Sunteți pe pagina 1din 11

Computers & Structures Vol. 16. No. 1--4,pp.

531-541, 1983 Pnnted in Great Britain

0045=7949/83/010531-11503.0010 Pergamon Press Ltd.

CUMULATIVE DAMAGE IN STEEL STRUCTURES SUBJECTED TO EARTHQUAKE GROUND MOTIONS


HELMUTKRAWlNKLERand MAHMUD ZOHREI Department of Civil Engineering, Stantord University, Stanford, CA 94305, U.S.A.

Abstract--Experimental data are presented from low cycle fatigue tests of structural steel components. In these tests the failure modes of local buckling in beam flanges and fracture at weldments were studied in detail. Cumulative damage models are proposed which permit a life prediction for arbitrary cyclic loading histories. For the local buckling failure mode a series of linear damage models is used to predict deterioration threshold and deterioration, whereas a single model is used for the fracture mode. Crack propagation at weldments is modeled with a crack growth rate model based on the plastic strain range. Adequate predictions of lives were obtained from the analytical models. the deterioration threshold (number of cycles without noticeable deterioration) is small and deterioration occurs at a low rate, indicating that some deterioration should be accepted in the failure definition. This type of behavior is characteristic for local buckling modes occurring in cyclically loaded flexural elements of steel structures. The behavior shown in Fig. l(b) is characteristic for crack propagation and fracture problems at weldments. Localized crack growth will not cause noticeable deterioration of the component strength and stiffness until the crack has grown considerably. Once Pf = P[D > 3'] = P C (AS,~)c > y (1) deterioration takes place, it occus at a very high rate leading soon to failure. Thus, it appears to be justified to ignore deterioration in this case and to associate failure where D = accumulated damage; y = limit value of acwith the onset of noticeable deterioration. ceptable damage; n = number of damaging reversals or The objective of the work discussed in this paper is to cycles; ASp~= plastic deformation range of reversal or identify the characteristics of localized failure modes cycle i; and 6', c = structural damage parameters. (local buckling and crack propagation in components of Ideally, a value of y equal to one should constitute steel structures) and to develop simplified cumulative failure. Considering the variability in material properties damage models which permit the assessment of comand workmanship, 3' as well as at least one of the structural damage parameters (C, c) should be taken as ponent reliability under arbitrary cyclic loading of the type experienced in severe earthquakes. To fulfill this objecrandom variables. In seismic response evaluations, the tive, two series of specimens were tested, using ten inelastic deformation ranges are usually of considerable identical specimens for each series. The specimens were magnitude which justifies the omission of elastic deforsubjected to monotonic loading, cyclic loading with conmations in damage evaluation. The parameters n and stant deflection amplitudes and cyclic loading with variLxSp~are also random variables due to the randomness of able deflection amplitudes. The constant amplitude test the seismic input. were used to develop cumulative damage models which Equation (1) is based on several simplifying assumptions. It is assumed that for constant amplitude cycling were then employed to predict deterioration and failure the relationship between the number of cycles to failure, for the test specimens subject to variable amplitude N~,, and the plastic deformation range is of the Coffin- loading. Manson type, i.e. Nf=C-t(L~8,) -~. As a further assumption, Miner's rule of linear damage accumulation is employed, that is, the damage in cycle i is taken as I/NI~ and the accumulated damage is Y, IlNt~. In addition, mean deformation effects are ignored in the formulation presented in eqn (1). It remains to be answered whether the simple cumulative damage model included in eqn (1) gives realistic results and whether a prediction of failure can be achieved with sufficient accuracy. The experimental work discussed in this paper indicates that deterioration is usually of one of the two types shown in Fig. 1. It can be seen from this figure that ! deterioration is not a linear process and that a definition N U ~ E R 0~- CYCLC.~ NUM.B~Ie O F CYCLES of failure should be associated with an acceptable level Fig. 1. Modes of deterioration. of deterioration or damage. In the case shown in Fig. l(a) 531 I. I~"rRODUCTION In order to assess the reliability of structures subjected to severe ground motions, it is necessary to evaluate failure modes which lead to cyclic deterioration in strength, stiffness and energy dissipation capacity. It is convenient to use cumulative damage models to predict the probability of failure in cyclically loaded materials or structural elements. In the simplest case, the probability of failure, P~, can be expressed as

532

H. KRAWlNKLER "~. ZObiREI and

2. ~ SPlIClMIgNS gXIPEiIIMEN'I'AL AND MEASUREMENTS Cantilever spcimens were used for both test series. A typical test setup is shown in Fig. 2. The specimens consisted of hot rolled structural wide flange shapes (ASTM A36 steel) which were welded to a column stub (BI specimens) or to a base plate (B2 specimens). Details of the connections and the instrumentation are shown in Fig. 3. The BI specimens (Fig. 3a) were designed so that crack propagation at the weldments was the predominant mode of deterioration. A W 4 x 13 section with a small b/t ratio ( b i t = 11.5) was selected to prevent, as much as possible, the formation of local buckles. Strain gages were applied close to the welds to record nominal strain measurements. Several techniques were employed to measure the crack dimensions during and after the tests. The length of the cracks and the crack mouth opening displacement, as well as the crack depth for shallow cracks (up to about 0.5 mm depth), were measured from impressions made with a silicon-base precision impression material [1]. The depth of deep cracks was measured aplaeXimately by inserting a 0.001 in. thick copper plate into the cracks. Accurate measurements of

Fig. 2. Experimental setup.

#. 2 4

~OS" ~#3S/ "

i_ _____~

~J
'i

i
I

5T

....

(~) 5 P E C / M E N

.27I

$92 " t 02//"

3..9,$"

t~
e I

1
[ :!

\
I !

(6)

SPEC/MEN

22

Fig. 3. Details of test specimens.

Cumulative damage in steel structures subjected to earthquake ground motions crack geometry, for deep cracks, were obtained from a fractographic study of the fracture suface. The B2 specimens (Fig. 3b) were designed so that local buckling in the beam flanges was the predominant mode of deterioration. A W 6 9 section with a b/t ratio of 18.9 was selected for this purpose. Strain gages were applied to both sides of the flanges in order to detect the onset of local buckling. Rotations, in the plastic hinge regions were deduced from pairs of displacement gages attached as shown in Fig. 3(b). All specimens were subjected to monotonic or cyclic loads at the tip of the cantilever. Tip deflections were used to control the loading histories. In the process of data reduction, the plastic deflections were converted into plastic hinge rotations by using the point hinge concept, i.e. 0p = 6pll. This conversions was done because for flexural elements the plastic rotation range is a more

533

relevant global response parameter. However, the conversion is only approximate because it disregards the effect of moment gradient on plastic hinge formation and localized failure modes.
3. DAMAGEDUE TO LOCAL BUCKLING

3.1 Experimental results Local buckling of the type shown in Fig. 4 was the cause of deterioration in the B2 specimens. Buckling occurred either during the first excursion or, for small deflection ampitudes, after a small number of reversals. In the constant amplitude tests, the buckle size (maximum deflection of the flange due to buckling) increased almost linearly for several cycles and then the rate of growth decreased steadily until a stage of stabilization was reached. This can be seen from Fig. 5 which shows

Fig. 4. Local bucklingfailure model

/.0
V

V
0

V 0 0

0.5

o v
I

~Po.r/Pivc P E A K LO.4D3 NE6AT/VE P E A K LOADS


i I

/o

,~0 NUM2~ER

30 o F IPEY',~'~AL$

4,.0

Fig. 5. Buckle size--specimen B2--6.

334

!~. KR.~,WlNKLI~P. M. ZOI4REI and


RANGI

the growth of the buckles in both flanges versus ;he number of reversals for specimen B2-6. The effect of this local buckling on the load-deflection response is a continuous decrease in strength, stiffness and hysteresis loop area as shown in Fig. 6. An evaluation of the deterioration of these response parameters is discussed next. In order to assess deterioration, an accurate prediction of the undamaged response for each cycle was needed. This was achieved by using experimental data for the monotonic and cyclic stress-strain relationships of the base material [2] and the heat affected zone [3] and by employing a cyclic hardening function of the type

R A N G E .~

RANG.4Z~"

II
/-O

f
O.8 ~V~Oo,,

I
c~

aoo

P,,(n) = P,(I)[1 + H(1 -e-S~"~")]

(2)

in which P,(n) is the nondeteriorated (undamaged) peak load at reversal n, H is the amplitude dependent hardening (difference between monotonic and cyclic stabilized peak load) and B is a hardening factor which was taken as 0.52. Hysterctic energy dissipation (areas enclosed by the hysteresis loops) was evaluated by fitting a RambergOsgood function to the predicted response. For each constant amplitude test the predicted undamaged response was compared to the experimental response by plotting ratios of peak loads (PIP.), elastic stiffncsses (KIK.) and hysteretic energies (E/E,) for each reversal. A typical result (PtP, for specimen B2--6) is shown in Fig. 7. All plots showed the same pattern in which four ranges can be identified, a short range of deterioration threshold in which flange buckling has not occurred or is insignificant, and three deterioration ranges (Ranges I-III in Fig. 7). In range I deterioration proceeds at a high rate which is associated with the continuous growth of the flange buckles (Fig. 5). In range II deterioration proceeds at a slow and almost constant rate due to the stabilization in buckle size. These two ranges are followed by a range of rapid deterioration which is caused solely by crack propagation at the welds. Although small surface cracks formed early in the his-

0.21

IVEI~4TIV~EAK I I

LOADS I I 40 i

to 20 /vUI.~.BC~

5O

OF A~EV~'~SA~S

Fig. 7. Strength detm'ioration--specimen B2--6. tory, those cracks had no noticeable effect on strength and s ~ s s until tbey grew ~ the thickness of the ~ g e and ~ io~y. With a r e ~ degree of accuracy the deterioration ranges I and II can be represented by two straight lines as shown in the e ~ in Fill. 7, When the slopes of these fines are plotted on a ~ paper against the plastic rotation ~ , u s i q all ~ t a n t amplitude tests, the results shown in Fig. 8(a) m ol~ined. This graph represents the rate of s t m ~ deterioration for ranges I and II. Similarly, rates of sti~ne, deterioration and of the detericcation in hysteretic energy can be plotted as shown in Figs. 8(b, c). In all three cases straight regres-

E N D t. 0 D

12.1IN L

8~I

4 N

-4.

-8

J,~ W

-I.N

e.~
m E P ~

l.m
, IN.

a.~

Fig. 6. Load-deflectionresponse--specimen B2-6.

Cumulative damage in steel structures subjected to earthquake ground motions


0.1 ,q .... /

535

Q(

///+
.4+
j/ ~ z _ _. _ ,I.69~

0.01 0.00/

I 11 I I 001

1 ~ ) ~ o.I

..,eTRENGT/'f

Z)'7"~RIORAT"ION PER ~EVCRSAL~ ,~o//.,/

~)
q

//
/ /

,,d~ = o . , , , [ A , ) ~
z
-

./.e/s

///7" S ///
%

O01

I I O.O00S

I I I 0.00I

I I I 1 0.01

O.OS

3T/~'FNE~',,

.DETERIORATION

PE,i~

REVR,.~AL , A o t g

O.I ~-., ,q

//'I

S~ (3

--7-

(--.)% -"
//// + ~

P< ~o

D.

O.OJ /

I" 1 / I

.00/

I I I O.OI

f
0.1

I I

(~)
Fig. 8. Rates of deterioration. sion lines can be fitted rather accurately to the data points of range I and, with much less confidence, to the data points of range II. The lines so obtained represent deterioration models which can be used for damage prediction. 3.2. The deterioration pattern of all constant amplitude tests of this test series is of the type shown in Fig. lta). Thus, a single cumulative damage model cannot be used for the full range of interest. It appears to be feasible. however, to use a series of damage models which describe individually the different ranges. The deterioration threshold range in the test series was unfortunately too small to permit the development of a rational damage model. Since this range did not have much effect on total damage, a simple Coffin-Manson model was assumed together with linear damage accumulation. In deterioration range I the rate of deterioration per reversal, Ad, for constant amplitude cycling can be expressed accurately by a function of the form Ad = A(ASp) (3)

Damagemodels/orlocalbuckling

where A and a are parameters which depend on the properties of the structural component, and A5, is the

536

~ KraWlnKLer and M. ZOHRE!

plastic deformation range which in this case is taken as A0p. A function of this type applies to deterioration in strength and also to deterioration in stiffness and hysteretic energy (Fig. 8). If the hypothesis of linear damage accumulation is accepted for reversals with variable amplitudes, the accumulated deterioration can be expressed as

d = ~ Ad, = A ~ (A~p~)~
i=l i~l

(4)

where n is the number of reversals. Let us assume that the deterioration model given by eqn (3) applies for a range of deterioration from zero to x (say, x = 0.1, i.e. 10% deterioration) and that we define failure as attainment of this deterioration. Then the number of reversals to failure for constant amplitude cycling is of the Coflin-Manson type, i.e.
Nf = xA-'(A~p) -a = c - I ( A ~ p ) - c

/vvv j,
2O

,,,A ^A/

a/

Fig. 9. Variableamplitude history. and a range of almost linear deterioration extending over more than two blocks. This gives some credence to the hypothesis of linear damage accumulation for range L Close to the end of block three the rate of deterioration decreases, indicating that a considerable portion of this block is spent in deterioration range II. The predicted deterioration is somewhat larger than the experimental one, but only by a small amount when the crossover into range II is considered. For practical purposes, the prediction based on range I deterioration should be adequate. In the damage models presented here, the effects of mean deformations have been neglected. A pilot test of a specimen subjected to constant amplitude cycling with a mean deflection equal to approximately half the deflection range has shown that the rate of deterioration in the first reversal is relatively high, but for subsequent reversals the rate is close to that obtained from cycling with zero mean. Thus, it apl~ars that for the local buckling deterioration mode the mean deformation is not of foremost importance.
4. DAMAGg DUg TO CRACK PROPAGATION

(5)

and with linear damage accumulation the total damage can be expressed as

o;

l/N,,-- c

(6)

where C = Alx and c = a. This formulation of cumulative damage is identical to that used in eqn (1). A value of D equal to one would constitute attainment of deterioration x which was defined as failure. Thus, the formulation given in eqn (1) can be applied for a range of linear deterioration. Contrary to customary practice, the coefficients A and C are based here on N and n being the number of reversals rather than cycles. This was done because the deterioration was approximately equal for the two reversals of each cycle. If cycles are used rather than reversals, A and C should be replaced by 2A and 2C. The models formulated for deterioration range I can be applied also to range II although with larger uncertainties. Thus, at any stage of cyclic loading, the amount of deterioration can be predicted by first exhausting the damage model used for deterioration threshold (D, = 1) and then using the cumulative deterioration relationship given in eqn (4). A cross-over from range I to range II can be accomplished in this process but is cumbersome because the deterioration at which crossover occurs depends on the deformation amplitude. The use of only the first deterioration range is always conservative and should give satisfactory results. In order to test the proposed model, one test specimen was subjected to variable amplitude loading and the deterioration obtained in the experiment was compared with the prediction. The cyclic deflection history for the variable amplitude test, shown in Fig. 9, represents the response of a single degree of freedom system subjected to a realistic earthquake ground motion. The response was scaled so that the maximum deflection range (from point 11 to point 20 in Fig. 9) corresponds to a plastic rotation range of A0p =O.038rad. This history was applied repeatedly to the test specimen. [n the prediction process the rain-flow cycle counting method [4] was used to order the plastic deformation ranges. The comparison of predicted and experimentally obtained strength deterioration is shown in Fig. 10. The experimental data points show a deterioration threshold

4.1 Experimental results Crack propagation at the beam flange welds, as shown in Fill. 11, was the cause of damage and fracture in the BI specimens. The behavior of the specimens was consistent in some respects and inconsistent in others.

20

,$ - - - -

! XPWR/M~'~/7"AL ~,PED~7".e'.O, RANGE .T / f/

~.

//o

i '
0 / 2. ,3

IVUM.~#R O F

8t O C K 5

Fig. 10. Damage in variable amplitude test--localbuckling mode.

Cumulative damage in steel structures subjected to earthquake ground motions

537

Fig. 11. Crack at beam flange weld.

Differences in workmanship were responsible for large variations in notch geometries at the weld toe and, correspondingly, for large variations in initial crack sizes. In most of the specimens, the predominant crack propagation occurred at the centerline of the flange (below strain gage 3 in Fig. 3a) with a single crack growing through the flange thickness. In a few specimens, however, edge cracks which propagated along the flange were equally or more important. But also in these specimens the centerline cracks grew to considerable sizes to that the specimens would have fractured at the center line soon after the actual fracture initiated by edge cracks. Thus, the effects of the edge cracks are ignored in this discussion. Consistently, cracks propagated from surface imperfections (notches) at the weld toe and in none of the

specimens internal imperfections in the welds were of importance. Small surface cracks were observed very early in the history, in some cases already during the first excursion. After a few reversals small surface cracks joined, forming a relatively long but shallow crack. This crack propagated through the heat affected zone at an increasing rate until it either grew through the flange thickness or joined with a smaller crack which initiated at the opposite side of the flange at the coping. The through thickness growth did not lead to a noticeable deterioration in the strength of the specimen, but once a through crack was formed, the crack propagated rapidly along the flange leading to severe deterioration in resistance. This is evident from Fig. 12 which shows the rapid deterioration in resistance in the first cycle after the through crack has formed.

SI=EC E N D L 0 A D 1 2 O0

]11-3

$ OO

4 00

~e

- 4 Oe

- 8 00

- 4 ~IQ

END l~'kl[CrZO~ . ZN Fig. 12. Load-deflectionresponse--specimen BI-3.

338

~]. KRAWlNKLERand M. ZOHREI

Much futile effort was invested in attempts to identify initial crack sizes and early crack growth through measurements using the previously mentioned siliconebase impression material. In fact, it became evident that the measurement of initial crack size and early crack growth at weldments is an imposssible task because of the irregular geomtry of imperfections and the random joining of surface cracks. Thus, it was decided to deduce an equivalent initial crack size from crack measurements at a more advanced stage of cracking. Solomon [5] has shown.that the log of the crack depth, a, is linearly related to the number of cycles provided that all cycles are of equal strain amplitude close to the crack. His tests indicate that this linear relationship holds true for the full range of crack propagation, from very small cracks to cracks approaching unstable crack growth. In the component tests performed in our study, tip deflections and not strains were controlled in the constant amplitude tests. However, the nominal strain recorded close to the crack plane (Fig. 3a) did stabilize to a constant amplitude after a small number of reversals. Thus, it can be justified to consider the constant deflection amplitude tests as constant strain amplitude tests. For all constant amplitude tests the crack depth, a, was potted against the number of cycles, N, on a loglinear paper, using primarily the data for large crack sizes which could be measured more accurately. The most reliable data were obtained from photos of the magnified fracture surface which showed clean striation lines for large cracks. A typical fracture surface, with a scale in centimeters, is shown in Fig. 13. The plots of crack depth versus N confirmed the hypothesis that log a and N are linearly related as can be seen from the example presented in Fig. 14. These plots were used for two purposes. The straight line placed through the data points for each test was extended backwards to N equal zero, resulting in a predicted initial crack size ao. For the six constant amplitude tests performed in this study, the mean value of ao was 0.0016 in. and the standard devia-

/" r
o./: ~

/
/

:
Q.

/
/

"3.01

/
~00~

i
20

i, i 4tO

l
60

NUMJE~

O~ C/CLE,~

Fig. 14. Crack propagation--specimen BI-8. tion was 0.0012 in. The large scatter comes as no surprise because of the large differences in initial imperfections in the specimens. The values so obtained can be used in lieu of a measured ao for crack propagation studies. In order to trace crack growth, a crack propagation model was developed using the plastic strain range as the basic deformation parameter. For this purpose, the slopes of the lines of the a - N graphs were plotted against the plastic strain ranges, Aep, using the natural tog for crack depth and a log-log paper. The results are shown in Fig. 15. With good accuracy a straight line can be placed through the data points, indicating that the resulting crack propagation model can be applied with confidence. The crack propagation model so obtained is of the form d-~ = aa(A~J.
do

(7)

Fig. 13. Fracture surface--specimen Bt--6.

In this equation, the parameters a and # depend on the material properties, the geometry of the component, the shape of the crack, and on the location at which stratus are measured. In our study, strains were measured at a distance of 3/8 in. away from the crack plane, resulting in a nominal reference strain. If the strains would have been measured closer to the crack plane, larger strains would have been obtained and smaller crack growth rates (as a function of A,p) would have been predicted. This explains, in part, the difference between the value of the coefficient a obtained from this study (a = 130) and that obtained by Solomon [5] (a = 19.2). Since the coefficient # is almost identical in both studies (1.91 vs 1.86), the crack propagation model obtained in this study would be equal to that obtained by Solomon if the plastic strain ranges are multiplied by a factor of approx. 2.75.

Cumulative damage in steel structures subjected to earthquake ground motions /:o

539

"b
0.I

-/30

(,,~,) ,9, I
I I I I
II 1

Life predictions for the specimens can be obtained by integrating eqn (7) between the initial crack size ao and a eritica! crack size a~. For the specimens tested in this series, the critical crack size was in the order of half the flange thickness. At about this stage, unstable crack growth occurred through the thickness of the flange. An exact value for ac is not needed because variations in a~ have little effect on life predcitions. An alternative interpretation of the test data is obtained by plotting the number of cycles to failure against a representative plastic deformation range, using all constant amplitude tests. These plots are shown in Fig. 16, using AG and A0p as plastic deformation ranges. One can place straight regression lines through the data points, but a considerable scatter has to be accepted. This scatter is most evident at ~x0p equal to 0.12rad., where three tests with identical plastic deformation ranges (identical displacement histories) led to lives of 8, 14 and 15 cycleso However, when the mean lives are used for the tests with identical rotation ranges, then a regression line fits well to the three data points. This indicates that a Coflin-Manson relationship can be used to predict lives for constant amplitude cycling, but the previously discussed differences in initial imperfections will lead to large uncertainties in life predictions. 4.2. Damage Models [or Crack Propagation at Weldments

~oJ ~.oI

0.1

I~LAST"IC ..,eT"~hlH ;PAHGE ~Ep

Fig. 15. Crack propagation model.

The damage pattern of the specimens of this test series is of the type shown in Fig. l(b). In the overall response, deterioration is not noticeable until the life is almost

to~ 0"1

~2

O.Ol

1o N U M , BLmR

too

O~r CYCLE5

7"0 T'AILU1PEs N~

o)

L~

0./

tb

v~

~. o.o~
io /oo

IVUI.'I.BE'~ 0~" CYCLES

";'G .TAILL/~Es

IV~

Fig. 16. Number of cycles to failure for constant amplitude tests--Bl Specimens.

540

}t. KaAwt?v~zL~eand M. Z(~HR~i

exhausted. Thus, the deterioration phase usually can be neglected and a deterioration threshold model will suffice in most cases. Such a model can be based on a conventional low cycle fatigue approach or on concepts of plastic fracture mechanics. Low cycle fatigue approach. It appears to be feasible to use a function of the type given in eqn (5) to predict the number of cycles to failure for constant amplitude cycling (see Fig. 16). However, the coefficient C may be a random variable with a considerable scatter. It becomes then a matter of philosphy whether the hypothesis of linear damage accumulation can be accepted for variable amplitude cycling. If damage is related directly to crack size, linear damage accumulation breaks down because the rate of crack growth increases rapidly with crack size. However, there is no evident reason why this relationship should be made. As long as the purpose of a cumulative damage law is to predict likelihood of failure. a crack size below a~ is not at all related linearly to damage. Thus, the authors see no evident reason why linear damage accumulation is less acceptable for crack propagation problems than for other low cycle fatigue problems. It appears that the acceptance of linear damage accumulation for variable amplitude cycling is equivalent to the acceptance of a single da/dN law for crack propagation. . Provided that this reasoning is sound, a damage model of the type expressed by eqn (6) should be acceptable. Since crack propagation is governed by local strains, the plastic deformation range in this equation should be the plastic strain range Ae~, close to the crack. The use of a global response parameter, such as A0~,, involves further inaccmacies unless A~, and A0z, are related linearly. The use of the rain-flow cycle counting method is recommended in the low cycle fatigue approach. Fracture mechanics approach. Provided that a crack propagation model of the type shown in eqn (7) can be found, the number of cycles to failure can be obtained by integrating this equation between a,, and a~.. In eqn (7) the coefficient a should be taken as a random variable to account for the uncertainties in material properties and the uncertainties in the evaluation of the plastic strain

range Ae,. As was mentioned previously, the Iact propagation model depends on weld geometry, crack shape and location of strain measurement. The model could be improved, probably, by using the 2,J integral rather than Ae,, as the deformation parameter. The use of AJ would make the model more general, because differences in specimen geometry and crack shapes can be accounted for in the formulation of Z This approach was not pursued in detail because of difficulties in evaluating M in the presence ,,)f crack closure. Most of the large scatter in lives of the test specimens must be attributed to differences in initial imperfections. Thus, the most critical parameter in life prediction is the equivalent initial crack size ao, which is a random variable with large scatter. Much more work needs to be done to obtain statistically acceptable data for ao for different weld sizes and geometries. The approach used in this study, that is, extension of the a - N line ',o N = 0, appears to be appropriate for this purpose. The cumulative damage model as well as the crack propagation model were used to predict the behavior of a specimen with variable amplitude loading. The deflection history used in this test is that shown in Fig. 9, scaled to a maximum plastic rotation range (points 11-20) of 0.125 rad. (A~, = 0.05l). Three experimentally obtained crack depths as well as the predictions are shown in Fig. 17 If the crack propagation model is used together with the mean of ao, the life is underestimated by a significant amount. The use of mean minus standard deviation for initial crack size gives a good prediction of crack growth and life to fracture. The use of the cumulative damage model, using the regression line shown in Fig. 16(a), results in a life which is close to that predicted by the crack propagation model with the mean of a,,. Thus. the models appear to give conservative but reaiistic life predictions and the predictions based on the two models are in close agreement when mean values are used. Mean strain effects were neglected in the prediction of crack propagation and life. A pilot test of a specimen subjected to constant amplitude cycling with a tensile mean strain equal to approximately half the strain range

a/ac , EZ P1~ IM1v'rA I.,


. . . .

o/oc, Oo. O.oo#~r" ( M c A a ) ~ / = = , a. = =ooo4" ( ~ r A ~ - ~ ' J


C UMULA FIVE ,,~AMACK MO~E~'L

,3

4 0~" ~ Z O C K 3

5"

NUMBER

Fig. 17, Damagein variable amplitude test--crack propagation mode.

Cumulative damage in steel structures subjected to earthquake ground motions led to a life which was larger than that for the test with zero mean strain. Thus, it appears that a tensile mean strain has no detrimental effect as long as the strain range is sufficiently large to cause crack closure. 5. CONCLUSION The tests discussed in this paper have shown that standard low cycle fatigue damage models can be used to predict the life of structural components subjected to variable amplitude cycling with large inelastic deformation amplitudes. In the case of local buckling modes, two separate damage models can be employed to account first for deterioration threshold and then for deterioration. In crack propagation and fracture modes, the deterioration threshold life is large compared to the deterioration life, and the latter can often be neglected. Thus, a single damage model wilt suffice in most cases. In life predictions for crack propagation and fracture modes, the plastic fracture mechanics approach is an attractive alternative to low cycle fatigue modeling. The advantage of this approach is that the use of a statistically acceptable initial crack size together with the use of the AJ integral could lead to a general mathematical formulation which can be applied to components with

541

different geometries and crack shapes. This will require, however, much more research on the evaluation of the initial crack size and the AJ integral. Acknowledgements--The research discussed in this paper was supported by the Nationa! Science Foundation through grant PFR-7902616. This suppor', and the assistance of the graduate student N. Cofie are gratefully acknowledged.
REFERENCES

1. Y. W. Cheng. I. H. McHenry and D. T. Read, Crack-opening displacement of surface cracks in pipeline steel plates. ASTM Syrup. Fracture Mechanics, Los Angeles (June 1981) 2. M. Zohrei, Cyclic stress-strain behavior of A36 steel. Engineer Thesis. Department of Civil Engineering, Stanford University (Dec. 1978). 3. Y. Higashida, Strain controlled fatigue behavior of weld metal and heabaffected bae metal in A36 and A514 steel welds. Ph. D. Thesis. Department of Metallurgical Engineering, University of Illinois, Urbana-Champaign,Illinois (1976). 4. H. Krawinkler, Selection of loading histories for seismic testing of structural components. Proc. SESA-JSME lni. Con/. on Experimental Mech, Hawaii (May 1982). 5. H. D. Solomon, Low cycle fatigue crack propagation in 1018 steel. J. Materials 7, 299--306(1972).

S-ar putea să vă placă și