal environment required for such
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorize
66 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions autho
HASSELMAN ET AL ON BRITTLE CERAMICS 67
1
«
2000
lOOC
6on
2 N
lb 90 too
^
CRACK DEPTH </im)
Acknowledgment
This study was conducted as part of a long-range research program on
thermal stress fracture of brittle materials supported by the Army Re-
search Office, Durham, under grant DA-ARO-A-31-124-73-G45.
References
[/] Williams, L. S., Transactions of the British Ceramic Society, Vol. 55, 1956, p. 287.
[2] Krohn, D. A. and Hasselman, D. P. H., Journal of the American Ceramic Society,
Vol. 55, 1972, p. 208.
[3] Gurney, C. and Pearson, S., Proceedings, Physics Society, London, Section B, Vol. 62,
1958, p. 537.
[4] Kossowsky, R., Journal of the American Ceramic Society, Vol. 56, 1973, p. 531.
[5] Sarkar, B. K. and Glinn, T. G. J., Transactions of the British Ceramic Society, Vol.
69, 1970, p. 199.
[6] Tumanov, V. I., Goldberg, Z. A., Chernyshev, V. V., and Pavlova, V. L, Porosh-
kovaya Metallurgiya, Vol. 10, 1966, p. 71.
[7] Fracture Mechanics of Ceramics, Vol. 2, Microstructure, Materials and Applications,
Bradt, R. C , Hasselman, D. P. H. and Lange, F. F., Eds., Plenum Press, New York,
1974.
[S] Evans, A. G., Journal of Materials Science, Vol. 7, 1972, p. 1137.
[9] Evans, A. G. and Fuller, E. R., Metallurgical Transactions, Vol. 5, 1974, p. 27.
[10] Badaliance, R., Krohn, D. A., and Hasselman, D. P. H., Journal of the American
Ceramic Society, Vol. 57, 1974, p. 432.
[U] Wiederhorn, S. M. and Bolz, L. H., Journal of the American Ceramic Society, Vol. 53,
1970, p. 543.
[12] Armann, C. L., Doherty, J. E., and Nessler, C. G., Journal of Materials Science, to
be published.
[13] Evans, A. G., Russell, L. R., and Richerson, D. W., Metallurgical Transactions, Vol.
6A, 1975, pp. 707-716.
[14] Nessler, C. G., Transactions of the American Society of Mechanical Engineers, to be
published.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
68 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
[15] Evans, A. G., Linzar, M., Johnson, H., Hasselman, D. P. H., and Kipp, M. E., Jour-
nal of Materials Science, to be published.
[16] Carslaw, H. S. and Jaeger, J. C., Conduction of Heat in Solids, Clarendon Press, Ox-
ford, England, 1969.
[IT] Boley, B. A. and Weiner, G. H., Theory of Thermal Stresses, Wiley and Sons, New
York, 1960.
[18] Sih, G. C , Handbook of Stress Intensity Factors, Lehigh University, Bethlehem, Pa.,
1973.
[19] WeibuU, W.,Akad. Handl., No. 151, 1939.
[20] Wiederhorn, S. M., Journal of the American Ceramic Society, Vol. 52, 1969, p. 99.
[21] Wiederhorn, S. M. in Fracture Mechanics of Ceramics, Bradt, R. C , Hasselman, D. P.
H., and Lange, F. F., Eds., Plenum Press, New York, 1974.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
D. A. Spera^ andE. C. Cox^
Description of a Computerized
Method for Predicting Tliermai
Fatigue Life of Metals
ABSTRACT: A computer program called TFLIFE which can be used to predict the
thermal fatigue life of metals and structural components from conventional metal
properties is described. This program is used as a subroutine with a main program sup-
plied by the user. The main program calculates input cycles of temperature and total
strain for TFLIFE which then calculates a stress cycle, creep and plastic strain damage,
and cyclic life. A unique feature of TFLIFE is that it calculates lives according to
several different failure criteria for the same input data. These criteria are surface crack
initiation, interior crack initiation, and complete fracture of both unnotched and
notched fatigue specimens. Sample output tables are shown, together with results for
two typical problems: (a) thermal-mechanical fatigue of bar specimens of the tantalum
alloy T-111 eind (6) thermal-stress fatigue of wedge specimens of the nickel alloy
B-1900. Thermal fatigue lives calculated using TFLIFE have been verified by com-
parison with a variety of laboratory test data on different types of alloys. The com-
puter program is now ready for more extensive evaluation on structural components
as well as additional laboratory specimens.
KEY WORDS: thermal fatigue, fatigue (metals), crack initiation, computer programs,
thermal stress, life prediction, creep
69
The original version of TFLIFE was written more than eight years ago to
aid in the development of a creep damage theory for thermal fatigue [1,2].^
Since that time, it has grown in scope and function, serving as a useful tool
for research on high-temperature metals [3]. It now includes recent develop-
ments in creep-fatigue theory and input/output formats which make it
suitable for general use. To date, TFLIFE results have been verified by ther-
mal fatigue tests on laboratory specimens such as bars, tapered disks,
wedges, and simulated turbine blades [J-4]. In these tests, thermal and
mechanical loads were applied by fatigue machines, fluidized beds, and
burner rigs to specimens of nickel, iron, cobalt, and tantalum alloys.
TFLIFE is now ready for more extensive evaluation by the technical com-
munity on complete components as well as laboratory specimens.
A unique feature of TFLIFE is that it calculates thermal fatigue lives ac-
cording to each of three failure criteria, based on one set of input data.
These criteria are (a) surface crack initiation, (b) interior crack initiation
(applicable to coated metals), and (c) complete fracture of thermal-
mechanical fatigue specimens, both unnotched and notched. Including
these three criteria in a single computer program produces a very general
and complete analysis of thermal fatigue resistance.
A complete set of the equations used in TFLIFE is given in Ref 5,
together with derivations. Computer time for a typical problem is 8 to 17
s on a time-sharing system (IBM 360/67). Required storage capacity is ap-
proximately 12 000 words.
' The italic numbers in braclcets refer to the list of references appended to this paper.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
SPERA AND COX ON COMPUTERIZED FATIGUE LIFE PREDICTION 71
TYPICAL
USER-SUPPLIED
MAIN PROGRAM
"TFLIFE" SUBROUTINE
V I . OUTPUT TABLES
temperature: (a) pounds per square inch and Fahrenheit degrees, (d) grams
per square millimeter and Celsius degrees, or (c) newtons per square cen-
timeter and Kelvin degrees. SI (Systeme International) units can be specified
for the bulk of the output, irrespective of the input unit system. TFLIFE
then returns to the main program.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
72 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
temperature distributions and the strains which result from them. In this
case, a linear elastic analysis is usually sufficient, according to the hy-
pothesis of total strain invariance proposed in Ref 6. This hypothesis
states that, under conditions of thermal stress, the displacements in a
body are substantially independent of local inelastic strains. Thus, dis-
placements obtained by a linear elastic analysis can be used to calculate
mechanical strains which can later be divided into elastic, plastic, and
creep components. For a complex component, it may be necessary for the
main program to call a finite-element analysis subroutine. In all cases, a
local temperature and a local uniaxial mechanical strain are calculated
for each time increment. The system then calls TFLIFE a second time.
In Block III, an incremental stress-strain analysis is performed which
divides the mechanical strain cycle into elastic, plastic, and creep com-
ponents. This analysis is repeated until shakedown occurs, when the stress-
strain hysteresis loop becomes stable and repetitive. In this way, a stress, a
plastic strain increment, and a creep strain increment are calculated for each
time increment. Approximately two thirds of TFLIFE's computing time is
used in these calculations. Upon completion of Block III, the mechanical
behavior of the material at the potential failure location has been described
thoroughly in terms of temperature, stress, and strain. Fatigue damage
can now be calculated.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
SPERA AND COX ON COMPUTERIZED FATIGUE LIFE PREDICTION 73
Output Tables
In Block VI, the results of the life analysis are printed in the form of six
tables. Samples of some of these tables are shown in Fig. 2 for a thermal-
mechanical fatigue problem which will be discussed later. A table which is
not shown documents the applied cycle in terms of the mechanical
strainrange plus temperature, strain, and time limits. Figure 2a contains the
actual life predictions. These are presented for each of the three failure
criteria just discussed. This figure also contains damage fractions which
show the theoretical proportions of the damage associated with creep strain
and elastic-plastic strain for each predicted Ufe. The sum of the two damage
fractions is always unity. Also, interior crack initiation and complete frac-
ture are assumed to occur simultaneously in a thermal-mechanical fatigue
specimen because of the nominally uniform strain and temperature con-
ditions in the test section.
Figure 2b presents temperature, strain, and stress as functions of time
during heating and cooling. If the user wishes, this figure can be doubled or
tripled in length, resulting in smaller time increments and more detailed
data on the cycle. Figure 2c contains the monotonic tensile data for the
material as a function of temperature. This figure is constructed from the
actual input data using linear interpolation. The number, order, and
spacing of temperatures for which property data are given are arbitrary.
The cycUc hardening ratio in the footnote to Fig. 2c relates cycUc and
monotonic stress-strain data. This ratio is an optional scaling factor for
proportionately increasing or decreasing yield and ultimate tensile stress
values at all temperatures. It is specified by the user according to one of the
following three criteria: (a) If no cyclic stress-strain data are available, the
hardening ratio is unity, which represents neutral hardening. Also, unity is
the nominal value used by the program in the absence of a specified value.
(b) If partial cyclic stress-strain data are available (for example, at only one
temperature) the hardening ratio is adjusted until first-order differences
between calculated and observed stresses are eUminated. (c) If complete
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions author
74 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
1 I I I 1
1. FAILURE CRITERinN I STRAIN I CYCLES I DAMAGE FRACTIONS 1
1 1 rnNC 1 rn i . . 1
1 1 FACTOR 1 FAILURE I 1 1
1 I I I PLASTIC 1 CREEP 1
I
1 TIME, TEMP, 1 STRAIN I STRESS,
.1
1 MIN DEG r. 1 MECH. I PLASTIC j CREEP I M/SQ CM
J I 1.
1 o.ono Ii78 1 I I I
1 0.039 506 1 -.009570 I -.000000 I 0.000000 I -56118
1 0.078 f 555 1 -.008622 I -.00000"
1 0.117 565 1 -.007662 I - '
.^o3 I 0 . 0 0 2 2 5 5 I Ii2962
.^ I 0 . 0 0 9 3 6 5 I 0 . 0 0 2 7 9 2 I UOSkC
I u.0087110 I 0 . 0 0 9 3 6 5 I 0.0OJ317 | 3868li
I 1.290 i lli21 I 0.009138 I 0.009565 I 0.005828 I 36739
I I I I I I
1 I
TEMP, I YOUNG'S YIELD ULTIMATE TENSILE POISSON'SI THERMAL I
I MODULUS, STRESS TENSILE DUCTILITYI RATIO I EXPANSIONI
(.002), STRESS, I FROM RT,|
DEG K I N/SQ CM N/SQ CM N/SQ CM I PER DEO K|
.1 I
I I
200 I 18.7E 06 56000 65500 0.800 I 6.05E-06 I
269 I 18.3E 06 51000 60100 1 nnn
339 I IS.IE 06 li8075
1(08 I 17.9E 06
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
SPERA AND COX ON COMPUTERIZED FATIGUE LIFE PREDICTION 75
cyclic Stress-Strain data are available, they are used to calculate equivalent
monotonic data for a cyclic hardening ratio of unity.
Figure 2d is & summary of the empirical equations which are used in
TFLIFE to calculate creep and rupture times [7]. The coefficients are
calculated from test data by means of regression analysis. Most of the com-
monly used time-temperature parameters are included in these general
equations as special cases, either exactly or to a close approximation. In this
way, the equations in Fig. 2d represent an attempt to select the optimum
2 1
1 LOG(Z) - A • B«LOG(X) • C«X • D»X • E«Y Z - T l » TR I
1 T l • TIME TO 1 PCT CREEP, HR TR - TIME TO RUPTURE, HR |
-1 1
1 CREEP RATE - 0 . 0 1 / T l , HR X - STRESS/1000 N/SQ CM I
BETA
1 Y • (TEMP/1000 K) OR LOG(TEMP/1000 K ) , IF BETA - 0 |
1 1 1 1 1 1 1
1 Z 1 A 1 8 1 C 1 D 1 E 1 BETA 1
1 1 1 1 1 1 1
1 1 t i l l !
1 Tl 1 - 9 . 1 9 8 1 -1I.BB70 1 0.00000 I-.001569 I 23.791 I -1.00 j
time-temperature parameter for the available data. If only limited creep rate
data are available, simplifying assumptions can be made by the user so that
some of the curve-fit constants in the equation for time to 1 percent creep
are the same as those in the equation for time to rupture.
A final table (not shown) contains data which document the iterations
required to achieve convergence. Each iteration is a new calculation of the
stress and inelastic strain cycles. The stress at the end of the cooling cycle is
carried forward to start the incremental stress analysis for the next iteration.
A residual inelastic strain is recorded which is the sum of the plastic and
creep strains at the end of each iteration. (Both these quantities are set to
zero at the start of each iteration.) A residual inelastic strain of zero
represents a completely stable, repetitive stress-strain hysteresis loop. When
the residual strain becomes less than ten microunits, a curve-fit procedure is
used to accelerate convergence, and iteration is terminated after two ad-
ditional cycles.
Options
After output tables have been printed for one set of input data, there is a
minimum of three options which the user can choose to include in the main
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
76 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
program: (a) terminating the program, (b) starting the solution to a new
problem by reading a new set of data, or (i) varying parameters in the input
data which are already stored in the computer and starting the solution
to a modified problem. The third option enables the user to perform a
parametric study from a single set of input data. As mentioned pre-
viously, the operations in the main program are determined by the needs
of the user. The TFLIFE system, shown schematically in Fig. 1, includes
sufficient flexibility to satsify a wide variety of such needs.
Sample Results
The application of TFLIFE to different types of thermal fatigue con-
ditions will now be illustrated by showing some typical results for the
following two problems: (a) thermal-mechanical fatigue of bar specimens of
the tantalum-base alloy T-111 tested in vacuum [76] and (b) thermal-stress
fatigue of wedge specimens of the nickel-base alloy B-1900 tested in
fluidized beds [17,18]. Figure 3 shows the geometry of these two types of
specimens. The slot notch shown in Fig. 3b has a theoretical elastic stress (or
strain) concentration factor of 3.0 [19]. Predicted lives are given for un-
notched and notched bars of T-111 as a function of imposed strain range
a 64 CM
5/8"-18
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
SPERA AND COX ON COMPUTERIZED FATIGUE LIFE PREDICTION 77
a 02 CM RAD
a 25 CM--
raiOCM R
r0.07 CM R 1
10 CM
J
C,^
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
78 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
cycle time and edge radius for both the uncoated and aluminide-coated con-
ditions of the alloy.
/
// 'A'l\ T,
/ If \
.005 / If
tl '
500 ^
If A
MECHANICAL rZ^J
if
STRAIN, 0 il \\\\ 0
Em
il
// \\u
II ^2 )C DIAMETRAL STRAIN (INPUT)
1
-.005 "7
V
\\ ^ A X I A L STRAIN (OUTPUT)
-.010 1 1 ^>
0 1 2 3
ELAPSED TIME, t, MIN
FIG. ^—Typical thermal and strain cycles applied during thermal-mechanicalfatigue testing
of the tantalum-base alloy T-111.
to the stress analysis (Fig. 1, Block III). TFLIFE resolves this difficulty by
using an equivalent axial strain equal to twice the diametral strain plus an
equivalent elastic modulus equal to Young's modulus divided by twice
Poisson's ratio [5].
Figure 5 illustrates the stress-strain behavior calculated for both the
isothermal and in-phase thermal cycles with equal strain ranges. The
significantly different shapes of these two calculated hysteresis loops are
confirmed by test data [16\. Cyclic hardening ratios were selected according
to the criteria described previously for cases with partial cyclic stress-strain
data (see output tables). Two different cyclic hardening ratios were required
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
SPERA A N D COX O N C O M P U T E R I Z E D FATIGUE LIFE PREDICTION 79
60X10^
1422" K7
IN-PHASE THERMAL
CYCLE
ISOTHERMAL CYCLE,
1422° K
TRUE STRESS,
for these two cycles because the stresses in the in-phase thermal cycle were
found to be significantly larger than those in the isothermal cycle for
equal strain ranges. A hardening ratio of 1.43 was needed to correlate cal-
culated and observed stresses at Point A in the in-phase loop. A much
smaller factor, 1.08, matched the calculated stress amplitude in the iso-
thermal loop (Point B) to the measured value. These hardening ratios
reflect the complex cyclic hardening which has been observed in T-111.
For behavior of this type, an accurate estimate of the hardening ratio
is essential to accuracy in the prediction of creep damage and life.
Figure 6 shows the results of the life calculations for isothermal cycling as
a function of axial mechanical strain range. The failure criterion for these
tests was complete fracture of the bar specimen. Predicted behavior with a
strain concentration factor of unity is in good agreement with the un-
notched specimen data, as shown by the open symbols. This correlation be-
tween theory and experiment gives a measure of confidence in the ex-
trapolated behavior for lives greater than 1000 cycles. For the notched
specimens (closed symbols), predicted behavior is somewhat conservative
when the strain concentration factor is equated to its theoretical elastic
value of 3.00. The available data were best fit using an empirical strain con-
centration factor of 2.15. Life predictions based on this factor are given by
the dashed line in Fig. 6. In the absence of notched specimen data, the
theoretical elastic strain concentration factor would be used as input in
TFLIFE.
In Fig. 7, predicted and observed lives are compared for tests with in-
phase thermal cycling. Unlike the isothermal tests, the in-phase tests of T-
111 showed little notch effect. The TFLIFE predictions are in agreement
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
80 T H E R M A L FATIGUE O F MATERIALS A N D COMPONENTS
1422° K, 0.0067 Hz
MECHANICAL
STRAIN
RANGE,
STRAIN CONCENTRATION
FACTOR
,>-1.00(UNNOTCHED)
2.15 (EMPIRICAL)
3.00 (THEORETICAL)
.001
10 10^ 10^ 10* 10^
CYCLES TO FRACTURE, N,
FIG. 6—Comparison of predicted and observed fatigue lives for isothermal cycling of the
alloy T-Ul in the unnotchedand notched conditions (1422 K, 0.0067Hz).
MECHANICAL
STRAIN
RANGE. ^ ^
t£_ F ^ ^ 1 ^ STRAIN CONCENTRATION
"• FACTOR
LOO (UNNOTCHED)
\ 2.15 (EMPIRICAL)
3.00 (THEORETICAL)
.001
10^ KP 10*
CYCLES TO FRACTURE, Nf
FIG. 1—Comparison ofpredicted and observed fatigue lives for in-phase thermal cycling of
the alloy T-Ul in the unnotched and notched conditions (478 to 1422 K, 0.0067Hz).
with the data, whether the strain concentration factor is assumed to be 1.00,
2.15, or 3.00. This insensitivity to strain concentration is related directly to
an assumption in the creep damage model in TFLIFE. The assumption is
that interior creep damage is not affected by a surface strain concentration
but depends only on the nominal stress. Thus, the strain concentration af-
fects only the elastic-plastic strain component of damage. During in-phase
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
SPERA AND COX ON COMPUTERIZED FATIGUE LIFE PREDICTION 81
thermal cycling at relatively large strain ranges (lives less than 1000 cycles)
most of the damage results from creep (Fig. 2a). Therefore, the model
indicates that strain concentration has only a minor effect on the predic-
ted lives at these strain ranges for in-phase cycling. However, as the strain
range decreases below about 0.004, elastic-plastic strain damage becomes
dominant over creep damage and a significant notch effect is predicted
for this alloy. Figure 7 is an example of an extrapolation of data by
TFLIFE that indicates a notch effect at long life which is not present in
the short-life data.
1200
EDGE TEMP,
HEATING IN 1360° K BED
T, ICOO
— COOLING IN 590P K BED
800
600
1 2
J m3
ELAPSED TIME, t, MIN
FIG. 8—Typical temperature transients at the edge of wedge specimens of the alloy B-1900
during thermal-stress fatigue testing.
Strain inputs for thermal-stress fatigue problems are quite complex. Con-
sequently, Block II in the main program (Fig. 1) is much more complex for
thermal-stress fatigue than for thermal-mechanical fatigue problems. For
instance, in this B-1900 problem, thermocouple data taken at several points
in the interior of the wedge specimen are curve-fit in Block II for each of the
approximately 100 time increments during heating and the same number of
time increments during heating and the same number of time increments
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
82 T H E R M A L FATIGUE O F MATERIALS A N D COMPONENTS
COOLING IN
590" K BED
MECHANICAL STRAIN,
-.002
-.004
600 800 1000 1200 1400
TEMPERATURE, T, °K
80b(lO^
-950PK
during cooling, as described in Ref 18. These curve-fits are then used to
calculate the temperature and mechanical strain cycles at the specimen
edges.
Figure 8 shows typical edge temperature transients during heating and
cooling. Note that the length of the heating and cooling periods is not suf-
ficient for the temperatures to reach steady-state levels equal to the bed tem-
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
SPERA A N D COX O N COMPUTERIZED FATIGUE LIFE PREDICTION 83
590P TO 13MP K
EDGE RADIUS,
r,
MM
0.6
HEATING LO
& PREDICTIONS USING
COOLING TFLIFE
TIMES,
MIN
10 10^ 10?
CYCLES TO CRACK INITIATION
590P TO 1360P K
lOr—
I MM
— EDGE
RADIUS,
HEATING & r, \ OND-
COOUNG TIMES, — MM \ o ^
MIN O a6 CJ [ J t
— D LO
- PREOICTia
TFLIFE
1 1 \ 1
10 10^ lo' 10*
CYCLES TO CRACK INITIATION
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
84 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Conclusions
The TFLIFE computer program has been found to be a useful tool for
analyzing thermal fatigue data in terms of conventional metal properties
and for predicting the thermal fatigue life of metals. It contains models for
the major damaging phenomena which occur during a general thermal-
fatigue cycle, at least to a first-order approximation. Thermal fatigue lives
calculated using TFLIFE have been verified by comparison with a variety of
laboratory test data on different types of alloys. This computer program is
ready for more extensive evaluation by the technical community, not only
on additional laboratory specimens but also on structural components.
References
[1] Spera, D. A., "A Linear Creep Damage Theory for Thermal Fatigue of Materials,"
Ph.D. thesis, University of Wisconsin, Madison, 1968.
[2] Spera, D. A., "Calculation of Thermal-Fatigue Life Based on Accumulated Creep
Damage," NASA TN D-5489, National Aeronautics and Space Administration, Oct.
1969.
[3] Spera, D. A. and Grisaffe, S. J., "Life Prediction of Turbine Components: On-going
Studies at the NASA Lewis Research Center," NASA TM X-2664, National Aeronautics
and Space Administration, Jan. 1973.
[4] Spera, D. A. in Fatigue at Elevated Temperatures, ASTMSTPS20. American Society for
Testing and Materials, 1973, pp. 648-657.
[5] Spera, D. A., "Equations for Predicting Thermal Fatigue Life," NASA TM (to be
published). National Aeronautics and Space Administration.
[(J] Mendelson, A. and Manson, S. S., "Practical Solution of Plastic Deformation Problems
in Elastic-Plastic Range," NACA TN 4088, National Advisory Committee for
Aeronautics, Sept. 1957.
[7] Palmgren, A., Zeitschrift des VereinesDeutscherIngenieure, Vol. 68, No. 14, April 1924,
pp. 339-341. (English translation, NASA TT F-13460, National Aeronautics and Space
Administration, March 1971.)
[8] Taira, S. in Creep in Structures, N. J. Hoff, Ed., Springer Verlag, Berlin, 1962, pp.
96-124.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
SPERA AND COX ON COMPUTERIZED FATIGUE LIFE PREDICTION 85
[P] Basquin, O. H., Proceedings, American Society for Testing and Materials, Vol. 11,1910,
pp. 625,630.
[10] Manson, S. S., "Behavior of Materials under Conditions of Thermal Stress," NACA
TN-2933, National Advisory Committee for Aeronautics, 1953.
[U] Manson, S. S., Experimental Mechanics, Vol. 5, No. 7, July 1965, pp. 193-226.
[12] Coffin, L. F., Jr., Transactions, American Society of Mechanical Engineers, Vol. 76,
Aug. 1954, pp. 931-950.
[13] Robinson, E. L., Transactions, American Society of Mechanical Engineers, Vol. 74,
1952, pp. 777-781.
[14] Hoffman, C. A., Proceedings, American Society for Testing and Materials, Vol. 56,
1956, pp. 1063-1080.
[15] Swindeman, R. V/., Proceedings, Joint International Conference on Creep, Institution of
Mechanical Engineers, London, 1963, pp. 3-71 to 3-76.
[16] Sheffler, K. D. and Doble, G. S., "Influence of Creep Damage on the Low Cycle Ther-
mal-Mechanical Fatigue Behavior of Two Tantalum-Base Alloys," NASA CR-121001,
National Aeronautics and Space Administration, May 1972.
[17] Howes, M. A. H., "Thermal Fatigue Data on 15 Nickel- and Cobalt-Base Alloys,"
NASA CR-72738, National Aeronautics and Space Administration, May 1970.
[18] Spera, D. A., Howes, M. A. H., and Bizon, P. T., "Thermal-Fatigue Resistance of 15
High-Temperature Alloys Determined by the Fluidized-Bed Technique," NASA TM X-
52975, National Aeronautics and Space Administration, March 1971.
[19] Manson, S. S. and Hirschberg, M. H., "Low Cycle Fatigue of Notched Specimens by
Consideration of Crack Initiation and Propagation," NASA TN D-3146, National
Aeronautics and Space Administration, June 1967.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
M A. H. Howes'
A Study of Thermal
Fatigue Mechanisms
KEY WORDS: thermal fatigue, fluidized bed processing, thermal cycling tests, nickel
alloys, steel, crack propagation, plastic deformation
86
Test Procedure
The equipment used for thermal fatigue evaluation consists of two flu-
idized beds, one heated and one cooled. The specimens are cycled between
the two beds at preset intervals. A full description of the equipment is
contained in an earUer paper [1].^
The specimens are disk-type configurations 16.35 cm diameter and 1.27
cm thick. The fin section is approximately 3 mm thick. These specimens
are used either unnotched or notched [1]. A notched specimen has two V-
notches of the same dimensions as the notch in a Type A Charpy impact
bar at 180 deg. The stress intensification factor of this type of notch is ap-
proximately 2.8.
Thermal fatigue testing often requires attempting the duplication of
both the stress and thermal cycles that occur in an actual component situ-
ation. In this investigation, it was desired to find the range of stress vari-
ation that could be produced with the same thermal excursion limits. This
would be done by varying the fluid bed test conditions, making accurate
temperature measurements throughout the thermal fatigue specimen test
section, and using a computer program to calculate the stress gradients.
The stress levels would then be compared with the thermal fatigue results
to determine if the stress level to which the specimens are subjected causes
cracking. Test results from other programs would also be examined where
mechanistic effects had occurred.
The specimens used for the stress det3rminations were made from two
materials chosen to be representative of steels requiring a hardening and
tempering treatment and materials requiring solution treatment and aging.
The compositions are shown in Table 1.
The materials were processed in the following way: the starting point
for each material was a 15-cm-square billet. From this billet, transverse
samples were removed at or near the midradius position for tension test-
ing. A portion of the billet was press forged to a 7-cm-round bar and used
for thermal fatigue specimens. The remainder of the 15-cm billet was re-
duced by forging and rolling to a 2.5-cm-diameter round bar for use in
creep, strain tempering, and static tempering tests. In this way, all mate-
rial of one composition came from the same billet, ingot, and heat.
The mechanical properties of both materials were measured up to 700 °C
and are shown in Figs. 1 and 2. Creep data are given in Fig. 3 and 4. Tem-
pering and aging curves are shown in Figs. 5 and 6.
The specimens were all thermally cycled in a manner that caused the
^The italic numbers in brackets refer to the list of references appended to this paper.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
88 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Analysis, %
C 0.399 0.018
Mn 0.33 <0.I0
P 0.018 <0.01
S 0.003 0.007
Si 0.93 <0.01
Cr 5.27 ND
Mo 1.13 5.10
V 1.00 ND
Ni 0.17 18.2
W <0.01 ND
Al ND° 0.10
Ti ND 0.65
Co ND 8.8
Zr ND <0.01
B ND 0.003
Melting conditions air melted induction vacuum melted
and vacuum arc melted
outer periphery of the fin to fluctuate between 593 and 204 °C. However,
the temperature excursion was produced in two different ways.
1. A severe heating cycle was produced by immersing the specimens in
a hot bed at 1038 °C and removing them when the outer edge of the fin
had reached 593 °C.
2. A mild heating cycle was produced by immersing the specimen in the
hot bed at 649 °C and removing them when the outer edge of the fin
reached 593 °C. In addition, a lower degree of bed fluidization was used.
The thermal fatigue test conditions are given in Table 2. It is seen that
there is a considerable difference in total cycle time between the two cy-
cles.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
HOWES ON THERMAL FATIGUE MECHANISMS 89
100
75
200
<
50 S.
IS
150 -S
25 O
2000
100
1500
£
8
£ 1000
500
- j _ - i -
100 200 300 400 500 600 700
Test Temperature , ' C
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
90 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
100 -1250
75 I- /
(9
S* 50 -o"-2?
<
0.«lt O / 1
150 S
25
RA.
2000 100
-'^%
1500
1000
500
_1 I L.
0 100 200 300 400 500 600 700
Test Temperature , 'C
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further rep
HOWES ON THERMAL FATIGUE MECHANISMS 91
1000
TIME, HOURS
10 100 1000
TIME, HOURS
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
92 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
H 13
45
o
-J 4 0
UJ
538*C
u
o
"^35
Ul
§30 \
< 566 °C
^ 593° C
25 -
20 1 1 1 1 1 - 1
3 4 5
I8NI
55
o
^ ^
ll 50
UJ
^ - - ^ ^ ~ - _ sse'c
0= 4 5 ^ ^ - - ~ - 566°C
tf)
^ V.^
UJ ^*-~ 593° C
§40
<
I
35
20
3 4 5
TIME, HOURS X 100
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
HOWES ON THERMAL FATIGUE MECHANISMS 93
LOCATION OF
THERMOCOUPLES
IN TEST SPECIMEN
TIME. SECONDS
FIG. 7—Heating and cooling of HI 3 thermal fatigue specimen fin during the severe cycle.
shows the mild cycle where temperature changes are much less severe. In
many cases, the temperature changes at the top edge of the fin are identi-
cal to changes at the bottom edge.
The temperature data are not of prime interest in themselves but are
obtained in order to predict stress levels in the specimen. This can be done
using finite element technique to calculate both thermal and stress gradi-
ents throughout the specimen.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
94 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
N
CM
o
*
<M
•si
ti
S'
o §
CM
s
«j
•s:
o«
O •^
CO 3
t3
— .C
-^
« S
u o
o
UJ
?
i
i?
Si
en 4
c:
O lU •^
CM •s •Si
H
8
e
f
o
3 930 '3aniVy3dW31
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reprod
HOWES ON THERMAL FATIGUE MECHANISMS 95
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorize
96 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
^I^
^
3
•Sf
is
a
-s:
-4^
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
HOWES ON THERMAL FATIGUE MECHANISMS 97
500
f c
i sV co
1
Z 1^
400 o
1 ^
'
300 -
200
100
r
9 ^ ^ 2 or 6
0 7^
s. > _-,-^"^ • ^
K
\
-100 ^
1
^
-200 3or7
^
-300 - \
2ar6
1 ^/
-400 -
\ /
-500 - \ lor 5
— ttcin 1 1 1 1 1 1 1
10 15 20 25 30 35
Time, Seconds
FIG. 10—Stress cycle of HI3 specimen fin during the severe cycle.
elastic strain but at a stress level near the point of plastic working. A
possible explanation is that the stress concentration at the notch was
higher than expected.
During testing, additional specimens were cycled and removed at inter-
vals until sufficient data had been obtained to observe the hardness changes
across the fin section. Specimens were removed at 250, 500, 1000, 2000,
4000, 8000, and 16 000 cycles. The specimens were sectioned and micro-
hardness surveys performed on the fin. These hardnesses were converted
to Rockwell C values, and isohardness curves were drawn across diagrams
representing the fins. The information is summarized in Fig. 14. This
figure also gives an estimate of the equivalent time that the outer edge of
the fin has spent at 593 °C. The estimates were made from tempering
curves by summing the effects of small temperature increments from 204
to593°C.
If the change in hardness (Fig. 14) is compared with the tempering
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions autho
98 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
lor S
. 1
2 or 6
TV®
-..JorT ^ ^ 5 v
(/'
l/fiorS
VL 4ore
h) '°
1 3 or 7
1——-^
2 or 6
-
\^rl^ , 1.- '
100 ISO
Tim* Sftconds
TABLE 3—Maximum stress levels at the surface of the thermal fatigue specimens.
Stress, MPa
Ultimate
Severe Cycle Mild Cycle Tensile 0.2%
Strength Yield
Alloy Plain Notched Plain Notched at593°C at593°C
" Values in parentheses are for specimens tested after a 200 h exposure at 593 °C. 200 h is
equivalent to 25 000 mild cycles or 3.2 x 10' severe cycles for H13 and 28 500 mild cycles or
3.9 X 10' severe cycles for 18Ni maraging alloy.
curves in Figs. 5 and 6, it will be seen that stress acts to accelerate the
normal tempering changes in the structure. In the case of the 18Ni marag-
ing alloy cycled under the severe thermal cycle, there is evidence that
strain hardening occurs in the body of the fin due to the high stresses im-
posed in this cycle. The effect observed in the HI3 tool steel and in 18Ni
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized
HOWES ON THERMAL FATIGUE MECHANISMS 99
11 /
17*
If
/ /
//' /;
fi
1 / / / , 1 // 1
15,000 20,000 2S,000 0 SpOO
NUMBER OF THERMAL CYCLES
FIG. 12—(a) Crack growth of successive cracks in one unnotched HIS specimen, (b) Aver-
age crack growth of four cracks in HIS and ISNi notched specimens. Specimens cycled be-
tween 59S and 204 °C under the severe conditions.
0 5,000 10,000
Number Of Thermal Cycles
FIG. 13—Average crack growth of successive cycles in four notched HIS and JSNi marag-
ing alloy specimens. Specimens cycled between 59S and 204 °C under the mild conditions.
maraging alloy exposed to the mild cycle is one of failure produced by the
onset of cracking in plastically deformed material.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
100 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
THERMAL CYCLES
18 Ni
SEVERE Q r \ r Rc53 1 Llnc53_] L_R||3_J |1_RC53_| |_Rc53_|
54
THERMAL
CYCLE
I H.
Equivalent Exposure 53 53 53 53
(Hours) (0.013)
).0i3) (0.025) (0.05 (Ol)
Ujl.A>.A>-l>AA.AJk,J LJ„K>.JUlJiJl,J'.*~AJ^ |JKJkAJULAA.A.AAAJ
18 Ni Rc50 \
MILD
THERMAL
CYCLE
Equivalent Exposure 49 47
(Hours) (1.75) (3.5)
FIG. 14—Hardness changed in thermal fatigue specimen fins after cycling 250 to 16 000
times.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized
HOWES ON THERMAL FATIGUE MECHANISMS 101
*^Tj:^ t - ^ - 4^-*^'
M^ ^ ^ > . . . ^ ^
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
102 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
FIG. M—Effect of structure due to different quench rates in steel specimens on crack
initiation during thermal fatigue testing. Specimens cycled between 550 and 260°C.
made to work to improve thermal fatigue resistance, and the prime ex-
ample of this is directional solidification and single crystal structures.
Great improvements have been reported when directional solidification
can be used in a cast part [5-7].
Even if it is assumed that a particular composition is processed to pro-
duce the structure with the highest thermal fatigue resistance, there is
another effect to be considered. During thermal exposure there is often a
degradation in properties and potential crack sites can also develop.
Figure 18 shows a section through a double-edged wedge IN-100 speci-
men that has been subjected to thermal cycling between 1088 and 316°C
[5]. It can be seen that the alloy is subject to oxidation and a surface layer
of material with a lower content of carbon, chromium, aluminum, and
titanium can be produced which is thus of lower strength than the base
material. It is less able to resist the stresses produced by the thermal cycUng.
Furthermore, at the points where the grain boundaries meet the surface,
the effect is intensified, and the oxidized or weakened layer tends to pene-
trate further at these points. This has the effect of creating notches or
stress intensification points at the surface, and, when oxidation has pro-
ceeded far enough, a crack will initiate and follow the line of least resis-
tance in the structure. The success of coatings in improving thermal fa-
tigue resistance is largely due to their ability to retard oxidation.
This is one reason why there is an apparent incubation period before
thermal cracking commences, but another reason is the degradation of the
material properties themselves. Most materials are strengthened by a mar-
tensitic or an age hardening mechanism. Thus, during thermal cycling,
there is a tendency to temper the structure or to cause overaging and
thereby lower the strength of the material. The decrease in surface hard-
ness of twelve superalloys during thermal cycling has been reported [6],
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
HOWES ON THERMAL FATIGUE MECHANISMS 103
S9
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
104 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
and this almost certainly indicates that the strength of the material at the
surface has been reduced. A further effect is that tempering and over-
aging are accelerated by stress [8], and thus the degradation of properties
occurs faster while the material is being thermal cycled.
Summary
To summarize the events leading to a thermal fatigue failure, it has
been shown that failures occur because the materials have been stressed in
the plastic range. If the stresses are purely elastic, then failure would only
normally occur after alignment of dislocations during an incubation period
causes a void which would exceed the critical crack length and initiate a
self-sustaining crack.
Under high stress levels or in situations where stress concentration ef-
fects cause a crack to propagate, crack initiation occurs at the start of
thermd cycling or very soon thereafter. However, the common experience
during thermal fatigue testing is that a definite incubation period is re-
quired before cracking commences. It is believed that a number of rate-
dependent processes can be in progress during this period, either singly or
concurrently, and it is this factor that makes the development of predic-
tive equations very difficult.
One process that occurs is plastic deformation, and it is fairly common
to see considerable local plastic deformation on the surface of a specimen
before cracking commences. When cracking does occur, it is almost in-
variably in the areas where plastic deformation was the highest, and this
may be regarded as an exhaustion of ductility effect. A further factor is
that the strength of the material will be slowly degraded due to overaging
or tempering effects, and this will be stress accelerated. At points of max-
imum stress, it is likely that plastic deformation will occur to a greater ex-
tent than in the bulk of the section, and the mechanical properties will
therefore be degraded faster in these areas. As the mechanical properties
decrease, plastic deformation becomes easier, and thus the weakening
process is self-accelerating in individual small areas which will ultimately
provide the sites for crack initiation.
A second mechanism that occurs is oxidation depleting the surface of
materials at the higher exposure temperature. A nickel-base superalloy
can expect to be depleted of carbon, chromium, aluminum, and titanium,
at least. Depletion will often be preferential at the grain boundaries, and
wedges of low strength material will be formed which act like a notch.
After some finite time, the notch will assume the dimensions of the criti-
cal crack size, and a thermal crack will initiate and propagate.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
HOWES ON THERMAL FATIGUE MECHANISMS 105
Acknowledgments
The author is grateful to Kin Yeung, Edward Welch, and Anne Hum-
phreys who developed the computer program for predicting stresses under
thermal cycling conditions.
References
[1] Howes, M. A. H. in Fatigue at Elevated Temperatures, ASTM STP 520, American So-
ciety for Testing and Materials, 1976, pp. 242-254.
[2] Farhoomad, I. and Wilson, E. L., "Nonlinear Heat Transfer Analysis of Axisymmetric
Solids," Report PB 199169, University of California, Berkeley, 1970.
[3] Farhoomad, I. and Wilson, E. L., "A Nonlinear Finite Element Code for Analyzing the
Blast Response of Underground Structures," Report No. N-70-1, University of Califor-
nia, Berkeley, 1970.
[4] Beck, C. G. and Santhanam, A. T., this publication, pp. 123-140.
[5] Howes, M. A. H., "Additional Thermal Fatigue Data on Nickel- and Cobah-Base
Superalloys," NASA CR-121212, National Aeronautics and Space Administration,
Washington, D.C., 15 March 1973.
[6] Howes, M. A. H., "Thermal Fatigue Data on 15 Nickel- and Cobalt-Base Alloys,"
NASA CR-72738, National Aeronautics and Space Administration, Washington, D.C.,
15 May 1970.
[7] Howes, M. A. H., "Thermal Fatigue and Oxidation Data on TAZ-8A, MAR-M200, and
Udimet 700 Superalloys," NASA Cr-134775, Washington, D.C., 15 Jan. 1975.
[8] Howes, M. A. H., "Investigation of Stress-Accelerated Tempering and Aging of Fer-
rous Alloys," presented at XVI International Heat Treatment Conference, Stratford,
England, May 1976.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
p. T. Bizon^ andD. A. Spera^
KEY WORDS: thermal fatigue, thermal stress, oxidation, coatings, nickel alloys,
cobalt alloys,fluidizedbed, heat resistant alloys, mechanical properties
106
Alloys
The 35 combinations of composition, grain structure, and surface protec-
tion which were studied in this program are listed in Table 1. Table 2 lists
the compositions of the dloys (chemical analyses provided by vendor or in-
dependent laboratory of heats used for specimens) along with the various
heat treatments applied to them. Mechanical properties are given in Table 3
for the same heat of alloy that was used for the thermal-stress fatigue
specimens.
A new alloy included in this investigation is directionally solidified NASA
TAZ-8A. In an earlier investigation [4], the random polycrystalline form of
this alloy was found to have the highest resistance to thermal-stress fa-
tigue among 14 uncoated alloys. This indicated that directional soUdifica-
tion or coating or both of TAZ-8A might produce a material with even
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized
108 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
NOTE—DS itidicates that the alloy was cast with a directionally solidified
grain structure.
"Specimens cast at NASA Lewis Research Center.
' Specimens with and without coatings supplied by Pratt & Whitney Air-
craft Corp.
'Specimens supplied by General Electric Corp.
•*Specimens supplied by Cabot Corp.
Surface Protection
Brief descriptions of the four protective coatings used in this study are as
follows:
Jocoat—This is a commercial silicon-modified nickel aluminide coating
(Pratt and Whitney Aircraft (PWA) proprietary process specified as PWA
47).
RT-IA—RT-IA is a commercial chromium-aluminum duplex coating
(Chromalloy American Corporation Research and Technology Division
proprietary process similar to specification PWA 32 but with a lower
process temperature).
RT-XP—RT-XP is a coating containing an aluminide with a case depth
of about 70 Mm (2.7 mils) (Chromalloy American Corporation Research
and Technology Division proprietary process).
NiCrAlY—This is a commercial Ni-15.2Cr-12Al-0.33Y electron-beam
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions a
BIZON AND SPERA ON SUPERALLOYS 109
vapor-deposited overlay coating about 135 /im (5.3 mils) thick (PWA
proprietary process specified as PWA 267).
Test Specimens
In this study, two types of specimen configurations (Fig. 1) were used:
bars with double-wedge cross sections and round cross-sectional tensile
bars. No significant differences were observed in the results of duplicate
tests of the double-wedge shape specimens whether end grooves or holes
were used for fastening them into the supporting fixtures.
Figure 1 also shows the designs of specimen used to measure conventional
tensile and stress-rupture properties. Designs A and B are similar except
that Design B has flat grips so that it can be fabricated from 0.635-cm (0.25-
in.) plate. The only difference between Designs C and A is the longer gage
length in Design C.
All cast alloy specimens were cast to size. For the random polycrystalhne
specimens, inoculated molds were used to produce fine grain structures.
Typical surface grain size at the test sections was about 1.6 mm (0.06 in.)
diameter. The directionally solidified polycrystalhne specimens were made
using a controlled solidification process similar to that detailed in Ref 16.
Only directionally solidified specimens with no grain boundaries in-
tercepting the leading edge were tested. Specimens of the two wrought
alloys (U 700 and TD NiCr) were machined with the specimen axis parallel
to the rolling or extrusion direction. All specimens were given radiographic,
fluorescent penetrant, and visual inspections before testing.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions au
110 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
M J= -C J= i' i-
«N •» '9- NO ri
t. 1- fc
0 5 ^ 0 -c 0 -c
t^ t— 4— 0
yd 8 0
NO
3—
? ?—
| Uh U
r — 1^1
I
^O 0 rr r- C- r. 3 o ^ O
00 (8 —• '-^ nj — ^ S3 "
flj r- 00 «
>
I 0
0
V
d 0
X
V 5 z £ £
5IN dS 0S -0
Si
£!?
s:
o r^ d d
Ia
a o .0
— VO •^
(N O
-: d 0 0 0 0
— . f^ •.s.
ID <*> I d « : S . — (N
00 «X
NO r~ O f^
^ p3 —
00 ~" I '
i
fo d «
00 X) X>
S Xi ^ JS Xt XI O -C -C X>
Ills
•3
X
8 ?
NO d
I -O "O 00 O ~-
§
-s: 0 0 0 0
•xs
• m
d d d
s d
'^ d d
~
d d d
Ic V V 9 V V
.0 0 0 r^ 0
• 0 0 ™ 0 0 0
d d d d d d d d d
V
V 0 V
I — 00 0
0 •t
0
0 i - 3 0
- r- •0 r~ ^ 00
NO 00
0 0 0 0 d d d d d d d d d d d d d d d
<3
I
W
_! ss 3 3::^
00
VO
io
no •V Q X U > f^ \ o «8 m
u-
X sa
< i- iii al
(N t
—• t o
^
*
00
— :s H u
£
< N <
i S +
u
S i
ss <
Is s S 8 | s i < 8 3 z 2
r-
00
s^ Xz ^<z II
(/5 O 2R —
7:
2 Z OS < * - z Q 2
Z X M S H
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
BIZON AND SPERA ON SUPERALLOYS 111
li
^^^
-S3 2^u S
8 s^S" 5-* «o g •-
o ^ C- •«
^1
m iU
s •
2 a!
o o
- s
s RS
8
3
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
112 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
U 8 •&. (S w^ vo O I
S o s sOTf\oa>»^r-)Ooo
S «" —
= 'vo (STfrf^O^ — &<rof0^r-«*>w^O
r- Z H
Si = i . •/% iC «^ 00 2 ^ ^ 00 —• IN -00
is 3
9 • S3 ? « Ss g - °
ih I <S <N —• — I S IN I — — M
• ^ o v o—
o —
<N
O w ^ ^ o—
<N
o f—
^tO
(S
'CsO
^ s
r ^ ^ O 0 \ O ( N ^ P ^ O ^ « N C T \ r ~ r - —
oo rr . ^O to (N <N VO <S m r- r-
\o — <= 0\ O lO «o
oo
. »i->
• 00 (S
—
—.—
Ov00 00 <s oo «-.» <Nr- ~^
fS 00 00
s2
• (*i
«
if!! m
• ^
:c QO S OS OS
fM
o • (S 00 f*^
M-1 00
w-1 QO 00
<o - U-1
OdOv'*lf~* — — O
J8
li.
I 00 — ^ > 0 " o O » * 1 0 \ ' ^ ^
— (NO-^-^r- — o o ^ > 0 ' « Q o o
I
mr^xn^OOsooMfS
S O i O O O t f s C S i n O — rrt O » o S - - r - - O O ' O Q 0 O t - - * O O '
BQ
3|
iili (N>*
o o
.OO^CTitOTf
-r-oo-^-* — -•I
— 1 roro
Oro
•-TOOf^oo'.O'O-g-v-ifNQ^
Or-OsOsr^t— OisOOOOi
• r- • 'O 00 . vo 1^ I
M«il
O u ^ s p v ^ O Q - O - O O M f S l N
• r- • * 00 • — o 1 S O O O O ^ O — O -00 • S 0 0 \ 0 0 v
OOr'^tri(NfnO'Oio — QO'*v-i'«tO
- H T f O m O — — • ^ O O Q O — <N —
O O 00 m I s p *0 OS r o '
Os TT t ^ OS I
• oooor-r^Of^osoocs
en
•^ i? S
*^ 8 <• X N s <<
<
CO = i s yz
? i 81 ^< +8 H<o or. ^ o f^
• < - < * - _ Q §l8
I z m zX m S H S2zzz«So< S D S S s D
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
BIZON AND SPERA ON SUPERALLOYS 113
i
•111
11 a
{|ii!L
iM^
f
pIlHIi
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
114 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
rl.OO
a 395)
L440
( 0 567).
© ^-RADIUS.
CLEARANCE
DIAMETER
7.«
(iOO)
r -o 9.53
a 75)
iai6
L —
0 6 4 ( 0 25)
DESIGN D L
A 0643 Z03
0.«3a375^.^ a 00)
(0253) (0801
RADIUS, B 0500 a 03
a318(ai25l (0197) (0 8 0
L440
C 0643 386
j Lj'-vP
o ( 0 567),
047
(0 253) (152)
( 0 19)-
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
BIZON AND SPERA ON SUPERALLOYS 115
i. ..Sw J
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions author
116 THERMAL FATIGUE O F MATERIALS A N D COMPONENTS
POSITIONED BY
PNEUMATIC
ACTUATORS-.
ALUMINA RETAINER
HEATING BED
123 CM DIA)
ALUMNA PARTICLES
REFRACTORY ^ -SPECIMENS
INSULATION- •'
r INSULATION
INSPECTION PORT-
RETORT
THERMOCOUPLE COOUNG AIR
AIR
l€AT EXCHANGER
INSPECTION PORT
A!R»
FIG. 3—Fluidized bed test facility at Illinois Institute of Technology Research Institute.
longer lives), duplicate or triplicate tests were run. The number of cycles to
crack initiation was taken as the average of the number of cycles at the last
inspection without cracks and the number of cycles at the first inspection
with a crack.
Thermal-Stress Fatigue
The numbers of cycles required to initiate cracks in the 35 combinations
of alloys and coatings are shown in Fig. 4. All results are for the 0.635-cm
(0.25-in.)-radius edge only. For the same specimen geometry and test con-
ditions, the lives varied from less than 25 to 12 500 cycles. Reproducibility
of test data was generally good.
The class of materials having the longest thermal-stress fatigue lives was
found to be cast alloys with directionally soUdified polycrystalline grain
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorize
BIZON AND SPERA ON SUPERALLOYS 117
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
118 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
10000
FIG. S—Comparison of thermal-stress fatigue resistances for alloys tested at both bed tem-
perature conditions.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authoriz
BIZON AND SPERA ON SUPERALLOYS 119
Oxidation
The materials exhibited a wide variation in oxidation resistance. In
general, a coating or vapor-deposited overlay on an alloy gave a greatly
reduced weight loss. Without surface protection, the directionally solidified
alloys oxidized much faster than the same compositions in the random
polycrystalline form [8], The severe oxidation of an uncoated directionally
solidified alloy can be seen by referring to Fig. 6b which shows direc-
tionally solidified IN 100 before testing and after testing for 2400 and
2900 cycles at bed temperatures of 1130 and 357 °C (2065 and 675 °F).
Figure 7 shows a comparison of the weight changes for directionally
solidified TAZ-8A and Mar-M200, the alloy with the next highest thermal-
stress fatigue resistance. All data are for double-wedge specimens tested at
bed temperatures of 1088 and 316°C (1990 and 600°F). Results are pre-
sented for both alloys without a coating and with the NiCrAlY overlay
coating. In addition, data for TAZ-8A with the RT-XP coat are also pre-
sented. The much higher weight loss of Mar-M200 compared to that of
TAZ-8A in the uncoated condition is apparent. The effect of applying
overlay coatings to both alloys is to greatly improve their oxidation re-
sistance. Note that TAZ-8A with the RT-XP coating exhibited only a Vi
percent weight loss after 14 OCX) cycles of testing, indicating excellent re-
sistance to oxidation.
Summary of Results
The comparative thermal-stress fatigue resistances of 26 nickel- and
cobalt-base alloys were determined using the fluidized bed technique. A
total of 35 combinations of compositions, grain structure, and coatings
were studied. All materials were evaluated by thermally cycling double-
wedge specimens between two fluidized bed furnaces maintained at 1088
and 316°C (1990 and 6(X)°F). Some materials were also exposed to bed tem-
peratures of 1130 and 357 °C (2065 and 675 °F). Immersion times were 3 min
in both the high and low temperature beds. Thermal-stress fatigue resist-
ance was based on the number of cycles required to initiate a crack. The
major results obtained are as follows.
1. Thermal-stress fatigue lives under identical test conditions ranged
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions autho
120 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
OtSCSIPTIOH, imCOATtB
cvcusroFiBsrMACK. ms
C V O i ^ TO PHOroSRAPH,. 0
iaiMA(t-Mmtis*aoy.
25 nm
iWINtOOCSftlUW.
FIG. 6—Directionally solidified thermal-stress fatigue specimens before and after testing at
bed temperatures of 1130 and357°C (2065 and 675 °F).
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
BIZON AND SPERA ON SUPERALLOYS 121
tEO TEMPS.
•---. C F
31^ lUft
1 THREE-MINUTE
IMMERSION IN
EACHIED
FIG. 7—Weight change for two directionally solidified alloys from fluidized bed tests.
References
[/] Spera, D. A. and Grisaffe, S. J., "Life Prediction of Turbine Components: On-Going
Studies at the NASA Lewis Research Center," Technical Memorandum X-2664,
National Aeronautics and Space Administration, 1973.
[2\ Bizon, P. T.'and Oldrieve, R. E., "Thermal Fatigue Resistance of NASA WAZ-20 Alloy
With Three Commercial Coatings," Technical Memorandum X-3168, National
Aeronautics and Space Administration, 197S.
[J] Spera, D. A., Howes, M. A., and Bizon, P. T., "Thermal Fatigue Resistance of 15 High-
Temperature Alloys Determined by the Fluidized-Bed Technique," Technical Memoran-
dum X-52975, National Aeronautics and Space Administration, 1971.
[4\ Howes, M. A., "Thermal Fatigue Data on 15 Nickel- and Cobalt-Base Alloys," IIT
Research Institute, Report IITRI-B6078-38, Report CR-72738, National Aeronautics and
Space Administration, 1970.
[J] Howes, M. A., "Additional Thermal Fatigue Data on Nickel- and Cobalt-Base
Superalloys," IIT Research Institute, Report IITRI-B6107-34 (Part 1), Report CR-
121211, National Aeronautics and Space Administration, 1973.
[6\ Howes, M. A., "Additional Thermal Fatigue Data on Nickel- and Cobalt-Base
Superalloys," IIT Research Institute Report IITRI-B6107-34 (Part 2), Report CR-
121212, National Aeronautics and Space Administration, 1973.
[7] Howes, M. A., "Thermal Fatigue and Oxidation Data on TAZ-8A, Mar-M 200, and
Udimet 700 Superalloys," IIT Research Institute, Report IITRI-B6124-21, Report CR-
134775, National Aeronautics and Space Administration, 1975.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
122 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
[8] Bizon, P. T. and Spera, D. A., "Comparative Thermal Fatigue Resistances of 26 Nickel-
and Cobalt-Base Alloys," Technical Note D-8071, National Aeronautics and Space Ad-
ministration, 1975.
[P] Glenny, E., Northwood, J. E., Shaw, S. W., and Taylor, T. A., Journal of the Institute
of Metals, Vol. 87, 1958-9, pp. 294-302.
[10] Glenny, E. and Taylor, T. A., Journal of the Institute of Metals, Vol. 88, 1959-60, pp.
449-461.
[11] Franklin, A. W., Heslop, J., and Smith, R. A., Journal of the Institute of Metals, Vol.
92, 1963-4, pp. 313-321.
[72] Howes, M. A., Die Casting Engineering, Vol. 13, 1969, p. 12.
[13] Howes, M. A. and Saperstein, Z. P., Welding, Vol. 48, 1969, pp. 543-545.
[14] Rostoker, W., Journal of Materials, Vol. 4, No. 1, 1969, pp. 117-144.
[75] Mowbray, D. F., Woodford, D. A., and Brandt, D. E. in Fatigue at Elevated Tem-
peratures, ASTM STP 520, American Society for Testing and Materials, 1973, pp.
416-426.
[75] Freche, J. C , Waters, W. J., and Ashbrook, R. L., "Application of Directional
Solidification to a NASA Nickel-Base Alloy (TAZ-8B)," Technical Note D-4390,
National Aeronautics and Space Administration, 1968.
[77] Howes, M. A. in Fatigue at Elevated Temperatures, ASTM STP 520. American Society
for Testing and Materials, 1973, pp. 242-254.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
C G. Beck^ and A. T. Santhamm^
ABSTRACT: A study was undertaken to determine whether the thermal fatigue re-
sponse of the cast cobalt-base alloy, Mar-M509, could be modified by changing the
casting parameters. Single-edge wedge-type specimens were tested in a fluidized bed
using a thermal cycle of 1915/52S''F with a cycle time of 6 min divided equally be-
tween heating and cooling. The results revealed that the thermal fatigue life of Mar-
M509, from the point of view of both crack initiation and propagation, could be sig-
nificantly improved by changing the casting variables. Cracking was found to be pre-
dominantly transgranular and often initiated at interdendritic carbides. A good cor-
relation was obtained between the secondary dendrite arm spacing and the number of
cycles for crack initiation, as well as the crack propagation rates. Small dendrite
spacings reduced the thermal fatigue life by accelerating both the crack initiation and
propagation. Precipitation of MC and MuCt carbide occurred during the thermal
cycling and affected the crack propagation rate. Based on these results, a model for
thermal fatigue cracking in Mar-M509 is presented.
In gas turbines, critical components such as blades and guide vanes are
subjected to rapid changes in temperature during startups and shutdowns.
The outer layers and thinner sections of the components respond more
quickly to changes in temperature than the interior. The resulting thermal
gradients give rise to thermal strains and stresses. Repetition of the stress
cycles during service can sometimes cause cracking of the component. This
type of cracking under fluctuating temperature conditions is termed ther-
mal fatigue. The susceptibility of blade and vane materials to thermal
fatigue is therefore a matter of vital concern to the designer.
123
Experimental
The cobalt-base alloy, Mar-M509, is presently used as one of the guide
vane materials for land and air gas turbines. The chemical composition of
the heat used is given in Table 1. Specimens were cast using different mold
and superheat temperatures which produced variations in grain size, den-
drite spacings, and amount of carbides. Single-edge wedge-type specimens
with geometry as shown in Fig. 1 were prepared from these castings.
The thermal fatigue testing was carried out in a fluidized bed test facility
at the Illinois Institute of Technology Research Institute (IITRI). It consists
of a high-temperature bed heated externally with silicon carbide elements
and a low-temperature bed equipped with an air- or water-cooled heat ex-
changer and an internal heat source of low power. The specimens are cycled
between the high- and low-temperature beds by means of automatically
controlled pneumatic cylinders. The bed consists of alumina particles in the
28 to 48 mesh size range. Air enters from the bottom and keeps the bed
fluidized. As a result, the rate of heat transfer to the specimen is high. The
heat content of the media is also high, so that a leuge number of specimens
can be heated and cooled rapidly without lowering the bed temperature
significantly.
A thermal cycle of 1915/525 °F (1046/274 °C) was used in the present
study. The specimens were immersed in each bed for 3 min. Thermocouples
were placed at five locations across the specimen section, and the tem-
peratures were recorded during heating and cooling. The time history of the
tip (thin trailing edge) temperature and that of the center of the specimen is
displayed in Fig. 2. Note that the tip gets heated and cooled more rapidly
than the center (Thermocouple Position 3) and gives rise to a transient tem-
perature difference which maximizes at short times. It is this transient tem-
^ The italic numbers in brackets refer to the list of references appended to this paper.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
BECK AND SANTHANAM ON A CAST COBALT-BASE ALLOY 125
o 13
U 03
m
<
ao
1
o
V
g
a. 1
d
V
3
<6
o V
a §
d
o
d
I H
00
d
Si
I d
^
m
n Z O
<
l-i en
o
1/1 8
d
K
i
d
V
u d
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
126 THERMAL FATIGUE OF MATERIALS A N D COMPONENTS
<zifn
.241
.231 K-.98-H
TW
3.94
.141,
.135'
(?)
.025"-| ..325'U».350",
00 I--040"
.030'|M n
270
K63< 282
1200
1
1
1 1000
800
(a) -
1 1 1 1 1
MO 0 60 120 180 240 300 360
' 1 1 1 I 1
+400
•»200 -
? 0
5 -200
(b)
1 1 1 1 1 1
-400
C 60 120 180 240 300 360
Elapsed Time, (sec)
FIG. 2—(a) Measured temperatures during rapid heating and cooling of wedge type
specimen of Mar-M509 and (b) transient temperature between the tip and the center of the
specimen.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
BECK AND SANTHANAM ON A CAST COBALT-BASE ALLOY 127
perature difference that gives rise to thermal strains and stresses and con-
sequent cracking of a component.
The specimens were removed from the test rig at regular intervals, and the
edges were examined for crack initiation using a x30 microscope. In-
spections were made at 25, 50, and 100 cycles, after every 100 cycles up to
2000 cycles, and at 2500 and 3000 cycles. When a crack was observed, its
length from crack tip to specimen edge was measured on both surfaces, and
the average value was taken as the crack length.
Superheat, "F
200 300 400
.10 I- ^2000 o 2.5
11900 ^
E
Sl800 o
.08 i0 3
i
I .06 1.5 2
c
I
.04 LO
.02 0.5
0 0
1000 2000 3000
Number of Cycles
FIG. 3—Crack length versus number of cycles for various casting parameters ofMar-M509.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproducti
128 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
.006
Superheat, "F
200 300 400
.005 I - **". 2000 a
t 1900 A
- 3
T .004
1 1800O
a
I
•s
.003
1 .002
.001
Numbers
0
1000 2000 3000
Number of Cycles
FIG. A—Area of crack versus number of cycles for various casting parameters ofMar-M509.
for the specimens tested. The thermal cycles required for crack initiation
were determined by extrapolating the curves to zero crack area (actually
0.001 in.^). Figure 4 demonstrates that the specimens representing the dif-
ferent casting variables differ in their response to thermal fatigue. There are
differences in both the number of cycles for crack initiation emd in the crack
propagation rate. Furthermore, crack propagation occurs in two stages.
(This is true for all the specimens except Specimen 2 in which only Stage 1
was observed within 3000 thermal cycles.) The first stage crack propagation
is faster than the second stage. Table 2 summarizes the crack initiation and
propagation data for each of the specimens. As can be seen in this table,
based on the measured cracked area, the thermal fatigue lives, from the
point of view of crack initiation, varied from 12(X) to 2100 cycles. The first
stage crack propagation rates differed by a factor of=1.5.
Microstructural Aspects
A microstructural study was performed on the specimens before and after
the thermal fatigue test with a view to identifying the structural features
associated with crack initiation and propagation. The specimens were
ground to the midplane, polished, etched, and examined by optical
microscopy.
In general, the as-cast microstructure of Mar-M509 consists of five
phases:
1. A face-centered-cubic (fee) cobalt-base matrix.
2. An interdendritic script-like MC carbide.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
BECK AND SANTHANAM ON A CAST COBALT-BASE ALLOY 129
p O 00 O "^
<N -^" O ~ ' —"
CO a E
22222
>n 00 VO t^ d
'^ o d o '^
OS
XX : X X
«S TF ( S 00
f o ooo o
X X X X X
1 f*^ f*l 00 rn vo
• ^ fn <S "*' ^*^
o o oo o
v~i trt *ri u^ v^
m Tt <s <^ -^
t-. -« (N rt -•
I R ^ ;8S
<s - ^ O
^ « (S ^ -^
fn >ri
a
a
<
Z < 5
iJi !S88i
I (S <N m '
ill! 800 S
^ ^
888
Ch O 00 00
(S ^ ^
*o Tt <N r- 00
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions
130 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
•'V'- -'
I %
E
bides. Very few cracks initiated at grain boundaries. The crack propagation
was predominantly transgranular and occurred in a direction normal to the
longitudinal axis of the specimen. This is illustrated in Fig. 7. Varying the
grain size from 0.5 to 2 mm did not affect either the crack initiation or
propagation as can be seen in Fig. 8.
Figure 9 shows the crack growth path after 3000 thermal cycles for
various soUdification conditions with associated variation in the secondary
dendrite arm spacings.' Two significant features emerge from this figure:
'Hereafter, the term "dendrite arm spacing" will be used to denote "secondary dendrite
arm spacing."
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
BECK AND SANTHANAM ON A CAST COBALT-BASE ALLOY 131
P^^'V/\':..
(o) cracks grow both interdendritically and across the dendrites and (b)
overall crack length decreases with increasing dendrite arm spacing.
In an attempt to monitor the structural changes that occur during thermal
cycling, the various carbide phases were carefully examined. It was ob-
served that the large script-like MC carbides were stable during the thermal
cycling test. However, fine M^jC. and MC carbides precipitated within the
dendrite areas (Fig. 10). To determine the total amount of carbide
precipitated during the test, the matrix was dissolved electrolytically [4],
and the insoluble residue was collected, weighed, and identified by X-ray
diffraction. Table 3 summarizes the variations in the type and amount of
carbides before and after the thermal fatigue test. The as-cast microstruc-
ture consists of two MC carbide phases which differ in their lattice pa-
rameter (MC #1 > MC #2) and one M7C, carbide phase. During thermal
cycling, the metastable M7C3 phase decomposes into MiaC* and MC car-
bides. Both of these carbides also precipitate from the matrix. As a result,
for a given dendrite arm spacing, the total amount of carbide increases
during thermal cycling. Table 3 shows that there is a greater increase in total
carbide in Specimen 6 (smallest dendrite arm spacing) than Specimen 2
(coarsest dendrite arm spacing). Also note that there is a greater increase in
the amount of MC #2 carbide in Specimen 6 than in Specimen 2. It was ob-
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
132 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
FIG. 7—Transgranutar nature of the crack growth in Mar-MS09. The longitudinal axis of
the specimen is vertical (x50).
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
BECK AND SANTHANAM ON A CAST COBALT-BASE ALLOY 133
I
I
fS
?1
y-s 00 a
11:
CO o ^
Oi
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
134 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
^
o
!2
I
d
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
BECK AND SANTHANAM ON A CAST COBALT-BASE ALLOY 135
v..
J* .« 4.1
•n
kf..
V
*^
t
!>: i
7
(a)
-**sJL
« * ;,
^ jii. ' * * ^"
:^r:
' ? « ? •
s - "•••
» .-. /..
^-
^ .y <i
'"'* -.
V^v •><•
f «
(b)
FIG. 10—£Xfec/ of thermal cycling on the microstructure of cobalt superatloy Mar-MS09.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
136 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
q *ri
en m
Ills
I c
s "!
I o
e^
<S
^ 00
t3 f*i •o
I vd >X
4> OU
>r) < N
o -S * o\ cK
I o
s
\o
d d
QO
H
<^
iL> bO
•S 'C
I 5^
T3 #
2
VO <N
d o
1 rt
J
^^
5t
u
r n r-;
•X •<t
:2
CO
«> iri r^
VO
•/^
(0
< M VO
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
BECK A N D S A N T H A N A M O N A CAST COBALT-BASE ALLOY 137
3000
i 2000
Present Study
1000
40 60 80 100
Dendrite Arm Spacing (|;m)
FIG. 11—Correlation between the number of cycles to initiate a crack and the dendrite arm
spacing for Mar-M509.
40 60
Dendrite Arm Spacing (pun)
FIG. 12—Correlation between the initial crack propagation rate and the dendrite arm
spacing for Mar-M509.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
138 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
is interesting to note that Howes' data point lies on the line extrapolated
from the present data.
The first stage crack propagation rate could also be correlated with the
amount of MC #2 carbide phase that precipitates during thermal cycling.
This correlation suggests that the effect of dendrite arm spacings on crack
growth rate could be an indirect one. No correlation was obtained between
the second stage crack propagation rate and the microstructure.
Discussion
The present study demonstrates that the response of a material to thermal
fatigue can be changed by changing its microstructure. In particular, it has
been shown that in cast cobalt-base alloy, Mar-M509, coarse dendrite arm
spacings improve the thermal fatigue resistance by delaying crack initiation
and also by reducing the crack propagation rate. It has also been shown that
thermal fatigue cracks initiate at the specimen periphery in the in-
terdendritic areas but propagate both interdendritically as well as across the
dendrites. Based on these observations, a model for thermal fatigue
cracking can be proposed as follows.
Let us consider Fig. 13. In this schematic drawing, two microstructural
conditions are depicted. Figure 13a shows a specimen with fine dendrite arm
spacing, whereas Fig. I3b refers to a specimen with coarse dendrite arm
spacing. In both cases, the cracks are shown to initiate at interdendritic
areas but to propagate both inter- as well as transdendritically.
The interdendritic regions contain a large number of script-like MC-type
Primary Dendrite
Surface
Cracks
Primary
Dendrite Arm
H' V Surface g
FIG. \3—Schematic of two microstructural conditions depicting a model for thermal fatigue
cracking in Mar-M509. Markers indicate secondary dendrite arm spacings.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
BECK AND SANTHANAM ON A CAST COBALT-BASE ALLOY 139
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
140 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Conclusions
1. The thermal fatigue resistance for a given heat of cast Mar-M509 can
be improved by employing solidification conditions that result in coarse
dendrite arm spacings.
2. This improvement occurs in both the number of cycles to initiate a
crack as well as the rate of crack propagation.
3. The carbides at the interdendritic areas act as preferred crack
nucleation sites and also provide easy paths for crack propagation.
4. The smaller number of interdendritic sites as well as a lower amount
of MC #2 carbide precipitated within the dendrite areas during thermal
cycling can explain the superior thermal fatigue resistance of the speci-
mens with coarse dendrite arm spacings.
Acknowledgment
The assistance of P. M. Yuzawich in preparing metallographic specimens
is gratefully acknowledged.
References
[1] Howes, M. A. in Fatigue at Elevated Temperature, ASTMSTP520, American Society for
Testing and Materials, 1973, pp. 242-254.
[2] Bizon, P. T. and Spera, D. A., "Comparative Thermeil Fatigue Resistances of Twenty-Six
Nickel- and Cobalt-Base Alloys," NASA TN D-8071, National Aeronautics and Space
Administration, 1975.
[3] Glenny, E. et al. Journal of the Institute of Metals, Vol. 84,1958-1959, p. 294.
14] Donachie, M. J., Jr., and Kriege, O. H., Journal of Materials, Vol. 7, No. 3, Sept. 1972,
pp. 269-278.
[5] Howes, M. A. H., NASA Report CR-121212, National Aeronautics and Space Ad-
ministration, 1973.
[6] Woulds, M. J. and Cass, T. R., Cobalt 42, March 1969, pp. 3-13.
[7] Sims, C. T., "The Superalloys—Chapter 5," C. T. Sims and W. C. Hagel, Eds., Wiley,
New York, 1972.
[8] Drapier, J. M., Leroy, V., Dupont, C , Coutsouradis, D., and Habraken, L., "Structural
Stability of Mar M-509—A Cobalt-Base Superalloy," International Symposium on Struc-
tural Stability in Superalloys, Seven Springs, Pa., 4-6 Sept. 1968, pp. 436-459.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
/. S. Laub'
REFERENCE: Laub, J. S., "Some Thermal Fatigue Characteristtcs of Mild Steel for
Heat Exchangers," Thermal Fatigue of Materials and Components, ASTM STP 612,
D. A. Spera and D. F. Mowbray, Eds., American Society for Testing and Materials,
1976, pp. 141-156.
ABSTRACT: Thermal fatigue of mild steel has received little attention until recently
even though this material is used extensively in heat exchangers for gas-fired appUances.
A survey of the thermal fatigue problem of thin shell mild steel structures subjected to
arbitrary temperature fields is given. Data from end-constrained thermally cycled test
specimens is also presented. These data have been validated through use in developing
production heat exchangers.
Nomenclature
A£, Total strain range, elastic and plastic
AT Temperature range
T„ Mean cyclic temp, T^^ + T^Jl
R Strain ratio £n,in/£max
a WeibuU location parameter
c Weibull shape parameter
M£ Microstrain, 10"* unit length/unit length
141
fired by a gas burner inserted into the lower opening to a thermal rating
of 33 000 Btu/h. The upper opening is attached to a flue collector box
which transfers the flue gas of this and an arbitrary number of other cells
to the stack. Residential and commercial gas furnaces have multiple num-
bers of these units determined by their capacity.
Heated internally by a gas flame, the cells are cooled by room air cir-
culated in most cases by a fan over their outside surfaces. Internal area
variations between adjacent cell surfaces or an outer cell surface and the
furnace casing produce variable and arbitrary air flow patterns which com-
bine with the flame thermal pattern to give a complex and time-varying tem-
perature pattern in the material. During typical furnace operation, any
given point in the cell will be subjected to a temperature cycle determined by
the operational functions plus the timing and duration of the flame and cir-
culating fan. The net result is that the thin shell of an appliance heat ex-
changer presents one of the most difficult and challenging thermal strain
and thermal fatigue problems to be found in engineering practice.
Economic considerations dictate that any new design must be well proven
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
LAUB ON MILD STEEL FOR HEAT EXCHANGERS 143
against thermal fatigue, but, until recently, the only accepted method for
doing this was the laborious, time consuming, imprecise, and inefficient life
test progrsim. Even life testing is subject to a wide divergence of opinion
since any given exchanger may be used in a multiplicity of appliance types
(upflow, downflow, crossflow, space heaters, and combination heating
and cooling units); and each installation has its own demand pecuHarities
which determine, to some extent, the cycle and temperatures encountered.
Further, by the time a life test program is complete, changing conditions
have often dictated new and previously unforeseen uses necessitating a
new, long-term test program. Although a component may cost only a few
dollars to manufacture, it will be made by the tens or hundreds of thou-
sands per year which necessitates a quicker and more efficient method of
proving structural integrity.
Experimental techniques for thermal strain analysis on thin shell struc-
tures, while difficult, have been developed in the author's laboratory to the
point of providing part of the thermal fatigue problem answer. Over a
period of several years, conventional experimental stress analysis methods
have been combined, refined, and adapted so that acceptably accurate
strain measurements can be made with readily available equipment.
Ceramic brittle coatings give initial external surface tensile strain patterns
above a threshold strain varying with the local temperature but ranging nor-
mally between 600 and 1200 fit. Medium temperature foil strain gages with
self-temperature compensation can be used in locations of the shell which
do not exceed 650 °F (343 °C). The manufacturer's apparent strain versus
temperature correction curve can be incorporated in the data reduction for
improved accuracy. For areas of the shell which exceed this temperature,
strain measurements require high temperature, wire strain gages applied
with flame-sprayed molten ceramic along with special temperature com-
pensation techniques and data reduction programs.
Acceptable strain measurement is one necessary part of the thermal
fatigue-reliability problem; the second is an accurate knowledge of the ther-
mal fatigue behavior of the exchanger material. Until recently, very little in-
formation was available for ordinary low-carbon or mild steel under con-
ditions of both varying strain £md temperature. Some work was published
on strain cycUng at constant elevated temperatures [1-3],^ for ex£unple, but
this did not really come to grips with heat exchanger operation where strain
changes in concert with temperature. As a consequence, thermal fatigue
tests were set up on specially constructed equipment following Coffin's
earUer work [4-7]. This paper presents the results of these tests. The ac-
curacy of the data has been estabUshed by successfully using it in designing
and qualifying rehable heat exchangers.
Recently, some new work has been pubUshed by Jaske [8] which confirms
^ The italic numbers in brackets refer to the list of references appended to this paper.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
144 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
the general accuracy of these data. His data were generated on equipment
allowing constant strain range and ratios and, as such, is more amenable to
conventional analysis. The data presented here were generated on much
simpler equipment and, consequently, required the proof of time and use.
Fortunately, the many uncontrollable and variable factors involved seem to
be similar for specimen testing, furnace life testing, and furnace field use.
Heat exchangers with reliability proven by service have been produced using
these data and the related experimental techniques, while other designs with
excessive strain levels were identified early in the development period and
corrected or discarded.
Procedure
Completely end-constrained axially loaded plain and notched thin-wall
tubes comprised the thermal fatigue specimens in this test. Figure 2 shows
the details of the test section. Large brass adapters were silver soldered into
CHEMISTRY
CARBON 0.05-0.15
MANGANESE 0.30-0.60
PHOSPHORUS 0.040 MAX.
SULPHUR 0.050 MAX.
25.4 DIA.-»
1.2
STANDARD MODIFIED
AREA 94.20inm^ HIGH STRAIN
LOW TEMR
AREA 64.52mm^
each end, providing alignment and gripping. These adapters were water
cooled to give a constant temperature heat sink at each end of the specimen.
They were also drilled axially to allow cooling air to be blown through the
tube interior during the cooling portion of the thermal cycle.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
LAUB ON MILD STEEL FOR HEAT EXCHANGERS 145
1000
1
Tm»600*F END OF HOT HOLt
aT=60< 3*F
800
• \ ^
600
<
400
END OF C}LD
< HOLD
200
^
BOTTOM 0 1 TOP
-X.in.—4 -I 0 I 4-•X.in.
SPECIMEN LENGTH
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reprod
146 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Cooling followed the hot hold time. In these data, the cold hold time
power was switched on at the same time as the cooling air. Room tem-
perature air was blown over the outside and through the inside of the
specimen until the low cycle temperature value was reached. At this point,
the air was switched off and only the cold hold time power applied to the
specimen. Redistribution of the low temperature followed for 5 s, allowing
the tensile thermal strain and stress to reach their maximum values. This
completed one thermal fatigue cycle.
Major restrictions with a test of this type are that compressive strain oc-
curs only by heating and tension by cooling; in order to vary the amount of
thermal strain available from a given temperature cycle, the specimen
geometry must be changed; and also, the strain ratio, R, is not controlled.
For these reasons, the data are presented on the basis of total cyclic strain
range, At,. R values generally fall between 0 and - 0.5.
On the other hand, the completely constrained test equipment is simple
and rugged. The concept was successful enough that other materials were
tested on a second similar machine. The total cost of the equipment at the
time it was built was approximately $5000.
Resistive heating was provided by a large transformer, the secondary coil
of which consisted of 1 Yi turns of heavy copper bar. Each end of the bar ex-
tended out to provide the connection to the ends of the specimen. A typical
electric power level during the heating portion of the cycle was 2500 A at
1.5 V. This power setting resulted in a heating rate at the center of the
plain tube specimen of 35 °F (20°C)/s. A wide range of control was pro-
vided for the transformer so that heating rates could be varied according
to the heating time required for a given test series and also to provide for
other specimen sections and sizes.
The end-constraint fixture consisted of an upper and lower platen which
gripped the specimen end adapters. Massive threaded tie bars and nuts with
faces ground parallel to the platen faces positioned the platens. Electrical
isolation between the holding fixture ends was accomplished by a set of
bakelite shoulder washers between the top platen positioning nuts and the
platen.
Specimens were loaded into the fixture platens but not immediately con-
strained. Several check and adjustment cycles could be run before actually
straining the specimen. These were used to obtain the total displacement of
the specimen and to make power adjustments when the thermal cycle was
changed. The movable platen was always constrained at the end of the hot
hold time so that the initial part of the strain cycle was tensile.
After constraint was imposed, no further attention was required until
machine shutdown indicated specimen failure or, on rare occasions, failure
of some component of the test equipment. Continuous temperature record-
ing at the center of the gage length provided a record for each specimen
which could be referred to at the end of the test. Any unusual temperature
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
LAUB ON MILD STEEL FOR HEAT EXCHANGERS 147
Results
Each data point on Fig. 4 consists of the average value of from four to six
specimens tested under identical mean temperature and temperature range
conditions. All of these data have 5-s hot and cold hold times, while the
heating and cooling times vary. On the average, cychc rate is about 45
cycles/h. At long lives, data are accumulated slowly. This accounts for the
scarcity of data beyond 10 000 cycles. Since a major effort was required to
obtain the data, a minimum statistical sample at each point was considered
necessary.
The data of Fig. 4 covers the area of greatest interest for furnace heat ex-
changers. While by no means complete, enough of the strain-life field is
covered to show general trends and conditions to be avoided if medium-
to long-term thermal fatigue reliability is to be achieved. Heat exchangers
are warranted for 10 to 15 years of life. From experience, 100 000 cycles
is considered an adequate design goal in most cases, so part of the de-
signer's task is to preclude low-cycle fatigue. For cooler areas of the ex-
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
148 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
u 0.51
^"-^
I 0.4, b.r^?^i=^^i=5[q
°= 0.3
O 0 4O0'F(2O5'C) "^^ B
K 0.2 "~-Tfc-
Q D 500rF(260'C)
• • 550'F(288'C)
^ — - A 600'F(3I6"0)
^ 0.1
1,000 apOO 5,000 10,000 20POO 50,000 100,000
N, .AVERAGE CYCLES TO FRACTURE
changer where the mean cycle temperature is 400 °F (260 °C) or below, an
upper limit of 0.2 percent strain is acceptable. In hotter areas, this must
be reduced to 0.1 percent or less.
As would be expected, thermal fatigue strength is greater at 400°F
(205 °C) than at any higher mean temperature. This corresponds to the
highest tensile strength for mild steels. It was necessary to use the second
(reduced cross section) type of specimen shown in Fig. 2 for these and the
450°F (232 °C) tests since, for plain tube specimens, fatigue cracks oc-
curred at the cool ends under these conditions.
All curves tend to converge in the high strain-short life area where the
strain range dominates the other peuameters. Here, the cyclic temperature
has less influence than at lower strains where temperature and time become
stronger factors. No fatigue limit indications are found in this data
although, under certain conditions, they may exist. All data shown here are
for material which goes well into the plastic range at the cycle extremes. The
change in slope between 500 °F (260 °C) and 550 "F (288 °C) may be due to
material characteristics, or it may be fortuitous. Further testing is needed
to clarify this situation.
Approximately 60 specimen tests make up the curves of Fig. 4. The
mean life value is presented without probability bounds since the sample
size at any given condition is too small to allow reasonably accurate ex-
treme values to be projected. In any case, there is so much uncertainty in
other parts of the fatigue analysis that more precision in these data
would be of doubtful value. As noted in the discussion section, use of
the data against full-scale furnace testing has shown the accuracy to be
completely adequate.
Figure 5 is an example of the way the data were treated. It was found in
some cases that the data fit the logarithmic normal cumulative distribution
function somewhat better than the WeibuU distribution. In other test
sequences, the two- or three-parameter (including the position parameter)
WeibuU distribution provided a better fit by the criterion of minimum
variance about the regression line. The ordered data for the 600 °F (316°C)
tests are shown by small dots. Log-normal population parameter estimator
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
LAUB ON MILD STEEL FOR HEAT EXCHANGERS 149
1 J. '
99.5
1 i^,\ ^ '^ 9 0 % CONFIDENCE
1 . <& ! BAND
99 f 1
1
1
/ / /
f
i 1 /(
/ / /» T
70 / / / / _4
/ ' :? V y ' ^
/I +'
U) 50 _^^ -« —4
i 40 ' P"
i 30 ,/
u- 20 ;
it 15.9
10
/ 1
n
)
5 / «
2
/ /1
T„=600 r
/ /
/ / '7
0.2 / / 7j
/ 3 4 5 6 7 81000 2 3 4 5 6 7SI0p00 3 4 5 6 7 OOPOO
CYCLES TO FAILURE
FIG. 5—Log-normal probability offailure with various temperature ranges at 600°F mean
temperature. Nonparametric lower 90 percent confidence band shown.
lines found by simple linear regression (least squares) are drawn through
these points. The median and one standard deviation below the median
value are noted by larger circled points. Population mean estimate points
are shown by the crosses denoted X. Because of the small sample sizes, it
was not considered appropriate to estimate the population extreme value
statistics by using an assumed distribution. Lower 90 percent confidence
bands (5 percent) obtained by the Kolmogorov-Smirnov nonparametric
procedure are shown. Statistical procedures just noted are all standard prac-
tice. The details of their use may be found in many texts; Refs 9 and 10
are examples.
For the mean life value, suitable confidence bands are available directly
from the log-normal Unes. Figure 6 is typical and shows the data of Fig. 5 as
they appear in the composite data, Fig. 4, but with the addition of a lower
90 percent confidence band on the mean. A good idea of the mean fatigue
life variability of the material can be obtained from this plot. Data from
other mean temperatures is similar, showing generally greater variability as
life increases at lower strain levels. The host of statistical time-dependent
processes increases the data spread as strain severity decreases and testing
time increases.
Discussion
The data presented here do not claim to give a complete picture of the
behavior of mild steel under thermal fatigue conditions. Thermal fatigue is
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
150 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
tf IJ3
0.5 * ~ ^ _ ^~^
ST
0.4
~> /MEAN
0.3 r*--
0.2 ^ LOWER 90%
~ CONFIDENCE
BAND 600 ' F T „
2
o
0.1
500 1,000 2,000 5,000 10,000 20,000 50,000
Nf,AVERAGE CYCLES TO FRACTURE
FIG. 6—A verage life as a function of strain range at 600°F (315 °C) mean temperature.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
LAUB ON MILD STEEL FOR HEAT EXCHANGERS 151
mm
JffiW
(a) Typical failures showing buckle in high temperature area at crack tip.
{b) Planar area of flat thermal fatigue (x 6).
FIG. 1—Detail of thermalfatigue test failures.
Fig. Ic. The through crack is at the left, and a series of more or less uni-
formly spaced surface cracks or checks occur along the top edge, the
outer tube surface in this case. These cracks have formed during the cool-
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
152 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
(c) Metallographic cross section adjacent to the through crack ('^ x 100).
(d) Detail of intermediate length crack of Fig. 7c showing oxidation (~ x 400).
FIG. 1—Continued.
ing (tension) part of the cycle. Bulging in the area of the cracks occurs
only after the crack has developed to the point of concentrating current
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authoriz
LAUB ON MILD STEEL FOR HEAT EXCHANGERS 153
flow and heat at its ends. Oxidation of the fracture surfaces is a uni-
versal feature, as is transgranular cracking. Experience indicates that, if
intergranular cracking is found in mild steel exchangers, some other fail-
ure mechanism such as corrosion or creep rupture is at work.
Figure Id shows the "oxide wedge" of the intermediate length crack of
Fig. 7c. Because the oxidation product has a greater volume than the
original material, there is a wedging action on the crack tip. In these tests,
there is a strong compressive load during heating which attempts to close
the cracks. After the oxide wedges have formed, the cracks cannot close,
and the effect is to shift R toward pulsating tension at the crack tip. It
follows that crack growth is not a simple function of the imposed strain
cycle. Further, the numerous cracks formed in the relatively uniformly
strained gage area interact until the time one crack is clearly dominant. At
this point in the tests, failure is only a few cycles away, so there is a gage
length size effect determined by the temperature profile which defines the
critically strained area. All of these factors influence the fatigue life of the
specimen, but they are also involved in actual exchangers, so there are com-
mon qualitative effects which are more or less equal in both situations.
Estimation of the fatigue performance of numerous heat exchangers over
the past five years has shown that the fatigue data hold up very well. Con-
sidering that the accuracy band of experimental strain measurements on
operating heat exchangers is fairly broad, great refinement of the thermal
fatigue data is not necessary. For strain measurements to about 650 °F
(343 °C), medium temperature foil strain gages give results with an error
band about ± 100 ME wide. High temperature wire gages with special tem-
perature compensation techniques are used from 650 to 1000 °F (343 to
530 °C) with a considerably wider error band estimated to be about ±300
[ii. on the average. Combination of experimental measurements with the
fatigue data to predict the response of several new heat exchangers, as well
as existing exchangers used in new applications, has estabUshed confidence
in both technical areas. Where strain and temperature measurements fall
within the limited hfe region of the data, failures Eire encountered, generally
at lives well within the statistical scatter expected. Usually in furnace Ufe
testing, failures £ire scattered over a much broader life range than for most
machine components because of the unavoidable vagaries of burner flame
and cooUng air flow patterns. Even for very severe strain conditions, the
failure rate in life test exchangers rarely exceeds 50 percent within
reasonable test times. These statistics are described by positively skew
Weibull probability of failure density functions (shape parameter 1.0 < c
< 3.5).
Several very reliable heat exchanger designs have been developed with the
help of the data. Furnaces have been manufactured for more than five years
which were given a clean bill of health by linear extrapolation as noted in
the Results section. Lack of field failures in combination with complete test
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
154 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
series survivals to 100 000 cycles has built confidence and established a
sound basis for using these data. Note that field results are a necessary part
of the qualification of the thermal fatigue analysis because direct
correlation of life testing to operational use is not available.
The most limiting features of completely constrained testing are the lack
of a constant compression-tension strain ratio and the difficulty of
changing the total strain range without changing the temperature range.
New, more sophisticated equipment has overcome these problems to a great
extent by providing strain control in a closed-loop feedback circuit. Tests
now in progress use a combination of strain and temperature control
providing more tractable, conventional low-cycle data on other materials of
interest. Because of previously noted shortcomings, no attempt has been
made to generalize the data or extract universal trends. There is a danger in
this which applies to even the most closely controlled laboratory data
because mild steel behaves in a nonhomogeneous, nonlinear manner under
high strain and temperature cycling. Even when a short gage length ex-
tensometer is used, the average strain measured is only an indication of the
section gross behavior once the plastic strain range is entered.
An example of nonhomogeneous behavior is shown in Fig. 8, a specimen
'1 * ( - -•ai^^'isi^jj*
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions auth
LAUB ON MILD STEEL FOR HEAT EXCHANGERS 155
Conclusions
The data presented in this paper provide one of the engineering elements
needed to eliminate thermal fatigue in gas appliance heat exchangers.
Another necessary ingredient is a reasonably accurate experimental strain
measurement technique which may be applied to operating exchanger
shells. Development of a strain measurement procedure has necessarily been
carried on concurrently with generation of fatigue data. Although mea-
suring quasistatic strains on thin shell structures at up to 1000°F (538 °C)
presents great difficulties, techniques have been evolved which give ex-
cellent predictive results in combination with the data. Prediction of failure
or survival under any normal operating conditions has been assured with
confidence through actual operating experience.
Application of these fatigue design techniques still requires a very high
level of technical and manual skill. High temperature measurements are
time consuming and relatively expensive when compared to the cost of the
heat exchanger element and normal experimental strain analysis. Further,
actual exchangers, preferably from production tooling, are necessary to do
the fatigue analysis. This means that the design is well along, and options
for change are limited and expensive by the time fatigue life prediction can
be made.
Finite element methods (FEM) have given hope of accelerating the design
of reliable heat exchangers by providing an analytical strain field solution
before tooling has been set up. This would be ideal because contours could
be changed in the computer model with minimal time and expense. To date,
however, FEM results obtained by the author's group during heat ex-
changer development have been disappointing, if not downright misleading.
The cause of this is that, to date, only a linear-elastic FEM code has been
used. Except for very preliminary indications of problems (where the stress
values exceed the local yield strength) this approach is not valid since critical
areas of the shells deform plastically and, in most cases, are entering
displacement controlled buckling at about the same time. Nonlineeu
material and geometry properties along with large deflection capabilities are
required in the finite element analysis. It remains to be seen if the nonlinear.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
156 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
References
[/] Mackenzie, C. T. and Benham, P. P., "Push-Pull Low Endurance Fatigue of En 25
and En 32B Steels at 20 °C and 450 °C," Proceedings }965-J966, The Institution of
Mechanical Engineers, Vol. 180, Part I, No. 30, London.
[2] Mackenzie, C. T., Burns, and Benham, P. P., "A Comparison of Uniaxial and Biaxial
Low-Endurance Fatigue Behavior of Two Steels," Proceedings 1965-1966, The In-
stitution of Mechanical Engineers, Vol. 180, Part 31, London.
[i] Benham, P. P., "High-Temperature Low-Cycle Fatigue: Survey of British Work," Paper
presented at 1967 Society for Experimental Stress Analysis, Spring Meeting, Ottawa,
Ont., Canada.
[4\ Coffin, L. F., Jr., and Wesley, R. P., "Apparatus for Study of Effects of CycUc Ther-
mal Stresses on Ductile Metals," Transactions, American Society of Mechanical En-
gineers, Aug. 1954.
[5] Coffin, L. F., Jr., "A Study of the Effects of Cyclic Thermal Stresses on a Ductile
Metal," Transactions, American Society of Mechanical Engineers, Aug. 1954.
[6\ Coffin, L. F., Jr., "An Investigation of Thermal-Stress Fatigue as Related to High-
Temperature Piping Flexibility," Transactions, American Society of Mechanical
Engineers, Oct. 1957.
[7] Coffin, L. F., Jr., "Design Aspects of High-Temperature Fatigue With Particular
Reference to Thermal Stresses," Transactions, American Society of Mechanical
Engineers, April 1956.
[<S] Jaske, C. E. and Porfilio, T. L., "Thermal-Mechanical Low Cycle Fatigue of AISI 1010
Steel, "Catalog Number H54375, American Gas Association, Cleveland, Ohio, 1975.
[9] Hald, A., Statistical Theory with Engineering Applications, Wiley, New York, 1952.
[10] Johnson, N. L. and Leone, F. C , Statistics and Experimental Design in Engineering
and the Physical Sciences, Vol. I, Wiley, New York, 1964.
[77] USA Standard for Gas Fired Gravity and Forced Air Central Furnaces, ANSI Z21.47-
1968, American Gas Association, Cleveland, Ohio, 1968.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
S. W. Hopkins'
KEY WORDS: thermal fatigue, mechanical tests, nickel alloys, crack initiation, crack
propagation
Thermal fatigue has been defined by Garden [ly as that process by which
cracks originate and are propagated in materials that have both cyclic tem-
perature and cyclic mechanicsd strain at a critical location. There are essen-
tially two experimental techniques used to obtain such material data. The
conventional thermal fatigue (TF) technique entails cycling the test
specimen between a hot source and a cold source. With this method, it is the
specimen geometry, the heating rate, and the temperature difference which
produce the cyclic mechanical strains. The resultant mechanic2d strains and
stresses have to be calculated analytically. For anisotropic materials, such as
'Presently, mechanical engineer. Failure Analysis Associates, Palo Alto, Calif. 94304;
formerly, senior materials engineer. Materials Engineering and Research Laboratory, Pratt
& Whitney Aircraft, Middletown, Conn. 06457.
' The italic numbers in brackets refer to the list of references appended to this paper.
157
gas turbine blade materials, this calculation is extremely complex. The other
experimental technique, known as thermal mechanical fatigue (TMF),
cycles both the specimen's mechanical strain and its temperature in-
dependently. It is this second technique and equipment that will be presen-
ted in this paper, with emphasis on those items or procedures which differ
from isothermal low-cycle fatigue (LCF) testing.
Before proceeding, definitions of the different types of strains are in or-
der. Thermal strain is defined as a change of specimen dimension caused by
thermal expansion at zero load (free thermal expansion) divided by its
original dimension. Mechanical strain is defined as that strain which is
caused by or which generates stress. Total strain, which is the displacement
measured from the specimen divided by the original gage length, is defined
as the algebraic sum of both the thermal strain and the mechanical strain.
Testing Machine
Presently, there are four TMF machines operating around the clock in the
Mechanical Behavior Laboratory of Pratt & Whitney Aircraft. All of these
machines are standard closed-loop, servohydraulic systems and are equip-
ped with load, strain, and temperature controllers, along with strain and
temperature programmers. One of these systems is shown in Fig. 1. For the
most part, this equipment can be purchased from any of the closed-loop
fatigue machine manufacturers. These machines were chosen over the com-
mercial screw driven machines because of their ability to control total strain
rate and because of their response time. For fast temperature response and
good control, a completely electronic closed-loop temperature controller is
used which has much better dynamic control than the old-fashioned elec-
trical-mechanical temperature controller that nulls by a slide wire system.
The details of how the strains and temperatures are controlled, measured,
and programmed will be discussed later.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions a
HOPKINS ON LOW-CYCLE TESTING 159
reasons. The thin wall allows for fast heating and cooUng cycles with a
minimum temperature gradient across the wall thickness. The wall thick-
nesses can be varied to model a component wall thickness. The strains are
measured on the centerline of the specimen which mechanically averages the
strains as Tishler and Wells [3] have shown, and the specimen is suitable for
both crack initiation and crack propagation testing. The alignment of the
loading system is checked and maintained to keep the bending strains within
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
160 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
-PULL ROD
-I!<-I8 IIEF-21IID.
iSc^:
-COLLAR
~
at -SPLIT RIHOS
I^ r .027
-SPECIMEN
2.4S« 3000ia.
1.000 y -050 R
I. ] N 0<0 R
>y i \,,
(-.901 'Oio.-j
FIG. 2—Solid isothermal LCF specimen with one end assembled in the standard tension-
compression gripping system.
FIG. 3—Strain control tubular fatigue specimen showing the internal ridges and the quartz
rod and tube extensometry system.
the specimen to less than 3 percent of the appHed axial strain. At small
strains where percent bending is meaningless for a stiff gripping system,
the bending strains on the specimen's surface are maintained below 10 '*.
Extensometry
To minimize the imposed strain errors and measure only displacements in
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
HOPKINS ON LOW-CYCLE TESTING 161
the gage section of the specimen, fused quartz was chosen for the ex-
tensometer material. Fused quartz has a low coefficient of thermal ex-
pansion and a high melting temperature, making it ideal for this applica-
tion. Fig. 4 is a sectional view of the specimen and extensometry. The
displacements are measured between the two internal ridges by the quartz
rod and tube which are extensions to the Unear variable differential trans-
former (LVDT). There are two high temperature springs: one on the bot-
tom to hold the quartz rod firmly against the lower ridge and one on the top
to hold the quartz tube against the upper ridge. Each spring applies less than
5 N force to the ridges. With this technique, the displacements measured
reflect only the relative movements of the two ridges.
MYBirAULIC 1*11
FIG. 4—Sectional view of the loading train with the quartz extensometry installed and the
collars and split rings removed.
Load Cell
The loads are measured from a standard load cell in series with the
specimen. The cell can be mounted in one of two locations: either above the
specimen, attached to the frame, or below the specimen, attached to the
hydraulic ram. There are two advantages to mounting the load cell to the
hydrauUc ram.
1. The temperature of the load cell is easier to maintain. The heat due to
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
162 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Programmers
The temperature and strain programmers can be a number of different
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions autho
HOPKINS ON LOW-CYCLE TESTING 163
FIG. 5—Typical TMF setup with the cooling air manifold and heating coil surrounding
the specimen.
devices depending upon the test frequency and the complexity of the
mechanical strain versus temperature curve that is to be programmed. Fig. 6
shows three standard cycles used to obtain material and design data. Cycle I
is a linear relationship between strain and temperature with the tensile strain
occurring at the lowest temperature and the compressive strain occurring at
the highest temperature. Cycle II is also a linear relationship between strain
and temperature and is the exact opposite of Cycle I, the tensile strain oc-
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
164 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
^TE«P»nU*E
•«SEI«LL CYCLE
TEMrEMTIWE
FIG. 6—The three standard TMF cycles conditions used to obtain material and design
data. Cycle I, Cycle II, and Baseball cycle.
curring at the maximum temperature. The third standard cycle, the Baseball
cycle, is in between a Cycle I and a Cycle II with both the tensile and com-
pressive strains occurring at the same middle temperature.
The original technique uses a dual channel Data Trak' follower for
programming total strain and temperature and is as follows: The tem-
perature versus mechanic^ strain and temperature versus time relationships
are chosen. Following this, the temperature command input curve is hand-
drawn on the Data Trak chart and then, the temperature is cycled main-
taining zero load on the specimen. During this cycling, the thermal strains
are recorded. Having acquired this thermal streiin data, the mechanical
strain and thermal strain are added algebraically, and the resulting total
strain would then be plotted onto the other channel of the Data Trak syn-
chronous with the appropriate temperature. With this procedure, there is no
way to ensure that the total strain curve is entirely plotted correctly until the
TMF test is started and the loads observed. This entire procedure has to be
repeated for each TMF test.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
HOPKINS ON LOW-CYCLE TESTING 165
AMPLITUDE ADJ.
RECORDER
Ly.D.T.
'*'{ MECHANICAL STRAIN)
FEEDBACK
divided into two parts. The first part uses the temperature feedback signal
from the pyrometer as the input and outputs a voltage that is adjusted to
match the thermal strain feedback voltage. This adjustable voltage is fed in-
to the closed loop electronics for control through one of the two strain span
pots, usually Span Pot B. The mechanical strain signal from the Data Trak
is fed into the other span pot, usually Span Pot A. The second part of the
compensator has as inputs the adjustable voltage signal and the strain feed-
back signal from the LVDT. It is wired to subtract the adjustable voltage
from the strain feedback signal and produces the resultant mechanical
strain signal. Once the adjustable voltage is correctly adjusted, this me-
chanical strain signal will be zero during thermal cycling at zero load and
will be the applied mechanical strain during a TMF test.
In setting up a test utilizing the thermal expansion compensator, the
specimen is cycled through the temperature range at the desired test rate.
During this time, the loads are maintained at zero, and the thermal strains
are measured. These strains are also plotted on an x-y recorder against tem-
perature feedback. This is done to check the strain extensometry system and
the temperature gradient. Any hysteresis loops on the x-^ recorder indicates
one of these items has to be corrected. The extensometry system could either
rub against itself or not seat properly against the internal ridges of the
specimen, producing a hysteresis loop. Also, a temperature gradient along
the gage length of the specimen which is different for the heating and
cooling portion of the cycle will produce a hysteresis loop. This is produced
because the pyrometer only measures temperature at a point, and the ex-
tensometer measures the strain over the gage section. Adjustment of the in-
duction coil corrects this aspect.
Having made the corrections to eliminate the hysteresis loop, the ad-
justable voltage from the nonlinear amplifier is adjusted to match the ther-
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
166 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
mal expansion feedback signal. This has been achieved when it is observed
that the mechanical strain signal will not vary during thermal cycling at zero
load. The curve drawn on the strain channel of the Data Trak must now
represent the mechanical strain rather than the total strain to be controlled.
This minimizes the precision to which the curve needs to be drawn and in-
creases the overall test accuracy.
Digital Computer
It is difficult to reduce the hysteresis loop of thermal strain versus tem-
perature to exactly zero. With a digital computer as the programmer, this is
not necessary. The computer can acquire and store the thermal strains as a
function of both temperature and direction, whereas the thermal expansion
compensator's output is only a function of temperature and not direction.
The computer also allows more complex TMF testing to be conducted with
no additional effort from the operator. Such tests may include mode
transfers, strain holds for a given time or to a given stress relaxation, and
stress holds for a given time or to a given creep strain.
Discussion
Three different techniques for programming the strain-temperature-time
relationship were mentioned in the previous section: Data Trak only. Data
Trak and thermal expansion compensator, and computer. The Data Trak
only technique requires a great deal of setup time for each test. The ac-
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
HOPKINS ON LOW-CYCLE TESTING 167
THKKNESS.t
-RADIUS, r
AK-AoViw G(a/rt)
FIG. 8—The Erdogan and Ratwani stress-intensity factor used in reducing crack growth
data from tubular specimens.
curacy to which these Data Trak curves has to be drawn is quite great,
especially for small mechanical strain ranges. The Data Trak and thermal
expansion compensator technique reduces the accuracy to which the Data
Trak curves must be drawn, reduces the setup time for each specimen, in-
creases the accuracy of the test, and aids in data reduction during the test.
However, the computer is still the optimum device to program this type of
test.
The accuracy to which the programmed strain curve has to be drawn to
has been mentioned a number of times. The following simple calculation
will prove this point and show why good temperature control is just as im-
portant as good strain control. A typical range in free thermal strain and
mechanical strain for a TMF test is 0.012 and 0.003, respectively. If the test
needs to be conducted within 10 percent accuracy on mechanical strain, this
means that the temperature control must be such that the free thermal strain
does not vary more than 0.0003 out of 0.012. This also requires that the
temperature and temperature gradient be controlled within 2.5 percent over
the entire temperature cycle. If total strain is being programmed from a
Data Trak, it must be drawn within 2.5 percent; however, if only
mechEuiical strain is being programmed from a Data Trak, it must be drawn
within 10 percent. Obviously running a TMF test to only 10 percent ac-
curacy is not acceptable; however, to run the test within 1 percent accuracy
can not be achieved with only Data Trak control. To hand draw the total
strain curve on the Data Trak or control the temperature dynamically to
within 0.25 percent is virtually impossible. The real accuracy lies between
1 and 10 percent with only Data Trak control depending upon all the
control parameters. With the thermal expansion compensator, accuracy is
only limited by how closely it is adjusted to zero mechanical strain during
the thermal cycling. Obviously, for a given thermal strain error, the test ac-
curacy increases as the mechanical strain range increases.
To conduct these types of tests with a computer is difficult and expensive.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
168 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
If the job is done correctly, the payoffs are worth the expense. The dynamic
accuracy to which load and total strain can be controlled to, with closed-
loop equipment, is better than 1 percent. The problem in TMF testing is
matching temperature control equipment to heaters for dynamic closed-
loop control. As already mentioned, small changes in temperature can trans-
late into large changes in mechanical strain if the total strain is left un-
changed. However, a small change in temperature during a cycle without
changing the mechanical strain usually does not change the fatigue response
of the material. Therefore, a computer should be progrjunmed to take this
into account. The computer should almost continuously read the in-
stantaneous temperature and adjust its total strain output to compensate
for any temperature error. The frequency at which the computer does this
should depend upon the maximum heating rate of the temperature equip-
ment and the accuracy desired from the test. For the materials and strain
ranges associated with gas turbine apphcation, the computer has to update
its total strain signal for every 2 °C of temperature error to maintain a 1 per-
cent accuracy on the mechanical strain. Presently, this type of programmer
is being installed to program this equipment.
These control problems never surface for isothermal LCF tests because
static temperature control is much better than dynamic temperature con-
trol, and any change in temperature that does occur only reflects in a
mean strain shift rather than a strain range shift. Fatigue life, measured
as either cycles to crack initiation or the crack propagation rate, is af-
fected more by a small change in strain range than it is by the same
change in mean strain.
Summary
A TMF test method has been described which has the capability of con-
trolling any mechanical strain-temperature-time cycle. A tubular specimen
design was presented that can be used for either crack initiation or crack
propagation testing. The axial total strains were measured with a unique ex-
tensometry system and separated into two components (thermal strains and
mechanical strains) through the use of a thermal expansion compensator.
The necessity for precise temperature control was demonstrated in terms of
overall testing accuracy, along with how digital computer control can im-
prove this type of testing.
Acknowledgments
The test method described in this paper was developed over a period of
years due to the dedicated efforts of numerous people who are or have been
associated with the Materials Engineering and Research Laboratory of
Pratt & Whitney Aircraft, a division of United Technologies.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
HOPKINS ON LOW-CYCLE TESTING 169
References
[}] Garden, A. E. in Manual on Low Cycle Fatigue Testing, ASTM STP 465, American
Society for Testing and Materials, 1970, pp. 163-188.
[2] Wells, C, H. in Manual on Low Cycle Fatigue Testing, ASTM STP 465, American
Society for Testing and Materials, 1970, pp. 87-99.
[3] Tishler, D. N. and Wells, C. H., Materials Research and Standards, American Society
for Testing and Materials, Vol. 6, Jan. 1966, p. 20.
[4\ Rau, C. A., Jr., Gemma, A. E., and Leverant, G. R. in Fatigue at Elevated Tempera-
tures, ASTM STP 520, American Society for Testing and Materials, 1973, pp. 166-178.
[5] Gemma, A, E., Langer, B. S., Leverant, G. R., this publication, pp. 199-213.
[6] Erdogan, F. and Ratwani, M., International Journal of Fracture Mechanics, Vol. 6,
No. 4, Dec. 1970, pp. 379-392.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
C. E. Jaske^
Thermal-Mechanical, Low-Cycle
Fatigue of AISI 1010 Steel
KEY WORDS: thermal fatigue, thermal cycling fatigue, alloy steels, tests, precipita-
tion hardening
Nomenclature
AT Cross-sectional area of specimen at temperature
E Elastic modulus
' Senior researcher, Structural Materials and Tribology Section, Battelle's Columbus
Laboratories, Columbus, Ohio 43201.
170
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
172 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
mechanical strain because the total strain is zero. The maximum me-
chanical strain (t = 0) occurs when the temperature returns to the original
value, and the corresponding stress is tensile (positive) if yielding occurs.
This situation is usually referred to as out-of-phase, thermal-mechanical
strain cycling. At the other extreme is in-phase cycling where maximum
(positive) strain is at the maximum temperature and minimum (maximum
negative) strain is at the minimum temperature. In a complex structure,
the strain-temperature relationship may be somewhere between these two
extremes.
This paper describes the results of a study where the low-cycle fatigue
behavior of AISI 1010 steel was investigated for conditions of combined
thermal and mechanical strain cycling. Experiments were limited to the
two extreme conditions of in-phase and out-of-phase cycling. Detailed
tabulations of data developed during this program and results of a com-
panion study on the isothermal, low-cycle fatigue resistance of this same
steel are reported elsewhere [1,2]}
The objective of this study was to obtain information on thermal-
mechanical fatigue behavior of a commonly used low-carbon steel. Ex-
periments combining both thermal- and mechanical-strain cycling are
more representative of many service conditions than strain cycling at con-
stant temperature. However, since thermal-mechanical fatigue testing is
more time consuming, costly, and complicated than isothermal fatigue
testing, the approach taken in this program was to conduct a limited
study of thermal-mechanical fatigue behavior over the temperature range
of 93 to 538 °C (200 to 1000 °F) for correlation with previous isothermal
work [2] on the same material. An upper temperature limit of 538 °C
(1000°F) was considered to be reasonable for most practical uses of AISI
1010 steel. The lower temperature limit of 93 °C (200°F) was chosen as a
reasonable approximation of cooling to room temperature. This lower
temperature level was well below the blue-brittle temperature range of 149
to 316°C (300 to 600 °F) previously observed for this material [2], yet it
was high enough to permit thermal cycling at a reasonably fast rate.
Most laboratory studies of thermal-fatigue behavior, as described by
King and Smith [3], have been conducted either by thermal cycling of
end-restrained specimens or by strain cycling of axially loaded specimens
at constant temperature. It was also noted that two other techniques of
evaluating thermal-fatigue resistance have been used: (a) testing of com-
ponents under actual service conditions and (b) thermal cycling of wedges,
disks, thick-walled cylinders, and locally heated sheet specimens [3].
Although the testing of components provides very useful and practical
data, it is usually not economical at early design stages and does not
provide a knowledge of thermal fatigue behavior under well-defined
^ The italic numbers in braclcets refer to the list of references appended to this paper.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized
JASKE ON AISI 1010 STEEL 173
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorize
174 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
- C c
W3
^U.
ctf
E
1)
ai
HSI
111!
I
00
^ CM
o e
<u ••3
o. n
>>
HKJ
o
H
Si
*-•
111
Vl
U
•S9 H
(4-
o
4»
Q,
H>.
II
;-?^'^ ?!
(A
V
00
c:
S ill
2 S2 2
^
0
F
CO
05
u
W
3
"^^ s g -ac
.0
CO CQ
2
(U
8g|8 i
0
c 0
u 00
u f*l
a. 222 2
E
O o
§l§ li B v-1
ffl
H
o
§§§§ §8 lU VO
4S u^
•3
'C
V
4-1
tS u V)
0
o
#
s 00
m i
0
d -J
u
g 16
a
(U
St: Z
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
JASKE ON AISI 1010 STEEL 175
u .S
liU
li II
1^
111i! f ^^
tf
O
ki n
M
H
—• —c <N <N <N —. M
^ ^ m r^ OS ^ —^
22 2
<N rs *N r i
—« 0\ r^ O
<N f*^ »0 f*^
\ en ^ r^ S88S|8
g;
«
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
176 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Experimental Procedures
The material for this program was taken from the same heat of special
quality hot-rolled, 1.91-cm (3/a-in.)-diameter AISI 1010 steel bar used
in the previous study of isothermal low-cycle fatigue behavior [2]. The
chemical analysis and tensile properties of this steel were typical of those
expected for this alloy and are reported in that same study.
The thermal-mechanical fatigue specimens were machined to the con-
figuration shown in Fig. 1. This is the same external configuration used
for the isothermal-fatigue specimens. Use of this specimen configuration
allowed the temperature to be controlled at the minimum diameter where
strain was measured and reduced the influence of temperature gradients
on fatigue life. The present specimens were hollow, rather than solid as
for the isothermal studies, to facilitate cooling.
The test section of each specimen was polished with successively finer
grades of silicon-carbide paper to produce a surface finish of 0.4 jim
(16 /iin.) rms or better, with finishing marks parallel to the longitudinal
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized
JASKE ON AISI 1010 STEEL 177
-9.85(3.88)-
-4.67(1.94)-
_ 3.17 _
(1.25) •0.635 (0.250)
2.23 -0.33 (0.i30)
(0.88) -0.635 (0.250)
Iiniiiuiu/;.
r
in)in'///, •///rniuni
'/yjlllllin
\ '\ in. (i,59 cm)-i8 UNF tiid.
0.4 ^m (ie^in.) rms finish or
Dimensions: cm (in.) better In test section
Apparatus
The thermal-mechanical fatigue experiments were conducted using the
S2ime basic servocontrolled electrohydraulic test system that was used for
isothermal-fatigue experiments [2]. This type of system has been used in
numerous low-cycle fatigue studies (for example, see Ref 20). Details
of the equipment, method of heating, gripping and alignment procedure,
technique of strain measurement, and general experimental procedures
are given in any one of several reports [1,2,21]. In the present work, the
closed-loop axial strain control system was integrated with a closed-loop
temperature control apparatus, as illustrated schematically in Fig. 2. Thus,
both the axial mechanical strain and the temperature were controlled
throughout each test.
Load was applied to the specimen axially; diametral changes in the
specimen were measured using a special diametral extensometer; and load
was measured by a strain-gaged load cell in series with the specimen.
Diametral mechanical strain was obtained by subtracting the diametral
thermal strain from the total diametral strain using a simple analog cir-
cuit. The thermal strain was computed from the linear coefficient of
thermal expansion and the change in temperature. Diametral mechanical
strain and load signals were combined and converted to an axial mechani-
cal strain using an analog computer [22].
The axial mechanical strain and temperature were controlled to follow
one of the two basic constant-amplitude waveforms shown in Fig. 3. In
Fig. 3a, £max is in phase with Tmax; whereas, in Fig. 3b, Emm is out of phase
with TmsLx. (or £min Is in phase with Tmax). The control program waveform
was provided by an arbitrary function generator. Heating rate was not
constant, but it changed throughout the cycle as shown in Fig. 3. The
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized
178 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Signal
Conditioner
? Upper Crosshead S
Thermocouples
Proportional
Temperature
Controller
4 5 0 kHz
2 5 k W Induction
Heoting Unit
Hydraulic actuator
heating rate was determined by the thermal response of the specimen and
the heating equipment, with the length of the thermal cycle maintained at
about IVi min. Mechanical strain was programmed through appropriate
phasing and scaling of the same waveform used to control temperature.
In two of the tests, the waveform was modified by introducing 6-min.-
long temperature and strain hold periods at maximum temperature.
Before starting each test, controls of the thermal-strain computer were
set to give close to zero diametral mechanical strain when the tempera-
ture was cycled with the load at zero. At the start of each test, the
specimen was heated to the mean temperature to obtain values of com-
pliance and elastic Poisson's ratio needed in the analog computer. These
values at mean temperature were then used throughout the test. Since
these values were actually temperature dependent, assuming them to be
constant introduced errors, as discussed in the next section of this paper.
During each fatigue test, load was continuously recorded using a time-
base strip-chart recorder. Load versus axial extension hysteresis loops
were periodically recorded during each test using an x-y recorder.
Data Acquisition
Use of the hourglass-type specimen (Fig. 1) required that diametral.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions autho
JASKE ON AISI 1010 STEEL 179
l-).,T
rather than axial, strain be measured. Because axial mechanical strain was
of prime interest, diametral strain had to be converted before the signal
from the diametral extensometer could be effectively used in the control
loop. To accomplish this conversion, it was first necessary to compute
diametral mechanical strain.
Since diametral thermal strain was simply a AT, mechanical diametral
strain was given by the following relation
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authoriz
180 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
and Jraax. The size of this error, in terms of strain, is discussed after the
introduction of Eq 2.
To compute axial mechanical strain from diametral mechanical strain,
it was then necessary to determine the elastic and plastic components of
diametral strain. Once these were obtained, the elastic and plastic com-
ponents of axial mechanical strain were computed and the results summed
to generate the total axial mechanical strain. The necessary equations for
the computation of axial strain and the analog circuits of the computer
were described in detail by Slot et al [22].
In this study, the value of Poisson's ratio for inelastic deformation was
assumed to be 0.5, and the value of Poisson's ratio for elastic deformation
was determined for each specimen. The specimen was installed in the test
fixture and heated to the mean temperature of the thermal cycle. The
value of specimen compliance, K, was then set into the computer by load-
cycling the specimen for several cycles at low loads in the elastic region
and adjusting the K potentiometer until the value of Af^p was equal to
zero. The value of elastic Poisson's ratio was then determined from the
equation
where the values of Aeae and AP were determined from the slope of the
load versus diametral extension x-y plot recorded during a few load cycles
after setting the value of K. Values of E used in Eq 2 were the average
results of the tension tests reported previously [2]. Since values of AT and
E at the mean temperature of the thermal cycle were used to compute
V, at the mean temperature, an additional error was introduced in the
computation of the total mechanical strain at other temperatures.
Considering both the error introduced in Eq 1 by assuming constant
value of a and the error introduced in Eq 2 by assuming constant v^,
the magnitude of total possible errors were computed for each tempera-
ture range. For 93 to 316°C (200 to 600°F), it was about 0.03 percent
strain; for 93 to 427 °C (200 to 800°F), it was about 0.04 percent strain;
and for 93 to 538 °C (200 to 1000°F), it was about 0.05 percent strain.
Thus, for example, at 0.5 percent strain amplitude, these would be rela-
tive errors of 6, 8, and 10 percent, respectively. Although these are
fairly large errors, they are in the opposite direction for each side of the
mean temperature and cause little error in the total strain range. Also,
they are likely to have only a small influence on fatigue life for strain
ranges with relatively large inelastic strain components. However, if this
test procedure is to be used for low-strain range, long-life tests in the
future, it must be modified to account for variations in a and v^.
Examination of the load-time records revealed changes in the cyclic
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions aut
JASKE ON AISI 1010 STEEL 181
. . O a t £max O a t £min
A£, = A£. - — ^ — (3)
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized
182 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
fnorovorN»nOO*ol>«/-ii^OONi
r^oO'-^>n>nr^fI?w^Qoooooro(Si
I ' v^ w^ m in o (
I fN ^o OS r^ O <
I «ri <^ r- f<i rsi (
?
« 0 — ' O O O O O — - ^ — l O O O O O O
iS -S C!
— CO
^
s
•Sf
i<N—lOOO—• — «—'OO—'O—CO
I Eo
88888888888
OOOOOOOOOOOOOOOOVC'O^
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
88888888888888888
.SO 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
c
«
In
0. 0
u !2—
J. : SB g E 6 S E E g E E E 6 E e E S t=
b. k. c
3 e0
a
S S .S .= .S .S S S .S .3 .= S .E .S S .g .= 0
t- E g g E E e S S e E E g a E E E S
e •03
t 3 ns
0 <u
C 0
0c
«e Z
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
JASKE ON AISI 1010 STEEL 183
< s
E
b
in b
Number of Cycles
500
400
1
E
< ._ f M
< ^
300
1i 3
i=
(n
200 i b
oo
e < 7
-200
en E
b
-300 e
E
-400
Number of Cycles
W £ max ^t / max-
( D ) £ min ^t ^max-
FIG. 4—Cyc/fc stress response of AISI 1010 steel under thermal-mechanical strain cycling
with a temperature range of 93 to 427''C (200 to SOO'F).
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
184 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
FIG. 5—Stress-strain hysteresis loop for 93 to 538°C (200 to 1000°F) cycle with £„„ at
Tm„ (specimen 56 at cycle 40).
when £max o r £min W3.S 3.t J max, there was generally a difference between
omax (or omin) and o at £max (or o at Ernin). As shown by the cyclic stress-
strain curves in Fig. 6, this difference was quite significant at higher
temperatures and larger strain ranges. As temperature became lower or
strain range became smaller, such differences became much less pronounced.
For out-of-phase cyclic stress-strain response, the difference between
Omax and a at Emax o r between Omin and o at cmin was only about 3 percent
of Omax or Omin for 93 to 538 °C (200 to 1000 °F) and negligible for the two
lower temperature ranges. However, as shown in Fig. 6, there was a dif-
ference between such data for in-phase cyclic stress-strain response. For
a constant strain amplitude, this difference became larger as the tempera-
ture range increased. It also became larger as the strain amplitude in-
creased for a constant temperature range. At 93 to 316°C (200 to 600°F)
and 93 to 427 "C (200 to 800 °F), values of a at £max or £mi„ were still
above the isothermal cyclic stress-strain curves for the maximum tempera-
ture of that cycle. However, for 93 to 538 °C (200 to 1000 °F), values of
o at £max o r £min were about the same as the isothermal cyclic stress-strain
curve at 538 °C (1000°F).
As shown in Fig. 5, an instabihty in the stress-strain response was en-
countered near 149°C (300 °F). This sort of instability was related to
dynamic strain-aging processes that resulted in discontinuous yielding
behavior. These instabilities tended to become less prominent with cyclic
hardening for all 93 to 316°C (200 to 600 °F) and 93 to 427 °C (200 to
800°F) cycles, and no significant instabilities were noted in the stabilized
hysteresis loops for these temperature ranges. However, such was not the
case for the 93 to 538 °C (200 to 1000°F) cycles, where the instabilities
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorize
JASKE ON AISI 1010 STEEL 185
200 S
FIG. 6—Cyclic stress-strain behavior of AISI 1010 steel under in-phase thermal-mechanical
cycling.
were still observed in the stable loops as shown in Fig. 5. Thus, it was
hypothesized that the time at temperatures near 538 °C (1000°F) was
sufficient to reverse the processes that caused dynamic strain aging at
lower temperatures. On the other hand, such was not the case for lower
peak temperature cycles.
Although Omax (ox Omin) was not always at tmax (or tmin), the values of
Omax and Omin provided a quantitative measure of the cyclic stress-strain
response. Therefore, they were used to characterize cyclic stress-strain be-
havior in the remaining discussion. For the 93 to 427°C (200 to 800°F)
(Fig. 4) and the 93 to 538°C (200 to 1000°F), the cychc stress in phase
with Tmax was relatively stable after the first few cycles and was within 5
percent of the value at Nf/1 for more than 90 percent of the cyclic life.
In contrast, the out-of-phase stress for these temperature ranges showed
continual cyclic hardening and the stress after 10 percent of cyclic life
was about 10 percent less than the value at Nf/2. The stress response
for both in-phase and out-of-phase cycling at 93 to 316°C (200 to 600°F)
showed the same trends as that for the out-of-phase cycling at the two
higher temperature ranges.
Morrow [25] has shown that it is often possible to express the cyclic
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorize
186 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
where K' = £ and « ' = 1 for 0 < A£,/2 < £o. The parameters E, K',
and n' and £o are functions of temperature as listed in Table 3; the
parameter Oo, also listed in Table 3, is directly related to to (oo = £"£0).
The thermal-mechanical, cyclic stress-strain curves defined by Eq 4 are
shown by the dashed Unes in Fig. 7. For comparison, isothermal cyclic
curves from Ref 2 are shown in Fig. 7 as solid Unes to indicate that more
cyclic hardening occurred in the thermal-mechanical than in the isothermal
experiments.
Fatigue Resistance
Results of the thermal-mechanical fatigue experiments are compared
with the previous isothermal fatigue data [2]
{a) In terms of As, in Fig. 8.
(6) In terms of Aa/2 in Fig. 9.
(c) In terms of Omax A£,/2 in Fig. 10.
Based on At, as shown in Fig. 8 (also true for Afp as shown in Ref 2),
the thermal-mechanical results fell below the isothermal results. This dif-
ference became more marked at higher peak temperatures and was greater
for out-of-phase cycling than for in-phase cycling. It was also noted that
curves for out-of-phase cycUng were generally steeper than those for in-
phase cycUng. Thus, the out-of-phase cychng was more damaging than
the in-phase cycling for these conditions.
There may be three main reasons why the thermal-mechanical data fell
below the isothermal data. First, thermal-mechanical cycles were applied
at a variable and lower strain rate than the isothermal cycles. In the
former case, the slowest strain rate was comparable to a frequency of
about 0.1 cpm; whereas, in the latter case, depending on the strain range,
since strain rate was maintained constant, the frequency was about 6 to 60
cpm. At high temperatures, usually greater than 427 °C (800°F), such a
reduction in frequency can cause a reduction in fatigue life. For AISI
1010 steel at 500°C (1112°F), Coffin [24] has shown that the effect of
frequency can be described by the following equation
For Ae, = 1.0 percent, this equation predicts (a) A^ = 860 cycles at v
= 0.1 cpm, (b) Nf = 2280 cycles at v = 5.0 cpm, and (c) Nf = 4060 cycles
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
JASKE ON AISI 1010 STEEL 187
1^ -^ >n
lO vo —
o o o
t - o oo
r^ oo ' *
I »rj <s \o
"O 0> OO
cs r ) M
I
1
.5
«N (S ^
in ^ ^
m M (S
888
I
z
o
^£
0) o
Eu 888
H . vo p- oo
^ fS m
r<v ^ <n
Jo
E
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
188 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
JASKE ON AISI 1010 STEEL 189
z nr t Z i Z u : iZ _
O Q o
o u. c
ill •fc
1 1 1 1 1
i
8 izc M s=.
" 888888 s %^ /fi 1
o o o o o o
CO 00 N t^ ff) 0 /I / 1ft
srM
ID
ro ro c^ OJ ^ ro
IT) If) f ^ ro 1
1 1 1 1 1
fO ro K !S K ro
a> 01CT)<J) (Tl
o o o o o
/ 1
c
E E E E E
J
•S
t- 1- 1- 1- K
"5 "5 o o o O
S s S .= S c / , ' /// ili /
' / g
1
• ^ E E E E E <m g
g w w w w to / / /// / s;
•% o • < ^ 0 »
i, V ^ / / / /
///
11
o B
1 >l
E
1 = o c 3 // / "S^
1
X
i
o
E
1/ TVrr u- *
•jt 1 i
1i
V V
/
/ / /
1
j iT* f
11// 1 :§|i:
5i
11 / ! 1
y v^ O o <->
00 2
:5
/ /
7
f
/ o
1
a> i
2
Ifi / /
1^-f
x: g ^ . J
3
o
1 t Ap
J i J
/
4^
/ 1
1 11 ,' •5
°
i
if r o
1
^
-1 L V i— L _
g
*»V 'a6uDy ui.DJiS |D!XV |D|OI
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
190 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
I
s
-TT G
.i
I
I
a
I
d
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized
JASKE ON AISI 1010 STEEL 191
••"
%
1 I/ /
17
'o
1
'o
1 I
V / I
/
/
/
tz /
/ /
tzo / i
i1
/ / /
-8 O i
J /
o— 'J" / '
I D /
u.
r-1 O fO. / /
8^1
. o r ' <o
^hj > ro
^
/ ^
/ /
f ^ /
/ // ^ J^
\,
• / As /
/
r
/
/
o (/>
^o
>»
(J
!"
^
cs
•*3
•a
8" lr/\ P / y '' / g g o O o o 3-
Jf"
Z
^^••«
88g§og B ^3
r / £ •Sf
^
u\
1 1 1 t 1 1 U '4— t ^
/ 888888 1 - Zi •a.
^ r £
/
/ / \// -
f /f
/
/
/ } g po oo po f3 (5
O O O < J ^ O
00 GO f^ ^- crt i n
f^* ^ Sii ai S =
£
3
i
«)
.i>
t
o
[1
1
1 a>
1 CT*
1 (T>
1 <T)
1 01
1 K
f / / / ¥ / , 1
0^ E
• «
X X K M M j< O c
/ VJ / Q
O O O O O O
> 10 / » -E KE 1 E
- 1 -E ( -£ K £ .££ ^»
q
tj
01
^
'^1
>
•
>
3
3
0 B ^ B B o
S £ S £ S E 1
1 E E E e ^E ^ -
S
o
•*:
C
« •^ ~y ' £
- o
u E .a o
Is s e
•i2
E <u S
is
1 3
2
C
I
aI
LU1 ,
1
°o o
O
!s>| '2/'»v'"'"'j> 'jajaujDJDd uiDJiS-ssaiiS
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
192 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
(250 to 350 °F) temperature range allowed dynamic strain aging to take
place during each cycle. It is hypothesized that this repeated dynamic
strain aging was responsible for the larger amount of cyclic hardening
which occurred during the thermal-mechanical tests in comparison with
that which occurred during the isothermal tests. If the cycUc hardening
caused the stress to be higher for the same strain range without increasing
the fatigue resistance of the material, then a shorter fatigue life would
be expected for the higher stress level.
To study this hypothesis, the thermal-mechanical data were compared
with the isothermal data on the basis of stress amplitude in Fig. 9. There
was no significant difference between the in-phase and out-of-phase
results when compared on this basis. Data from 93 to 538 °C (200 to
1000 °F) cycling fell slightly above the 427 °C (800 °F) isothermal curve;
data from the 93 to 427 °C (200 to 800 °F) cycUng fell above the 371 °C
(700 °F) curve; and data from the 93 to 316 °C (200 to 600 °F) cycling fell
above the 316°C (6(X)°F) curve. The isothermal curves were taken from
Ref 2 where heavy soUd Unes are shown in Fig. 9 and were extrapolated
as shown by the Ught lines. The curve at 371 °C (700 °F) was obtained
by logarithmic interpolation between the curves at 21 and 427 °C (70 and
800°F). The curves at 482 and 593 °C (900 and llOO'F) were obtained in a
similar manner using the curves at 427 and 538 °C (800 and 1000 °F) and
at 538 and 649°C (1000 and 1200°F), respectively. To get the 371 °C (700°F)
curve, the 21 ° (70 °F) curve was used rather than the 427 °C (800 °F) curve
because it is beUeved that the strength dropped rapidly between 316 and
371 °C (600 and 700 °F) and that such a choice was more realistic. In view
of these comparisons, it appears that the increased stress due to cyclic
hardening was the primary reason for the lower fatigue life observed
previously on the basis of At, or A£p.
Smith et al [27] have used the following empirical parameter to assess
the fatigue resistance of metals
Omax(A£,/2)
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorize
JASKE ON AISI 1010 STEEL 193
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
194 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
r , s ^"^ '^ -^""", for r„ax < 400 °C (752 °F) (6a)
and
and
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions auth
JASKE ON AISI 1010 STEEL 195
10' 10'
Actual Fatigue Life, cycies
FIG. 11—Comparison of actual thermal-mechanical fatigue life with cyclic life calculated
from isothermal fatigue curves.
for both the 93 to 316 °C (200 to 600 °F) and 93 to 427 °C (200 to 800 °F)
data. Thus, for temperatures of 427 °C (800 °F) or less, this parameter
gave a reasonably accurate assessment of fatigue life, amd it could be
used successfully in this lower temperature range. However, the calcula-
tions were significantly unconservative at 538°C (1000°F). Therefore, life
calculations for this steel and these temperature ranges were more con-
sistent when made using the approach based on Aa/2 and T^, which gave
about the same results for all three temperature ranges 2md was conserva-
tive by a factor of about three on average cyclic life.
At first, it may seem that a factor of three is a large difference, but
it is important to consider how such calculations are used in design applica-
tions. The designer usually follows one of two approaches: {a) he deter-
mines the cyclic stresses and strains in the critical area of the component
under consideration and uses them to assess whether or not adequate
fatigue Ufe is available or (Jb) knowing what fatigue life is desired of the
component in service, he designs it so that the cyclic stresses and strains
will be low enough to assure achievement of desired life. Whichever
approach is used, it is always necessary to apply a safety factor to either
the stress and strain or to the fatigue life. Safety factors of 1.5 to 2.0
are normally applied to stress or strain, whereas factors of 10 to 20 are
usually applied to cyclic Ufe. The reason for larger factors on life than
on stress or strain is that fatigue life is usually treated as a logarithmic
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorize
196 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
variable and small changes in stress or strain produce much greater changes
in fatigue life.
When actual cyclic lives from the thermal-mechanical experiments were
used to compute Ao/2 values from the isothermal fatigue curves at Te, the
actual stress amplitudes were only about 18 percent greater than the cal-
culated ones on the average. Thus, a factor of three under calculation
of the average number of cycles to failure is not excessive in relation to
the relatively small variation in stress amplitude that can cause a factor
of three variation in cychc life. The most important point is that this
approach gives a consistent method of computing a lower bound on
thermal-mechanical fatigue life of this steel from isothermal fatigue infor-
mation. Computation of such a lower bound is more valuable for design
purposes than is prediction of average fatigue life. Although the actual
thermal-mechanical fatigue data might be considered for use, the limited
amount of these data makes the statistical confidence in them lower than
in the more numerous isothermal data.
It is important to note that either of the two life prediction methods
just discussed required a knowledge of the half-life stress amplitude and
that strain alone did not provide a sufficient correlation between thermal-
mechanical and isothermal fatigue data. In general, a knowledge of both
stress and strain were required to make accurate fatigue life predictions.
Thus, in studies of thermal-mechanical fatigue behavior of materials, it
is necessary to develop information on cycUc deformation behayior as
well as on cyclic life. Also, the methods are based upon information for
uniaxial loading and are not generally valid for multiaxial loading. Develop-
ment of a method for complex multiaxial loading situations that are often
encountered in design appUcations should be pursued in future studies.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authoriz
JASKE ON AISI 1010 STEEL 197
Acknowledgments
The work described in this report was sponsored by the American Gas
Association, Inc. (Project No. HA-3-1), and was performed at Battelle-
Columbus, Columbus, Ohio. Thanks are extended to Dr. Ab Flowers of
the American Gas Association, Inc., for his guidance of the program and
to R. W. Newell, G. C. Groff, and O. C. Davis of the advisory com-
mittee for their technical assistance.
Appreciation is EIISO extended to the following Battelle-Columbus staff
members for their assistance in conducting the experimental work: N. D.
Frey, J. J. Parks, and H. J. Mahk. T. L. Porfilio, formerly of Battelle-
Columbus and presently with Dialight, Brooklyn, New York, assisted
in technical conduct and planning of this program. Dr. L. E. Hulbert
and H. Mindlin of Battelle-Columbus provided direct technical super-
vision of the studies.
References
[1] Jaske, C. E. and Porfilio, T. L., "Thermal-Mechanical, Low-Cycle Fatigue of AISI
1010 Steel," Report to American Gas Association, Inc., Catalog No. H54375, Bat-
telle's Columbus Laboratories, Columbus, Ohio, 14 Feb. 1975.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
198 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
[2] Jaske, C. E. and Porfilio, T. L ., "Isothermal, Low-Cycle Fatigue of AISI 1010 Steel
at Temperatures Up to 1200 F , " Report to American Gas Association, Inc., Catalog
No. H54975, Battelle's Columbus Laboratories, Columbus, Ohio, 14 Feb. 1975.
[5] King, R. M. and Smith, A. 1. in Thermal and High-Strain Fatigue, Monograph and
Report Series No. 32, The Metals and Metallurgy Trust, London, 1967, pp. 364-378.
[4] Carden, A. E. in Manual on Low-Cycle Fatigue Testing, ASTM STP 465, American
Society for Testing and Materials, 1970, pp. 163-188.
[5] Manson, S. S„ Thermal Stress and Low-Cycle Fatigue, McGraw-Hill, New York, 1966.
[6[ "Thermal Fatigue of Metals in Naval Power Machinery," Wash 1178, "Sudostroyeniye"
Press, Leningrad, 1967, given to the 1970 U.S. Nuclear Power Reactor Delegation.
[7\ "Thermal and High-Strain Fatigue," Monograph and Report Series No. 32, The
Metals and Metallurgy Trust, London, 1967.
[8] Fatigue at High Temperature, ASTM STP 459, American Society for Testing and
Materials, 1969.
[9\ International Conference on Thermal Stresses and Thermal Fatigue, Berkeley, England,
1%9.
[10] Fatigue at Elevated Temperatures, ASTM STP 520, American Society for Testing and
Materials, 1973.
[JJ] "International Conference on Creep and Fatigue in Elevated Temperature Applica-
tions," American Society for Testing and Materials, American Society of Mechanical
Engineers, and Institution of Mechanical Engineers, Philadelphia, 23-27 Sept. 1973.
[12] Carden, A. E., "Bibliography of the Literature on Thermal Fatigue," N68-23652
(NASA CR-94605), University of Alabama, Huntsville, 1967.
[13] Wundt, B. M., Low-Cycle Fatigue Bibliographic Survey, ASTM STP 449, American
Society for Testing and Materials, 1968.
[14] Kawamoto, M., Tanaka, T., and Nakajima, H., Journal of Materials, Vol. 1, No. 4,
1966, pp. 719-757.
[15] Udoguchi, T. and Wada, T., "Thermal Effect on Low-Cycle Fatigue Strength of
Steels," International Conference on Thermal Stresses and Thermal Fatigue, Paper
No. 18, Berkeley, England, 1969.
[16] Lyle, J. M., "Forced Air Furnace Exchangers; Some Design and Reliability Evaluation
Considerations," Gas Appliance Engineers Society Meeting, June 1971.
[17] Taira, S. and Inoue, T., "Thermal Fatigue Under Multiaxial-Thermal Stresses," Inter-
national Conference on Thermal Stress and Thermal Fatigue, Paper No. 3, Berkeley,
England, 1969.
[/*] Taira, S. in Fatigue at Elevated Temperatures, ASTM STP 520, American Society for
Testing and Materials, 1973, pp. 80-101.
[19] Taira, S., Fujino, M., and Takashi, M., Journal of Society of Materials Science, Japan,
Vol.22, 1973, pp. 235-241.
[20] Jaske, C. E., Perrin, J. S., and Mindlin, H., Reactor Technology, Vol. 15, No. 3, Fall
1972, pp. 185-207.
[21] Jaske, C. E., Mindlin, H., and Perrin, J. S., "Low-Cycle-Fatigue Evaluation of Reactor
Materials," in USAEC Report No. BMI-1928, Battelle's Columbus Laboratories,
Columbus, Ohio, July 1972.
[22] Slot, T., Stentz, R. H., and Berling, J. T. in Manual on Low-Cycle Fatigue Testing,
ASTM STP 465, American Society for Testing and Materials, 1969, pp. 100-128.
[23] Morrow, JoDean, in Internal Friction, Damping, and Cyclic Plasticity, ASTM STP
378, American Society for Testing and Materials, 1965, pp. 45-87.
[24] Coffin, L. F., Jr., "A Generalized Equation for Predicting High-Temperature Low-
Cycle Fatigue, Including Hold Times," Air Force Conference on Fatigue and Fracture
of Aircraft Structures and Materials, Miami Beach, Florida, 1969, (Also General
Electric Research and Development Center Report 69-C-401, 1969.)
[25\ Hirschberg, M. H. in Manual on Low-Cycle-Fatigue Testing, ASTM STP 465, American
Society for Testing and Materials, 1969, pp. 67-86.
[26] Manson, S. S., Halford, G. R., and Hirschberg, M. H. in Design for Elevated Temper-
ature Environment, The American Society of Mechanical Engineers, 1971, pp. 12-28.
[27] Smith, K. N., Watson, P., and Topper, T. H., Journal of Materials, Vol. 5, No. 4,
Dec. 1970, pp. 767-778.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
A. E. Gemma,' B. S. Langer,' and G. R. Leverant^
KEY WORDS: thermal fatigue, fatigue tests, nickel alloys, crack propagation
' Senior assistant project engineer, senior materials engineer, and group leader, respectively,
Pratt & Whitney Aircraft, Middletown, Conn. 06457 and East Hartford, Conn.
^ The italic numbers in brackets refer to the list of references appended to this paper.
199
mental conditions under which the crack growth data has been obtained.
The empirical parameters, that is, the pre-exponential term and the ex-
ponent of the Paris-type crack growth relationship, encompass the mate-
rial and environmental information. Some investigators [2-4] have studied
the influence of various mechanical properties on fatigue crack propaga-
tion; however, Pearson [5] was among the first to recognize the strong in-
fluence of Young's modulus on the fatigue crack growth rate. Richards
[6] found a correlation between grain orientation and room-temperature
fatigue crack growth rate in silicon iron, and he attributed the effects to
the variation in the ultimate tensile strength and Young's modulus.
Woodford and Mowbray [7] conducted thermal fatigue tests using tapered
disks in a fluidized bed. They found that the crack growth rate correlated
with grain orientation for directionally solidified IN 738. The variation in
crack growth rate was attributed to carbide alignment and dendrite orien-
tation.
Rau, Gemma, and Leverant [8] have shown that, for strain-controlled
thermal-mechanical fatigue cycling of high strength turbine alloys, crack
growth may be analyzed on the basis of a linear elastic fracture mechanics
approach.' This paper will extend this approach to a directionally solidified
nickel-base superalloy tested at various angles to the grains growth direc-
tion. It will be shown that the orientation dependence of the modulus of
elasticity can be used to normalize the power-type crack propagation rela-
tionship. Thus, the dependence of crack growth behavior of an anisotropic
alloy on the angle between the grain growth direction and the tensile axis,
9, can be predicted from a single test of longitudinal (0 = 0 deg) material.
' The data in Ref 8 and much of the data in the present study are plotted as a function of the
strain intensity range [A^. = A£ V naf (geometry)] rather than the usual stress intensity range.
This is done primarily for convenience since thermal fatigue is a strain-controlled property,
and lifetime prediction of components is more easily accomplished when deaUng in terms of
strain distributions. (As shown later in Fig. 6, the stress intensity factor is still considered to be
the prime driving force for crack propagation.)
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
GEMMA ET AL ON A NICKEL-BASE SUPERALLOY 201
Growth Axis
and 1600°F (871 °C) for 32 h prior to machining. The B-1900 plus hafnium
specimens were machined from hollow bars and were heat treated at
1975°F (1079°C) for 4 h and 1650°F (899°C) for 10 h. The Mar-M2(X)
plus hafnium grain size ranged from approximately 0.040 to 0.150 in. (1
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
202 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
to 3.8 mm) and the B-1900 plus hafnium grain size was equal to or less
than 0.032 in. (0.8 mm) with preferred growth in the radial direction.
Thermal-mechanical fatigue tests were performed in strain control on
tubular specimens (Fig. 2a) with a 1 in. (2.54 cm) uniform section gage
length in closed loop testing machines. Axial strain was measured with a
calibrated linear variable differential transformer (LVDT) and its associ-
ated extensometry (coaxial quartz tubes) which rested on ridges machined
on the inside diameter of the specimen gage section, as shown in the cut-
away view in Fig. 2b. Temperature was controlled £md measured with the
Bs^SaT
FIG. 2—Strain control fatigue specimen: (a) macroetched Mar-M200 plus hafnium. The
angle between the grain growth direction and the tensile axis, 9, is 45 deg and (b) cutaway
view of specimen with quartz extensometer in place.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized
GEMMA ET AL ON A NICKEL-BASE SUPERALLOY 203
Experimental Results
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions autho
204 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
TRANSVERSE
DIRECTION
FIG. 3—(a) Schematic of thermal-mechanical fatigue cycle and (b) schematic of loading
and grain growth axes.
TABLE 1—Elastic modulus, yield strain, and yield stress of DS Mar-M200 plus hafnium and
conventionally cast B-1900 plus hafnium.
Temperature, °F 0 15 30 45 90 Cast
800 17 20 26 34 20 25.4
1500 14 17 23 30 17 22.2
1700 12 15 20 26 16 20.8
1900 11.5 13 18 23 12.5 19.4
Yield Strain," %
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
G E M M A ET A L O N A NICKEL-BASE SUPERALLOY 205
1 X10-'
HALF CRACK
GROWTH RATE
da/dN
(INCHES/CYCLE)
IX 10-«
5 I 7 I 910
STRAIN INTENSITY R A N G E , A K C ( V 1 N ! X 1 0 ' ' )
FIG. 4—Dependence of the fatigue crack propagation rate on the angle between the tensile
axis and the grain growth direction.
about a factor of eight faster than those for the longitudinal material
(0 = 0 deg). Conventionally cast Mar-M200 plus hafnium was not tested
in this investigation but would be expected to show crack growth rates
similar to conventionally cast B-1900 plus hafnium.
Fractography
Crack growth was transgranular and perpendicular to the tensile axis
(Stage 2) for all DS Mar-M2(X) plus hafnium orientations. At low strain
intensity ranges (tJC^, the crack propagation path was smooth. With in-
creasing tJCi, Fig. 5, the amount of interdendritic cracking increased,
resulting in a much rougher fracture surface for all orientations. The
strain intensity range for the smooth to rough fracture surface transition
decreased in the following order: 9 = 0, 15, 30 or 90, and 45 deg, Table
2.
Discussion
Background
The stress intensity factor of a crack oriented at an arbitrary angle
to the principal material directions of an orthotropic plate is not meaning-
ful within the now classical Irwin approach to fracture mechanics. Wu
[10,11] has shown that the stress intensity factor for a crack in an ortho-
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
206 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
f T
(b)
(c)
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
GEMMA ET AL ON A NICKEL-BASE SUPERALLOY 207
TABLE 2—Values of various fracture mechanics parameters at the transition from a smooth
to a rough fracture surface in DS Mar-M200 plus hafnium.
0 15 30 45 90
Strain intensity range A*:., sAm. X 10"' 1.76 1.75 1.31 1.23 1.26
1.96 1.34
avg 1.86 1.30
Estimated crack opening displacement
COD, in. X 10-' 6.0 8.8 6.0 6.0 7.2
7.6 6.9
avg 6.8 7.0
tropic plate is reducible to the Irwin approach, that is, uncoupled modes,
only if the crack is oriented along one of the principal material directions,
and Bowie and Freese [12] allude to the difficulties which arise when lines
of elastic symmetry do not coincide with the coordinate axes. Most
published analyses are specialized to avoid these difficulties and so fall
within the Irwin scheme of fracture mechanics [10-14]. Gandhi's analysis
[75] of an arbitrarily oriented crack in a finite orthotropic sheet indicates
a strong dependence of the stress field of an inclined crack on material
properties and the aspect ratio of the finite orthotropic plate. In addition.
Cook and Rau [76] critically examined the concepts of linear elastic
fracture mechanics and concluded that orientation and plasticity effects
were of significance in fatigue crack extension.
In light of these difficulties, a heuristic approach was taken to estimate
the crack propagation behavior of a directionally soUdified nickel-base
superalloy. The objective of the analysis was to retain the linear elastic
isotropic definition of the stress intensity factor by introducing the effects
of grain orientation in the Paris-type crack propagation relationship.
* The effective stress intensity range (A/fo)eff is derived from the load history of a strain-
controlled test and is calculated in exactly the same manner as the standard stress intensity
range for load-controlled cycling. It is not strictly equivalent to the stress intensity range
calculated from a load-controlled test because of the presence of complex time-dependent
plasticity at the crack tip during cycUng at elevated temperature. As a result, {U(o)crf would
not be material independent or independent of orientation for an anisotropic material as
required by the linear elastic fracture mechanics definition of stress intensity range.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
208 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
CyCLE SHAPE
HALF CRACK
GROWTH RATE,
da/dN
(INCHES/CYCLE)
/ g» = 90'
' 0 6 = 45*
D ClJ=30"
a j A9<a5-
— 9 = 0 * SCATTER >AND
J I I I II
FIG. 6—Convergence of the crack growth rates when plotted as a function of effective
stress intensity range.
one test of a specific orientation. The data in Table 1 indicate that there
is some relationship between the anisotropic elastic modulus and crack
growth rate. This approach is developed as follows.
A plate of directionally solidified alloy may be considered to be ap-
proximately transversely isotropic if the transverse columnar grain
orientation (<100> to <110> in the case of directionally solidified nickel-
base superalloys) is random. Thus, any lamina of the plate when loaded
uniaxially in the 1-1 direction, is orthotropic. The ratio of the longitudinal
modulus, EL, to the modulus of the load axis, £•, is [77]
EL I EL \
A"' = — = cos "0 + — sin''0 + ( IVLT ) cos^e sin^e
ET \GLT J
where
0 = angle between the load and grain growth axes,
E-T = transverse modulus (0 = 90 deg),
GLT = shear modulus, and
VLT = Poisson ratio.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
GEMMA ET AL ON A NICKEL-BASE SUPERALLOY 209
B' W 30'WI5V
HALF CRACK
GROWTH RATE
da/dN
(INCHES/CYCLES)
C«ICUI«T10 ( X A K , )
TEStOMH
O S = 0•
^ 9 • IS-
a e = 90-
C) 8 = 30*
O 8 • 45-
4X10
20 30 4.0 5.0 G.O 10 10.0
AK.cyliix 10')
COD = iiKJ/4Ea,
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions a
210 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
where
AATo = stress intensity range,
E = Young's modulus, and
Oy = yield stress [18].
The values for the estimated COD for various Q calculated with the
800 °F material constants are tabulated in Table 3 for a constant crack
growth rate of 2.5 x 10^* in./cycle. The calculated COD is reasonably
constant, with the exception of the value for 0 = 45 deg.
TABLE 3-—Crack opening displacement (COD) for ar given crack growth rate.
0 15 30 45 90
Estimated COD at a
constant growth rate
of2.5 X lO-Mn./cycle 1.43 x :lO- 1.73 X 1 0 - 1.53 X 10-" 0.86 X 10"" 1.95 X l O -
The ability of the elastic modulus to normalize the crack growth data
must be associated with the form of the equation for COD. Oy is a weak
function of d (Table 1), and there is a direct relationship between elastic
modulus and stress intensity range experienced by the crack tip. Since
the loading conditions for these tests resulted in no net section yielding
(the one exception being moderate shakedown in the initial portion of
the 45-deg orientation), the elastic modulus and AA",, are directly pro-
portional such that the A/f„^ term in the COD equation dominates.
Fractography
It was reported in the results section that the transition from a smooth
to a rough fracture surface occurred at decreasing AKe in the order
0 = 0, 15, 30 and 90, 45 deg. Table 2. Calculation of the COD at the
transition using the effective stress intensity range shows that the esti-
mated COD is essentially constant (6.0 to 8.8 x 10"' in.) at this location,
Table 2. This is consistent with the observation that COD is the control-
ling factor in determining the crack growth rate for various d.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions au
GEMMA ET AL ON A NICKEL-BASE SUPERALLOY 211
ditions. In these studies, fatigue crack growth rate decreased with in-
creasing elastic modulus. An example is shown in Fig. 8 for Udimet 700
[19], a superalloy similar to Mar-M200 and B-1900. For stress-controlled
n - ' .Y
1 X10-
S 4-
3-
ixio-'4fi5 i
6 7 t 9 10 IS 20 30
STRESS INTENSITY RANGE, AKJKSiyiN)
FIG. 8—Crack growth rate as a function of stress intensity range for single crystal and
wrought Udimet-700 tested in pulsating tension (R = 0.1) at nOO'F (927°C).
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
212 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Summary
It has been established that thermomechanical fatigue crack growth
rates in a directionally solidified nickel-base superalloy are a minimum for
loading parallel to the grain growth direction (0 = 0 deg). Crack growth
rates increase for other orientations in the sequence 0 = 15, 30 or 90,
and 45 deg until equivalence is reached with rates for conventionally-cast
superalloys.
It was demonstrated that a linear elastic fracture mechanics approach
can be extended to the prediction of strain-controlled thermomechanical
fatigue crack growth in an anisotropic material of this type. Prediction of
crack growth rates as a function of the angle between the loading direc-
tion and the grain growth direction can be achieved by utilizing the elastic
modulus as a normalization factor. The relative crack growth rates as a
function of orientation (0) and the smooth-to-rough transition in fracture
surface appearance can be explained by the orientation dependence of the
COD.
References
[1] Paris, P. C , Gomez, M. P., and Anderson, W. E., The Trend in Engineering, Vol. 13,
Jan. 1961, p. 9.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
GEMMA ET AL ON A NICKEL-BASE SUPERALLOY 213
[2] Grosskreutz, J. C , Physica Status Sotidi (b), Vol. 47, No. 1, 1971, pp. 11-31.
[3] Grosskreutz, J. € . , Physica Status Solidi (b). Vol. 47, 1971, pp. 359-396.
[4] Hickerson, J. P., Jr., and Hertzberg, R. W., Metallurgical Transactions, Vol. 3, Jan.
1972, pp. 179-189.
[5] Pearson, S., Nature, Vol. 211, Sept. 1966, pp. 1077.
[5] Richards, C. E., Acta Metallurgica, Vol. 19, July 1971, pp. 583-596.
[7] Woodford, D. A. and Mowbray, D. F., Materials Science and Engineering, Vol. 16,
No. 1/2, Oct./Nov., 1974, pp. 5-43.
[8] Rau, C. A., Jr., Gemma, A. E., and Leverant, G. R., in Fatigue at Elevated Tempera-
tures, ASTMSTP520, American Society for Testing and Materials, 1973, pp. 166-178.
[P] Erdogen, F. and Ratwani, M., International Journal of Fracture Mechanics, Vol. 6,
No. 4, Dec. 1970, pp. 379-392.
[10] Wu, E. M., Journal of Applied Mechanics, Dec. 1967, pp. 967-974.
[77] Wu, E. M. in Composite Materials Workshop, S. W. Tsai, J. C. Halpin, and N. J.
Pagano, Eds., Technomic Publishing Co., Stamford, Conn., 1968, pp. 20-43.
[12] Bowie, O. L. and Freese, C. E., International Journal of Fracture Mechanics, Vol. 8,
1972, pp. 49-58.
[13] Sih, G. C , Paris, P. C , and Irwin, G. R., International Journal of Fracture Mechanics,
Vol. 1, No. 3, 1965, pp. 189-203.
[14] Walsh, P. F., Engineering Fracture Mechanics, Vol. 4, 1972, pp. 553-541.
[15] Gandhi, K. R., Journal of Strain Analysis, Vol. 7, No. 3, 1972, pp. 157-162.
[16] Cook, T. S. and Rau, C. A., " A Critical Review of Anisotropic Fracture Mechanics,"
Proceedings, International Conference on Prospects of Fracture Mechanics, Delft
University of Technology, Netherlands, 24-28, June 1974.
[17] Calcote, L. R., The Analysis of Laminated Composite Structures, Van Nostrand Rein-
hold Co., New York, 1969, pp. 29-31.
[18] Hahn, G. T., Rosenfield, and Sarrate, M., "Elastic-Plastic Fracture Mechanics,"
AFML-TR-67-143, Part III, Air Force Materials Laboratory, Jan. 1970.
[19] Rau, C. A., Pratt & Whitney Aircraft, unpublished data.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
K. D. Sheffler'
KEY WORDS: thermal fatigue, fatigue (materials), iron alloys, tests, stainless steels
' Assistant project engineer. Materials Engineering and Research Laboratory, Pratt and
Whitney Aircraft, East Hartford, Conn. 06108.
'Sheffler, K. D. and Doble, G. S. in Fatigue at Elevated Temperatures, ASTM STP 520,
American Society for Testing and Materials, 1973, p. 491.
'Sheffler, K. D. and Doble, G. S., "Influence of Creep Damage on the Low Cycle
Thermal-Mechanical Fatigue Behavior of Two Tantalum-Base Alloys," Final Report, Con-
tract NAS3-13228, NASA-CR-121001, National Aeronautics and Space Administration,
May 1972.
214
Experimental Details
To separate the effects of temperature and environment on fatigue be-
havior, all tests were conducted in an ion-pumped ultrahigh vacuum
chamber at pressures below 1 x 10"' torr. Elevated temperature iso-
thermal fatigue tests were conducted at two frequencies (0.65 and 0.0065
Hz) to evaluate the contribution of frequency induced creep effects to the
fatigue process. Test temperatures for the isothermal tests were 1200°F
(922 K) for the Type 304 stainless steel and 1100 °F (866 K) for the A286
alloy. Tests with combined temperature and strain cycUng were of two
types, as illustrated in Fig. 1. The in-phase cycle involved isothermal
tensile strain imposed at a high temperature reversed by isothermal
compressive strain imposed at a low temperature. The out-of phase cycle
was similar except that the temperature-strain phasing was reversed. The
upper temperature for each material was the same as the isothermal
test temperatures. The lower temperature was 600°F (589 K) for both
alloys. Tests were conducted to failure (defined as a separation of the
specimen into two pieces) over a range of plastic strain amplitudes which
were measured by the width of hysteresis loop at zero load. Fractured
specimens were sectioned longitudinally and examined metallographically
to evaluate the character of the microstructural damage associated with
each of the applied cycle types.
Test procedures were essentially identical to those reported for pre-
vious tests.^^ The apparatus was designed to perform completely reversed
push-pull fatigue tests on hourglass specimens using independently pro-
*Manson, S. S. in Fatigue at Elevated Temperatures, ASTM STP 520, American Society
for Testing and Materials, 1973, p. 744.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
216 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
1fi r i I
II 1 1
M 1 I
Time ——
a .™ b
Test Results
Typical hysteresis loops generated in Type 304 stainless steel are shown
in Fig. 2. The isothermal loops are essentially symmetrical for both ma-
terials, with the load range developed at a constant diametral strain range
being larger at the higher test frequency. Hysteresis loops observed for
A286 were similar except that the inelastic strain ranges were smaller at
equivalent total diametral strain ranges because of the larger loads and,
consequently, larger elastic strains developed in the higher strength ma-
terial. Asymmetric hysteresis loops were developed with both types of
thermal fatigue cycles as a result of differences in flow stress at different
temperatures. In-phase cycling generated loops having a net compressive
stress, while out-of-phase cycling caused a mean tensile stress to be de-
veloped.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
SHEFFLER ON IRON-BASE ALLOYS 217
lOk SS
APPROXIMATE
TOTAL LONGITUDINAL
STRAIN RANGE
OF .008 I N . / I N .
+1)0
+30
H—I—h H 1—h
000 3000
DIAMETRAL STRAIN
/4IN./IN.
100
-30
-1*0
FIG. 2—Typical hysteresis loops observed for Type 304 stainless steel tested with various
strain-temperature cycles.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
218 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
t ORIGINA). tPECIHEN
3NT0UR ( 1 - 1 / 2 " RADIUS)
ORIGINAL HINIHUH
DIAMETER ( l A " )
ZONE OF HAXIHUH
HICROSTRUCTURAL
DAMAGE
minimum diameter (Fig. Ad) where strain was both unknown and uncon-
trolled. This qualification applies only to the out-of-phase Type 304 stain-
less steel results since geometry changes were not observed in the A286
alloy.
Fatigue life results for both alloys are summarized in Tables 1 and 2.
Type 304 stainless steel displayed cycUc strain hardening, while the A286
alloy exhibited cyclic strain softening. For both alloys, the majority of
hardening or softening occurred during the first few cycles of testing. The
stress ranges noted in Tables 1 and 2 represent stabiUzed loop sizes. For
the thermal cycled tests, where asymmetric loops were observed, both
tensile and compressive stress values are noted.
The longitudinal inelastic strain ranges noted in Tables 1 and 2 are
plotted against life in Figs. 5 and 6. These results indicate definite ef-
fects of both frequency and thermal cycling on fatigue life. Decreasing
frequency, which increases the potential for creep effects, reduces the iso-
thermal fatigue life of both materials. Combined temperature and strain
cycling causes further reductions of fatigue life in both materials as com-
pared to isothermal cycling at the maximum temperature associated with
the thermal cycle. For the A286 alloy, in-phase cycling is more damaging
than out-of-phase cycling. For Type 304 stainless steel, the reverse at
first appears to be true; however, the out-of-phase Type 304 results are
confused by the previously noted geometry change and are therefore not
considered meaningful.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorize
SHEFFLER ON IRON-BASE ALLOYS 219
I
If
If
II
il
!i|1
ft
S.i
i
5I
i
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
220 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
-O r- (N ^o w-i I
<N -"J- "j-i M (*) >0
— 00 r- (S I
oo 00 o f*! <*)
00 r-i v"! 1^ ^ oo
(j\ oo r- <N — CT\
r^ (N (N r-) rs —
O 0\
r^ » ^ r4
3 S
\0 ^O ^ \0 '"O
fl — 0\ »N —< 00
** ^ "1 m m (N
3
.50
o •= I S
iiiiiiiiiii
ON 0 \CT\ON 0^
I
BQ
<
888 88 8 8 8 888
tr\ *r\ tn
mmm
^i < <
< < <
< < M
< <
<< << << < < -t
Z.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
SHEFFLER ON IRON-BASE ALLOYS 221
^ O ^ 0\ w^
irj ry <*) r- >0
. . '. . : o o oo r- r~-
*N Q ^ •* r^
0\ '•D a^ tri —
r- ^ I/-, o oo
W^ Tf C-I -^ —
v^ <n \ 0 f^ oo
-H Ov oo K-l —
I
I o8S
d d o
08
do
088
o d d
08
dd
T ^ O^ ON
00 00 00
u-i ir\ in
O^
00 00
*n v~i
ON
cs
0 0 0 0 0 0 OOOO OOOOOO 0 0 0 0
:3
888 88 8 8 8 88
0 0 0 0 0 0 0 0 0 0
II mHiiiii
11
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
222 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
"T 1 1 1 1 I I ~i—I—I I I IJ
SOTHERMAL I 2 0 0 ° F c y c l e , 0 . 6 5 Hz
J I I ' l l J I I I I I I Ii J I I I I I I
CYCLES TO FAILURE
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
SHEFFLER ON IRON-BASE ALLOYS 223
Microstructural Observations
Microstructural damage observed in the fatigue tested specimens varied
with both cycle and test material, as shown in Figs. 7-9. In high fre-
quency cycling, both alloys exhibit mixed intergranular and transgranular
cracking (Fig. 9) with minimal microstructural damage localized at grain
boundaries (Figs. So and 9a). As shown in Fig. 9d, some of the secondary
cracks found in A286 are associated with carbide inclusions. While car-
bide cracking is assumed to have some influence on the cyclic life of
*>l^i^
20K.n
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
224 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
tf^mil
^ 3 ^ w - • •;•;*.•
Em.*' ^r.
(a) Isothermal 0.65 Hz.
(b) Isothermal 0.0065 Hz.
(c) In-phase cycle.
(rf) Out-of-phase cycle.
FIG. 8—Microstructural damage found in A286 tested with various cycles. Tensile axis
vertical.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorize
SHEFFLER ON IRON-BASE ALLOYS 225
m^:.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions auth
226 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Summary
Elevated-temperature, low-cycle thermal-mechanical fatigue tests con-
ducted in vacuum on Type 304 stainless steel and A286 alloy show sig-
nificant effects of frequency and of combined temperature-strain cycling
on fatigue life. At temperatures in the creep range, fatigue lives of both
materials are lower at 0.0065 than at 0.65 Hz. Metallographic examina-
tion of fractured specimens indicates mixed mode (intergranular and
transgranular) fracture at the higher frequency and exclusively inter-
granular fracture at the lower frequency. In-phase thermal cycling (ten-
sion at high temperature and compression at low temperature) causes
large life reductions in both materials. These Hfe reductions are attributed
to grain boundary cavitation caused by unreversed grain boundary sliding
(grain boundary ratcheting). Out-of-phase thermal cycling (tension at low
temperature and compression at high temperature) also causes large cyclic
life reductions in both materials, but the results from Type 304 stainless
steel are confused by the occurrence of geometric instabilities. In the
A286 alloy, out-of-phase life reductions are attributed to cavitation dam-
age resulting from unreversed compressive grain boundary displacements
(compressive ratcheting) which cannot be fully accommodated by intra-
granular deformation.
A cknowledgments
Support for this work was provided by the NASA Lewis Research Cen-
ter, Contract NAS-3-6010, with Dr. G. R. Halford as program manager.
Assistance with the experimental effort, provided by J. W. Sweeney, is
gratefully acknowledged.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authoriz
L. F. Coffin'
Instability Effects in
Thermal Fatigue
227
tt Oil-- -
m"^'-^
a-2-28 •i-2-1^7 H-2-50
538°v
6.1,-^ i.ti:^
no hold time • 56 s e c ( t e n s . ) 56 sec(eoinp.}
Nf= 176 Nf= 1,152 N^= 3,180
resulting from compression hold. Here rfo refers to the position of the
diametral extensometer. Using similar specimen geometries and strain
control procedures, the author has shown that AISI 304 stainless steel,
tested at 650 °C with unequal (sawtooth) total longitudinal strain ramp
rates, developed similar shapes [5]. With a plastic strain range, A£p, of
0.02, a period of 10 min, and a ramp whose tension going time was 0.1
min, followed by a compression going ramp time of 9.9 min, off-center
necking was found, while reversing the ramp rates produced off-center
barreling.
Shape instability was found in thermal fatigue tests performed on axi-
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
COFFIN ON INSTABILITY EFFECTS 229
ORIGINAI. SPECIMEN
CONTOUR ( 1 - 1 / 2 " RADIUS)
ORIGINAL HINIHUn
DIAMETER ( l A " )
ZONE OF MAXIMUM
MICROSTRUCTURAL
DAMAGE
FIG. 2—Schematic of shape changes observed with in-phase and out-of-phase thermal and
mechanical cycling of AISI 304 stainless steel. See Refl.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
230 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
• (a) EFFECT OF
Z STRAIN HARDENING
^ (b) EFFECT OF
Z STRAIN RATE
SENSITIVITY
• - ( c ) EFFECT OF
Z TEMPERATURE
result of the changes in the flow stress curves for the parameters consid-
ered here, the axial strain distribution may be diffuse or locaUzed with ap-
plication of axial stress. The locaUzation or diffuseness can also be ob-
served by comparing, in each case, the axial strains at Position 2 for equal
strains at 1.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
COFFIN ON INSTABILITY EFFECTS 231
the axial strain distributions shown in Fig. 3 to the separate legs of spe-
cific hysteresis loops. Three cases will be considered: (a) isothermal bal-
anced cycling (Fig. 4), (b) isothermal unbalanced cycling (Figs. 5,6), and
(c) thermomechanical cycling (Fig. 7c).
IT, IC
I^'PI
BALANCED ISOTHERMAL LOOP
•'ZCi ^-ZTl
OFF-CENTER BARRELING
FIG. 4—Cyclic stress-strain behavior and hysteresis loop for balanced isothermal cycling
showing conditions for off-center barreling.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
232 T H E R M A L FATIGUE O F MATERIALS A N D COMPONENTS
FIG. 5—Cyclic stress-strain behavior and hysteresis loop for unbalanced cycling produced
by isothermal slow-fast or in-phase thermal cycling showing conditions for off-center barrel-
ing.
Fig. 4, these points are identified. Also shown in Fig. 4 is the cyclic stress-
cyclic plastic strain amplitude response of the material, |oJ versus le^pi.
The Points 2T and 2C are identified. The cyclic stress-strain curve shown
is the locus of stable loop End Points 1, either in tension or compression,
for all strain ranges. Since the loop is balanced, the tension and compres-
sion curves are identical. Now since |2C| > |27l, Itzpzcl > k^pzrl- Thus, the
loop for Point 2 does not close, but rather a slightly excessive axial com-
pressive strain is produced. This increment of strain accumulates with
each cycle leading to progressive off-center barreling.
The ratcheting process described is difficult to predict quantitatively.
As the compressive strain accumulates at 2, the flow stress increases and
the increment of plastic strain decreases because of strain hardening. Cy-
clic strain softening limits strain hardening, and a steady-state ratcheting
rate is reached.
We can qualitatively determine the response of various materials to
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions author
COFFIN ON INSTABILITY EFFECTS 233
/HIGH €
'T \ L O W TEMP
LOW €
IC 1 HIGH TEMP
l^'J
ISOTHERMAL FAST-SLOW LOOPS
OR
OUT-OF-PHASE THERMAL FATIGUE LOOP
FIG. 6—Cyclic stress-strain behavior and hysteresis loop for unbalanced cycling produced
by isothermal fast-slow or out-of-phase thermal cycling showing conditions for off-center
necking.
shape changes under these conditions. First, a high strain hardening coef-
ficient is desirable in order to enhance the off-center strain range. Sec-
ondly, pronounced cyclic strain softening is necessary in order that the
strain hardening effects which would otherwise limit the ratcheting strains
be minimized. Because of these considerations, 1100 aluminium is particu-
larly sensitive to shape changes [7,5].
Isothermal Unbalanced Cycling
We now consider specimen shape changes for isothermal loops of un-
equal ramp rates, called slow-fast (Fig. 5) or fast-slow (Fig. 6). We first
note the hysteresis loop generated for the slow-fast ramp condition illus-
trated and observe that its unbalanced nature arises from the material's
response to strain rate. Fig. 3b [5], The cyclic stress-plastic strain ampli-
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorize
234 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
MECHANICAL STRESS
STRAIN OR TEMR
"ITEMP
TIME
,STRAIN
IN PHASE
STEPPED TEMPI
I 1 I TEMP
I
.STRAIN
L_J I I
OUT-OF-PHASE
STEPPED TEMP
DECREASING
TEMP
v
^STRAIN
OUT-OF-PHASE^
RAMPED TEMP
FIG. 7—Waveshapes for mixed mechanical and thermal cycling for in-phase and out-of-
phase conditions.
tude curves for the tensile and compressive stress portions of the loop (lo-
cus of Points i r and \C) are also shown. Since the tensile portion of the
cyclic stress-strain curve is controlled by the slow tension-going part of
the loop, it differs distinctly from the compressive portion, as indicated.
Values of stresses and plastic strain are shown for the off-center Position
2 as indicated, the stress assumed to be 90 percent of the stress at Position
1. No correction has been made for load differences at Position 1 between
tension and compression nor for differences in strain rate between Posi-
tions 1 and 2. Strain rate difference effects between 1 and 2 can be
avoided if the cyclic stress-plastic strain curve is obtained under condi-
tions of constant frequency. From Fig. 5 we note that |£,p2cl > kzparl; that
is, for each closed cycle of strain at Position 1, there is a net compressive
strain increment developing at Position 2, causing the specimen to become
progressively fatter there. Simultaneously, there will be a net shortening
of the specimen.
Conversely, the fast-slow loop produces just the opposite effect. Fig. 6.
The unbalanced cyclic stress-strain loci for Point 1 are now reversed, the
tensile curve being generated at high strain rate. The off-center stresses
are located at 90 percent of the end point values and, because of the slope
differences, |£z„2rl > k zp2C\ Thus, at Position 2, there is a progressive
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions author
COFFIN ON INSTABILITY EFFECTS 235
Special Cases
Earlier, the case of end thinning in axially constrained and thermally
cycled tubular specimens was referred to [6]. This situation requires a dif-
ferent treatment since the end thinning (and central thickening) was occur-
ring in a region of uniform stress. Nonuniform temperature along the
tube length, especially near the ends, was found and was particularly
severe at the higher temperature (600 "C) but was insignificant at the lower
temperature (100 °C). The end thinning and center thickening can be
traced to this axial temperature distribution. Since compression exists at
high temperature, greater thickening in the wall occurs there rather than
at the ends where the flow stress is greater. At the lower temperature and
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
236 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
the accompanying tensile stress, the ends are thinned further because of
the now developed nonuniform wall thickness. There is little to prevent
the process from continuing.
Another case mentioned earlier is that reported by Conway, Stentz, and
Berling [4] for isothermal hold-time tests on zirconium-copper in argon at
538 °C. It should first be noted that this temperature is very high for cop-
per alloys. Secondly, the shape changes were reported to occur at 1.4 per-
cent total strain range, but not at 5.0 percent. Hence the behavior is not a
general phenomenon. Because of the high temperature and very low strain
hardening, most of the plastic deformation during the ramping of the
strain is localized to a region very near the specimen center, as illustrated
by Fig. 3. Thus, at off-center Position 2, the cyclic plastic strain com-
ponent is very small. The tensile hold period requires that the total com-
puted longitudinal strain at Position 1 remain constant. The specimen is
subjected to a prolonged, relaxing tensile stress, allowing tensile creep de-
formation to occur at Position 2. Correspondingly, there will be elastic
unloading at 2 consistent with the constant total (elastic plus plastic)
strain conditions at Position 1. With the next strain reversal, again little
plastic deformation occurs at 2, but another creep increment develops
during the hold period. Hence, progressive necking develops at 2. As the
diameter becomes smaller and the stress increases at 2, the creep effects
accelerate, although no indication of this event is evident at the control
position. However, the specimen becomes progressively longer.
In the case of compression hold, the same process takes place, but in
compression. Off-center barreUng develops as a result of repetitive incre-
ments of compressive creep. Here, the stress decreases with increasing
barreling, so that the process tends to slow down at 2. However, the
shape change of the specimen causes the bulging process to move down
closer to the specimen center. Eventually the barreling extends almost to
the center, and the profile becomes highly localized. The highly concen-
trated necking and barreling is clearly seen in Fig. 2, together with spec-
imen shortening.
In the case of the hold experiments performed at higher strain ranges,
the zone of plastic deformation extends further from the center. The dif-
ferences in rapid straining in one direction and creep in the other direction
are now less, and the shape changes are smaller. Additionally, there are
fewer cycles.^
' Note added in proof: attention should also be called to the work of Skelton [J3] in which
uniform gage length test specimens of 20Cr/25Ni/Nb steel were subjected to longitudinal
strain control at 750 °C with tension and compression hold periods. It was found that tension
hold periods led to barreling at the midgage length position and necking near the end fillets,
while compression holds produced midlength necking and barreUng near the end fillets.
These observations are consistent with those of Conway, Stentz, and Berling [4].
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
COFFIN ON INSTABILITY EFFECTS 237
Discussion
The general condition that appears necessary for shape changes and in-
stabilities in mechanical and thermal cychng is a positional variation or
gradient in the cyclic stress-strain field caused by geometry, temperature,
or material variations. We have cited several examples of geometry varia-
tions where hourglass specimen geometries Eire used and one example of
temperature variation, that of the axially constrained and thermally cycled
tube. Instability can be minimized by maintaining axial uniformity both in
geometry and in temperature distribution wherever cyclic plastic strain is
present.
Although the observations and interpretations cited here relate to spec-
ific bulk geometries, it is interesting to speculate as to whether similar
geometric changes can occur on a much finer scale. Two intriguing ques-
tions come to mind. One relates to whether the intense strain locahzation
associated with slip-band extrusion and crack initiation develops as a con-
sequence of progressive microstructural changes in shape of surface
grains. Another deals with crack growth and the shape of the crack tip as
a function of the wave shape employed. Relative to the latter question,
substantial differences in the shape of mildly notched specimens occur de-
pending on the phase relationship between thermal and mechanical load-
ing (Fig. 5). Similar shape changes can be envisaged when the notch
sharpens sufficiently to become a propagating crack. Various investigators
[2,5] have reported substantial differences in life as a function of wave
shape, other conditions being equal. Can these differences be related to
differentes in crack tip shape changes and thus on crack propagation?
Relative to crack nucleation, McClintock's [12] quantitative model of in-
stability to explain surface roughness and crack nucleation should be re-
called. That work was inspired by observations of roughenmg on cyciea
plasticine models. Northcott and Baron's studies on thermal fatigue can
also be cited, where narrow edges of wedge-shaped specimens were alter-
nately heated and cooled to produce a regular pattern of deep grooves,
initiating crack propagation. Crazing in thermeil fatigue and local surface
roughening in isothermal fatigue may be closely related to the hourglass
specimen shape instabilities reported here. The craze-cracking problem is
unquestionably more compHcated, dealing with cyclic phase transforma-
tions, interplay with oxidation processes, two-dimensional stresses, etc. It
is, however, basic to many thermal fatigue problems and should be
studied in more detail.
References
[/] Coffin, L. F., Transactions, American Society for Mechanical Engineers, Journal of
Basic Engineering, D, Vol. 82, 1960, p. 671.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
238 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
[2] Sheffler, K. D., "Vacuum Thermal-Mechanical Fatigue Testing of Two Iron Base Al-
loys High Temperature Alloys," NASA-CR-134524, National Aeronautics and Space
Administration, Lewis Research Center, Cleveland, Ohio, Jan. 1974.
[3] Raymond, M. H. and Coffin, L. F., Acta Metallurgica, Vol. 11, 1963, pp. 801-807.
[4] Conway, J. B., Stentz, R. H., and Berling, J. T., "High Temperature, Low-Cycle Fa-
tigue of Copper-Base Alloys in Argon," NASA-CR-121260, National Aeronautics and
Space Administration, Lewis Research Center, Cleveland, Ohio, Aug. 1973.
[5] Unpublished General Electric Co. research.
[6] Coffin, L. F., Transactions, American Society of Mechanical Engineers, Vol. 76, 1954,
pp. 931-950.
[7] Coffin, L. F., Transactions, American Society of Mechanical Engineers, Vol. 79, 1957,
p. 1637.
[8] Carden, A. E., "Thermal Fatigue—An Analysis of the Experimental Method," ORNL
TM-405, Oak Ridge National Laboratory, Oak Ridge, Tenn., 1963.
[9] Thomas, W. S., "An Investigation of the Stages of Deformation of a Thin-Wall Tube
Caused by Cyclic Thermal Stress," M.S. Thesis, University of Alabama, University,
Ala., Aug. 1961.
[10] Hernandez, J. A., "An Analysis of the Thermal Fatigue Characteristics of Inconel,"
M. S. Thesis, University of Alabama, University, Ala., 1964.
[77] Sheffler, K. D. and Doble G. S. in Fatigue at Elevated Temperature, ASTM 520, Amer-
ican Society for Testing and Materials, 1973, pp. 491-499.
[12] McClintock, F. A. in Fracture of Solids, Interscience Publishers, New York, 1963, pp.
65-102.
[13] Skelton, R. P., Materials Science and Engineering, Vol. 19, 1975, pp. 25-30.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
G. R. Halford^ and S. S. Manson^
ABSTRACT: This paper describes the features of the method of strainrange partition-
ing that make it applicable to the life prediction of thermal-mechanical strain-cycling
fatigue. An in-phase (230 to 760 °C) test on Type 316 stainless steel is analyzed as an
illustrative example. The method utilizes the recently proposed step-stress procedure
of experimental partitioning, the interaction damage rule, and the life relationships
determined at an isothermal temperature of 70S°C. Implications of the present study
are discussed relative to the general thermal fatigue problem.
KEY WORDS: thermal fatigue, intergranular fracture, plastic strain, creep strain,
life prediction, strainrange partitioning, stainless steels
239
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
HALFORD AND MANSON ON LIFE PREDICTION 241
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized
242 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
HALFORD AND MANSON ON LIFE PREDICTION 243
Example Analysis
Material: AISI Type 316 stainless steel, hot rolled and mill annealed, 19 mm (0.75 in.) di-
ameter bar stock supplied by Viking Steel Co., Cleveland, Ohio. The heat number,
exact chemical composition, and annealing details are unknown. Tensile and creep-
rupture properties at 705 "C are given in Ref 9. An indication of the grain structure
can be found in Fig. 5a.
Specimen: Tubular, hourglass test section [20\, specimen number AYY-304.
Extensometer: Diametral [20].
Strain: Diametral strain cyclically ramped linearly with time between limits resulting in an
' axial total mechanical strainrange of 0.00923. The apparent thermal component of
strain (0.01381) was measured by temperature cycling the specimen under a cori-
trolled zero load condition. This thermal component was subtracted from the total
measured strain to determine the mechanical component. The elastic strainrange
was subtracted from the total mechanical strain range to get the inelastic strain
range of 0.00616. The cyclic period was 30 min.
Temperature: Silicon carbide internal heating element, temperature ramped linearly with
time between temperature limits of 230 and 760 °C with a period of 30 min. The
temperature was phased with the strain such that the maximum temperature oc-
curred at the peak tensile strain. This is referred to as an "in-phase" thermal-
mechanical strain cycle.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorize
244 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
A e , - A H = 0.00923
Aejn = 61 = 0.00616
TEMP: 230°-760° C
PERIOD • 1800 SEC
AISI TYPE 316 STAINLESS STEEL
FIG. 1—Stress-strain hysteresis loop for in-phase thermal-mechanical strain cycle illustrat-
ing stress levels and creep strains associated with experimental partitioning of strains by the
step-stress method.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
HALFORD AND MANSON ON LIFE PREDICTION 245
111
HU55
a
>•„ 2
t!
llf
5 u
i5 a3 1<s3 -
t^os'»
n^
>-i *0 • 00 O t>
oooo ' —vdodooooo
!lis
<^
so •t-i t«
c
•§ 1^12 :88 :S8S :S8
1s ^11
^ap a
•S e£:
§ — 5
•J- m
5> Tu? -H ( S <S <S| — —I — (S <S f l
-^
-*::
C
*-•
* i cj I I I I
^=4. M >.
U
c =»
1 IS o « vo <s 00 -^ •
»0 IS <S -" — I
^ ^ w% vo r^ I ills fS8 r^ 00
w 2«^
0»/^fnoo^o\»nQ(Sas'<tooo
H PI 4>
.s'R <pQOQWfcODa^-^^J<
o og
ft-
a> .
"y o
^z
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
246 T H E R M A L FATIGUE O F MATERIALS A N D C O M P O N E N T S
where
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions autho
HALFORD AND MANSON ON LIFE PREDICTION 247
9i-
-1800
u
<:
A R E A - a 00006
(STEADY STATE
COMPRESSIVE CREEP STRAIN)
^ ^
J
200 400 600 800 1000 1200 1400 1600 1800
TIME IN CYCLE, SEC
FIG. 3—Accumulation of tensile and compressive steady-state creep strain during in-phase
thermal-mechanical strain cycle (230 to 760°C) applied to AISI Type 316 stainless steel.
0.00471
0.00006
(2)
F,, = A£„/A£,„ = 0.010
0.00616
0.00139
F „ = A£„/A£,.„ = = 0.225
0.00616
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
248 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
Damage Rule
The strain fractions are used in the interaction damage rule which is
written as follows
N N N N N
•* ^pp ^ ^ cc ^ ^ cp •» ^pc •'• ^pr
where
Npp, Ncc, Ncp, and Np^ = cycUc lives determined from entering the life
relationships at a strainrange equal to the en-
tire inelastic strainrange of the cycle of inter-
est, and
Npr = predicted life.
This damage rule is based on assumptions described in detail in Ref 9.
Although all of the terms in the equation are Unear, we have continued to
use the term "interaction" in its designation in order to distinguish it
from the Unear damage rule used in Ref 8 and because of certain inter-
action assumptions entering the original derivation.
Life Relationships
The life relationships used in the life prediction of the example thermal-
mechanical strain-cycling problem are shown in Fig. 4. These curves are
based upon the use of the interaction damage rule and the interpretation
of creep strain as being only the steady-state (secondary) portion of the
time-dependent strain. All of the tests used to establish these life relation-
ships were conducted at a single isothermal temperature of 705 °C.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
HALFORD AND MANSON ON LIFE PREDICTION 249
105
CYCLES TO FAILURE
FIG. 4—Strainrange partitioning life relationships based on use of interaction damage rule
and steady-state creep strain.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authori
250 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
out-of-phase, one could expect exactly the same amount of creep and
plastic strain as measured previously but with the role of tension and
compression reversed. That is, the term Fcp/Ncp in Eq 4 would be replaced
with FpJNp^ where Fp<, is 0.225 and Np^ is 410 cycles. The predicted life
for out-of-phase cycling would be 870 cycles which is 2.6 times greater
than the predicted life of 334 cycles for in-phase cycling. The result of
this calculation is in line with our experience in conducting a number of
other thermal-mechanical strain-cycling tests on AISI Type 316 stainless
steel.''
Application of the time and cycle fraction approach as used in ASME
Code Case 1592 [16\ does not differentiate between in-phase and out-of-
phase thermal-mechanical straining since compressive creep and tensile
creep are taken to be equally damaging. This approach would predict the
same life for these two extremes in cycling which can lead to the following
undesirable situation: consider that the design curves were established
with just the proper amount of conservatism to handle an in-phase thermal-
mechanical cycle such as the one just discussed. By applying the same
criteria and design curves to an out-of-phase cycle, the designer would
predict an unnecessarily conservative life. An inefficient design results.
The foregoing discussion helps to illustrate the emphasis that strain-
range partitioning places upon some of the details within thermal-mechan-
ical strain cycles.
Assessment of Error
It should also be pointed up that these calculations do not include an
assessment of the added damage to the specimen as a result of the step-
stress partitioning procedures. A series of calculations have been made
to determine the damage done by having held a constant creep stress for
the time intervals listed in Table 2. Each cycle during which the creep
' A n out-of-phase test conducted at an inelastic strainrange of 0.00484 failed at 1449
cycles. The frequency and temperature range were the same as listed in Table 1. Since this
particular test was conducted prior to the development of the step-stress method, the parti-
tioning of the creep and plastic strains was not accomplished. However, since the strainrange
is only 20 percent less than that of the in-phase test, it is reasonable to assume that the
partitioning would have been approximately the same. Hence, the predicted life is given by
F F F 1
PP ' cc ^ pc *
A^ N N ~ N
*^ PP ^^ cc ^^ pc ^^ pr
0.765 0.010 0.225 1
2022 + 524
... + 540 N,r
1
0.000378 -t- 0.000019 + 0.000417 = 0.000814 =
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
HALFORD AND MANSON ON LIFE PREDICTION 251
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorize
252 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
INTERGRANUUR
CRACKS OPEWO
BY LAST CYCLE
(a)
aiMM
mRECTION
Of AXIAL
STRAIN
i:
INSire OlAM
OUTSIM OlAM
WALL THICKNESS WALL THICKNESS
FIG. 5—Photomicrographs of failed specimen of AISI Type 316 stainless steel subjected
to thermal-mechanical cycling (230 to 760°C) in phase.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
HALFORD AND MANSON ON LIFE PREDICTION 253
Conclusions
We have applied the method of strainrange partitioning to a thermal-
mechanical strain-cycling problem on AISI Type 316 stainless steel. The
sample problem can be used as an initial guide in applying the method to
the general thermal fatigue problem. To date, the approach has been
limited to cycles in which the inelastic strains can be determined with
some degree of confidence. Extension of the method would be required in
order to deal with cycles for which the elastic strains dominate and the
inelastic strains are too small to calculate or measure accurately.
Although the sample problem involved a material whose life relation-
ships are insensitive to temperature, some suggestions were given for
dealing with materials for which this is not the case. Obviously, further
experimentation is needed to test out the applicability of these suggestions.
Additional research is also needed to guide in the selection of appropriate
techniques for analyzing materials that exhibit metallurgical instabilities
such as dynamic strain-aging or that experience strong interactions with
the environment.
Extension of the method to predict the thermal fatigue life of an actual
component would not be a difficult task, provided analyses were available
for the temperature and equivalent axial strain histories at critical loca-
tions. An axially loaded laboratory specimen could then be programmed
to follow these histories. From that point on, the determination of the
creep and plastic strains, the partitioned strainrange components, etc.,
could follow the procedures described in this paper.
Rather than measuring the creep and plastic strains within a cycle as
illustrated in this paper, they could be calculated using procedures and
constitutive equations such as those proposed by Spera and Cox [19] and
Mowbray and McConnelee [27]. Of course, upper and lower bounds on
life can be estimated as discussed earlier without having to resort to either
experimental or analytical partitioning.
References
[7] Taira, S. in Creep in Structures, N. J. Hoff, Ed., Springer-Verlag, Berlin, 1962, pp.
96-119.
[2] Manson, S. S. and Spera, D. A., Quarterly Transactions of the American Society for
Metals, Vol. 58, 1965, pp. 749-751.
[3] Manson, S. S., IntemationalJoumal of Fracture Mechanics, Vol. 2, 1966, pp. 327-363.
KJ Manson, S. S. and Halford, G. R. in Thermal and High Strain Fatigue, The Metals
and Metallurgy Trust, London, 1967, pp. 154-170.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
254 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
[5] Spera, D. A., "A Linear Creep Damage Theory for Thermal Fatigue of Materials,"
Ph.D. thesis. University of Wisconsin, Madison, 1968.
[6] Coffin, L. F., Proceedings, 2nd International Conference on Fracture, Chapman and
Hall, London, 1969, pp. 643-654.
[7] Ellis, J. R. and Esztergar, E. P. in Symposium on Design for Elevated Temperature
Analysis, American Society of Mechanical Engineers, 1971, pp. 29-43.
[8\ Manson, S. S., Halford, G. R., and Hirschberg, M. H. in Symposium on Design for
Elevated Temperature Analysis, American Society of Mechanical Engineers, 1971, pp.
12-23.
[P] Manson, S. S. in Fatigue at Elevated Temperatures, ASTM STP 520, American Society
for Testing and Materials, 1973, pp. 744-782.
[10] Coffin, L. F., Jr., Transactions of the American Society of Mechanical Engineers,
Vol. 79, No. 7, Oct. 1957, pp. 1637-1649.
[11] Morrow, J. and Halford, G. R., Proceedings, Joint International Conference on Creep,
Institution of Mechanical Engineers, 1963, pp. 3-43 to 3-47.
[12] Swindeman, R. W., Proceedings, Joint International Conference on Creep, Institution of
Mechanical Engineers, London, 1963, pp. 3-71 to 3-76.
[13] Manson, S. S., Halford, G. R., and Spera, D. A. in Advances in Creep Design, A. 1.
Smith and A. M. Nicolson, Eds., Applied Science Publications Ltd., London, 1971, pp.
229-249.
[14] Halford, G. R., Metallurgical Transactions. Vol. 3, No. 8, 1972, pp. 2247-2256.
[15] Manson, S. S., Halford, G. R., and Nachtigall, A. J. in Advances in Design for Ele-
vated Temperature Environment, American Society of Mechanical Engineers, 1975, pp.
17-28.
[16] Code Case 1592, Boiler and Pressure Vessel Code, American Society of Mechanical
Engineers, April 1974.
[17] Halford, G. R., Hirschberg, H. M., and Manson, S. S. in Fatigue at Elevated Tempera-
lures, ASTM STP 520, American Society for Testing and Materials, 1973, pp. 658-669.
[75] Manson, S. S., Thermal Stress and Low-Cycle Fatigue, McGraw-Hill, New York, 1966.
[19] Spera, D. A. and Cox, E. C , this publication, pp. 69-85.
[20] Hirschberg, M. H. in Manual on Low Cycle Fatigue Testing, ASTM STP 465, American
Society for Testing and Materials, 1969, pp. 67-86.
[21] Mowbray, D. and McConnelee, J., this publication, pp. 10-29.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
STP612~EB/Nov.1976
Summary
The objective put forward by the committee which organized the Sym-
posium on Thermal Fatigue was to exchange ideas on methods for deter-
mining and improving the thermal fatigue resistance of both materials and
components. A total of 16 technical papers were presented at the sym-
posium, of which one was a review of terminology and historical develop-
ments in the field of thermal fatigue, thirteen were devoted to research on
materials, and only two dealt directly with components. Because of the
large quantity and variety of information assembled by those who con-
tributed to this symposium, it is appropriate to summarize here the major
conclusions offered, to evalute the success of the symposium in achieving its
goals, and to offer some suggestions for future thermal fatigue research.
A nalysis Papers
Mowbray and McConnelee compared the results of two-dimensional
finite-element analysis of temperature, stress, and strain (elastic and
inelastic) in a thermal-stress fatigue specimen with similar results using a
simpler uniaxial, elastic strain invariance (UESI) method and found that
stress-strain behavior trends were the same. Moreover, in most temperature
regions of the specimen, identical behavior was calculated by the two
methods. Therefore, it was concluded that the simpler and less costly UESI
method was adequate for further parametric studies of three damage
calculation methods.
255
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
SUMMARY 257
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
258 THERMAL FATIGUE OF MATERIALS AND COMPONENTS
parameter is equal to the product of the modulus in the load direction and
the strain intensity range.
Sheffler demonstrated the drastic reductions in Ufe which thermal-
mechanical cycling can produce in two iron-base alloys in comparison to
isothermal cycling at the maximum cycle temperature. In most cases, the
Hfe reductions are attributed to grain-boundary creep damage. A few
specimens of a high ductility alloy subjected to out-of-phase cycling showed
pronounced geometry changes which led to premature failure. This same
behavior has been observed previously in two very ductile tantalum alloys,
suggesting that high-ductility alloys may be more susceptible to geometric
instability.
Halford and Manson demonstrated the feasibiUty of extending the
techniques of strainrange partitioning to include the calculation of thermal-
mechanical fatigue life. It was proposed that further development of this
analysis will remove present restrictions such as requirements for
measurable inelastic strain, temperature insensitivity of life relations, and
metallurgical stabiUty.
According to Coffin, geometric instabilities such as those observed
during thermal-mechanical strain cycUng must be preceded by a gradient in
the cyclic stress-strain field. This gradient may be caused by geometry, tem-
perature, or material variations. It was also suggested that macroscopic
geometry changes occurring during thermal-mechanical fatigue may be
analogous to microscopic surface roughening and crazing observed during
thermal-stress and isothermd fatigue.
Critical Experiment
An important experiment which remains to be performed is the direct
comparison between the life of a thermal-stress fatigue specimen and that of
a thermal-mechanical fatigue specimen subjected to identical cycles of
strain and temperature. While it is convenient to assume that these two lives
are roughly equal, correlation between thermal-stress fatigue and thermal-
mechanical fatigue has not been critically evaluated in the literature.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions auth
SUMMARY 259
However, all the necessary tools are now available, as evidenced by papers
presented at this symposium. These tools are (a) equipment for well-
controlled thermal-stress fatigue tests, (b) analyses for determining cycles of
temperature and stress at crack locations in a thermal-stress fatigue
specimen, and (c) thermal-mechanical fatigue machines for faithfully
reproducing these cycles in a material test specimen. This critical experiment
is important, not only for the technical information which it would
produce, but also for the unifying effect it would have on research and
development in the field of thermal fatigue.
D. A. Spera
Fatigue Branch, National Aeronautics and
Space Administration Lewis Research
Center, Cleveland, Ohio; symposium co-
chairman and also coeditor of this publi-
cation.
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.
STP612-EB/NOV.1976
Index
A Thermal stress, 3, 10, 30, 38, 55,
. , . , , , . ^- . . 69,86,106,123
Analysis (calculation, prediction) ci j - j u j ir> o^ in/: m
r> \ • 1 10 Ir. iin Fluidized bed, 10, 86,106,123
Creep strain, 1, 38, 69, 239 Fracture 55 199
Fatigue life, 38, 55, 69, 239 i^racture, 55, 199
Finite element, 11, 32, 40, 71, 93,
155 H
Plastic strain, 1,38, 69, 239 Heat exchangers, 141
Stress, 10, 30, 69 Heat resistant alloys, 106
Thermal (temperature), 10, 59, 88
I
Intergranular fracture, 239
Ceramics, 55 Iron alloys, 86, 141, 170, 215, 229,
Coatings, 70, 106, 143 239
Cobah alloys, 12, 106,123
Computer program, 11, 32, 42, 62,
69, 166 L
Crack initiation, 69, 157 Life prediction, 38, 55, 69, 239
Crack propagation, 55, 86, 157, 199
Creep strain, 38, 69, 239
M
Materials
D Ceramic
Definitions, 3 Silicon nitride HS-130, 57
Directional solidification, 107, 199 Soda-lime-silica glass R-6, 56
Disk Cobalt alloys, 106
Test specimens, 10, 87 FSX414, 12
Turbine, 31, 38 Mar-M509, 123
Copper, zirconium, 228
P Iron alloys
A286,215
Fatigue AISI 1010, 170
Thermal mechanical, 3, 69, 141, H-13, 86
157, 170, 199, 214, 227,239 Low-carbon mild steel, 141
261
Copyright by ASTM Int'l (all rights reserved); Sun Dec 27 13:13:08 EST 2015
Downloaded/printed by
University of Washington (University of Washington) pursuant to License Agreement. No further reproductions authorized.