Multiaxial Fatigue Life Prediction Methods for Notched Bars of Ti-6Al-4V ABSTRACT

Similar documents
Cyclic Event Identification and Fatigue Damage Assessment for Multiaxial Mission Loadings

Evaluation of HCF Multiaxial Fatigue Life Prediction Methodologies for Ti-6Al-4V

* Many components have multiaxial loads, and some of those have multiaxial loading in critical locations

Fatigue and Fracture

Multiaxial Fatigue. Professor Darrell F. Socie. Department of Mechanical Science and Engineering University of Illinois at Urbana-Champaign

A Critical Plane-energy Model for Multiaxial Fatigue Life Prediction. of Homogeneous and Heterogeneous Materials. Haoyang Wei

Predicting Fatigue Life with ANSYS Workbench

Definition of Stress and Strain Ranges for Multiaxial Fatigue Life Evaluation under Non-Proportional Loading

Life Prediction Under Multiaxial Fatigue

Volume 2 Fatigue Theory Reference Manual

This guide is made for non-experienced FEA users. It provides basic knowledge needed to start your fatigue calculations quickly.

Calculating and Displaying Fatigue Results

Numerical methods of multiaxial fatigue life prediction for elastomers under variable amplitude loadings

ME 2570 MECHANICS OF MATERIALS

Structures and Multiaxial Fatigue Analysis. Timothy Langlais

DEVELOPMENT OF TEST GUIDANCE FOR COMPACT TENSION FRACTURE TOUGHNESS SPECIMENS CONTAINING NOTCHES INSTEAD OF FATIGUE PRE-CRACKS

Prediction of Elastic-Plastic Behaviour of Structures at Notches

Chapter 7. Highlights:

Plasticity R. Chandramouli Associate Dean-Research SASTRA University, Thanjavur

Fatigue Algorithm Input

Fatigue Life Assessment of 30CrNiMo8HH Steel Under Variable Amplitude Loading

Critical Plane Approach to Multiaxial Variable Amplitude Fatigue Loading

Mechanical Engineering Ph.D. Preliminary Qualifying Examination Solid Mechanics February 25, 2002

Stress-Strain Behavior

The science of elasticity

A REVIEW ON FRETTING FATIGUE CRACK INITIATION CRITERIA

Stress Concentration. Professor Darrell F. Socie Darrell Socie, All Rights Reserved

6.4 A cylindrical specimen of a titanium alloy having an elastic modulus of 107 GPa ( psi) and

ARTICLE A-8000 STRESSES IN PERFORATED FLAT PLATES

MULTIAXIAL ASSESSMENT METHOD FOR FATIGUE CALCULATIONS IN COMPOSITE COMPONENTS

PREDICTION OF FATIGUE LIFE OF COLD FORGING TOOLS BY FE SIMULATION AND COMPARISON OF APPLICABILITY OF DIFFERENT DAMAGE MODELS

Module 5: Theories of Failure

Stresses Analysis of Petroleum Pipe Finite Element under Internal Pressure

Examination in Damage Mechanics and Life Analysis (TMHL61) LiTH Part 1

EMA 3702 Mechanics & Materials Science (Mechanics of Materials) Chapter 2 Stress & Strain - Axial Loading

NONLINEAR LOCAL BENDING RESPONSE AND BULGING FACTORS FOR LONGITUDINAL AND CIRCUMFERENTIAL CRACKS IN PRESSURIZED CYLINDRICAL SHELLS

NORMAL STRESS. The simplest form of stress is normal stress/direct stress, which is the stress perpendicular to the surface on which it acts.

Samantha Ramirez, MSE. Stress. The intensity of the internal force acting on a specific plane (area) passing through a point. F 2

Available online at Fatigue Fatigue life prediction for large threaded components. Ali Cetin*, Gunnar Härkegård

5. STRESS CONCENTRATIONS. and strains in shafts apply only to solid and hollow circular shafts while they are in the

Lecture 8. Stress Strain in Multi-dimension

APPLICATION OF DAMAGE MODEL FOR NUMERICAL DETERMINATION OF CARRYING CAPACITY OF LARGE ROLLING BEARINGS

A novel approach to predict the growth rate of short cracks under multiaxial loadings

TMHL TMHL (Del I, teori; 1 p.) SOLUTION I. II.. III. Fig. 1.1

Model of Fatigue Crack and Effect of Triaxiality

INTRODUCTION TO STRAIN

Investigation of Thermo-Mechanical Fatigue Characteristics for Cast Aluminum (AL319-T7)

IMECE CRACK TUNNELING: EFFECT OF STRESS CONSTRAINT

MECE 3321 MECHANICS OF SOLIDS CHAPTER 3

CHAPTER 6 MECHANICAL PROPERTIES OF METALS PROBLEM SOLUTIONS

Archetype-Blending Multiscale Continuum Method

2.1 Background of Piping Stresses

Appendix G Analytical Studies of Columns

Elastic-Plastic Fracture Mechanics. Professor S. Suresh

Siping Road 1239, , Shanghai, P.R. China

Reference material Reference books: Y.C. Fung, "Foundations of Solid Mechanics", Prentice Hall R. Hill, "The mathematical theory of plasticity",

MATERIALS FOR CIVIL AND CONSTRUCTION ENGINEERS

Fatigue calculations in ANSYS Workbench. Martin Eerme

Crack Tip Plastic Zone under Mode I Loading and the Non-singular T zz -stress

Transactions on Engineering Sciences vol 6, 1994 WIT Press, ISSN

Introduction to Engineering Materials ENGR2000. Dr. Coates

CHAPTER 7 FINITE ELEMENT ANALYSIS OF DEEP GROOVE BALL BEARING

Failure analysis and optimization design of a centrifuge rotor

7.6 Stress in symmetrical elastic beam transmitting both shear force and bending moment

MECHANICS OF MATERIALS. EQUATIONS AND THEOREMS

MIL-HDBK-5H 1 December 1998

Chapter 5 Torsion STRUCTURAL MECHANICS: CE203. Notes are based on Mechanics of Materials: by R. C. Hibbeler, 7th Edition, Pearson

FATIGUE ANALYSIS: THE SUPER-NEUBER TECHNIQUE FOR CORRECTION OF LINEAR ELASTIC FE RESULTS

Proposed Method for Evaluating Multiaxial Fatigue in ITER

STANDARD SAMPLE. Reduced section " Diameter. Diameter. 2" Gauge length. Radius

Stress Analysis Lecture 3 ME 276 Spring Dr./ Ahmed Mohamed Nagib Elmekawy

A PAPER ON DESIGN AND ANALYSIS OF PRESSURE VESSEL

Kliment Ohridski Blvd No 10, 1756 Sofia, Bulgaria; Fax:

Strength of Material. Shear Strain. Dr. Attaullah Shah

Comparison of Models for Rolling Bearing Dynamic Capacity and Life

A Model for Local Plasticity Effects on Fatigue Crack Growth

Failure from static loading

PSEUDO ELASTIC ANALYSIS OF MATERIAL NON-LINEAR PROBLEMS USING ELEMENT FREE GALERKIN METHOD

EMA 3702 Mechanics & Materials Science (Mechanics of Materials) Chapter 3 Torsion

Numerical Simulation of Fatigue Crack Growth: Cohesive Zone Models vs. XFEM

Nonlinear Analysis of Reinforced Concrete Shells Subjected to Impact Loads

NEW CONCEPT OF A LOCAL ELASTIC STRESS APPROACH FOR FATIGUE LIFE CALCULATION

Mechanical Properties of Materials

Unified Quiz M4 May 7, 2008 M - PORTION

Chapter 6: Plastic Theory

Structural Metals Lab 1.2. Torsion Testing of Structural Metals. Standards ASTM E143: Shear Modulus at Room Temperature

Low-Cycle Fatigue Crack Growth in Ti-6242 at Elevated Temperature Rebecka Brommesson 1,a, Magnus Hörnqvist,2,b, and Magnus Ekh 3,c

CHAPTER 3 THE EFFECTS OF FORCES ON MATERIALS

Non-linear and time-dependent material models in Mentat & MARC. Tutorial with Background and Exercises

MMJ1133 FATIGUE AND FRACTURE MECHANICS A - INTRODUCTION INTRODUCTION

EVAPRED A CODE FOR FATIGUE ANALYSIS OPTIMIZATION

Advanced Software for Integrated Probabilistic Damage Tolerance Analysis Including Residual Stress Effects

[7] Torsion. [7.1] Torsion. [7.2] Statically Indeterminate Torsion. [7] Torsion Page 1 of 21

Module-4. Mechanical Properties of Metals

Fig. 1. Different locus of failure and crack trajectories observed in mode I testing of adhesively bonded double cantilever beam (DCB) specimens.

Plastic Anisotropy: Relaxed Constraints, Theoretical Textures

JJMIE Jordan Journal of Mechanical and Industrial Engineering

Fatigue Life Analysis of an Oil Injector of a Diesel Engine

Solution: The strain in the bar is: ANS: E =6.37 GPa Poison s ration for the material is:

Analysis of local stress concentration at transitions

Transcription:

Multiaxial Fatigue Life Prediction Methods for Notched Bars of Ti-6Al-4V Dr. Alan Kallmeyer 1, Mr. Ahmo Krgo 1, and Dr. Peter Kurath 2 1 Department of Mechanical Engineering, North Dakota State University, Fargo, ND 2 Department of Mechanical Engineering, University of Illinois, Urbana, IL ABSTRACT Many critical engineering components are routinely subjected to cyclic multiaxial stress states, which may include non-proportional loading and multidimensional mean stresses. Initial efforts by the authors involved the identification of suitable multiaxial damage parameters for a Ti-6Al-4V titanium alloy utilizing smooth samples. As with uniaxial loading, the extension of the multiaxial data to a notched geometry poses challenges. Utilizing the multiaxial stress-strain state at the notch root, estimates of the fatigue life for multiaxially loaded notched members may be highly conservative, even for damage parameters that successfully modeled the smooth specimens under identical loading. By using an average of the multiaxial stress field over a small region near the notch in the calculation of the damage parameters, more realistic fatigue life estimates were obtained. INTRODUCTION Performance requirements in certain applications require that an adequate fatigue damage parameter be identified, rather than masking uncertainty in excessive factors of safety. Ideally, uniaxial data would be utilized to establish a baseline data set that could be extended to more complex loading with a multiaxial damage parameter. A previous investigation [1] scrutinized twenty-two existing fatigue damage parameters in light of approximately thirty multiaxial smooth bar tests which encompassed a variety of mean stress conditions and non-proportionality. Figure 1 highlights some of the non-proportional paths employed in the investigation. Both the level of loading and shape of the paths are intended to simplify portions of actual service events. The reader is referred to reference [1] for other details of the testing and specimen design. The parameters evaluated in reference [1] included both the scalar effective-stress type as well as the more recently introduced critical plane approaches. For the most part, any plastic deformation observed during the multiaxial testing was restricted to the first reversal, resulting in fatigue lives of 5 1 4 to 5 1 6 cycles. Only one effective-stress type parameter, the modified Manson-McKnight parameter [2, 3], showed promise for general multiaxial loading. This model uses an alternating stress range, Δσ eff, defined from the general form of effective stress with the range of each stress component, and an effective mean stress, σ mean eff, defined as the effective mean stress obtained using the mean stress for each component. Damage = 1 2 Δσ (w 1) eff σ w max, σ max = Δσ eff 2 ( ) ( ) + β 2 σ mean eff, β = σ 1 + σ 3 σ 1 σ 3 The sign and magnitude of the mean stress term in the model is modified by multiplying the effective mean by the term β/2. β is defined by Σσ 1 and Σσ 3, which are the sums of the first and third principal stresses, respectively, at the maximum and minimum points in the cycle. Then σ max is obtained by adding the effective stress amplitude to the modified mean stress term. If (1) 1

strains (rather than stresses) are known, a pseudo-stress range is defined that makes use of an effective strain range with an elastic Poisson s ratio multiplied by the linear modulus to obtain Δσ eff. Three critical plane models were identified [1] for extension to notched member analysis. They include the Findley [4] parameter, which modifies the maximum shear amplitude by the maximum stress normal to the orientation of maximum shear multiplied by an influence factor, k. While it has been shown that k may not be a constant, it is assumed to be for the life range under consideration. Damage = Δτ 2 + k σ max n The Fatemi-Socie-Kurath (FSK) [5, 6] parameter has a similar basic mechanistic interpretation, but is formulated in terms of the shear strain amplitude. Damage = Δγ 2 1 + k σ n (2) max ( ) (3) Finally, a modification of the Chu-Conle-Bonnen (CCB) [7] parameter showed promise. It may be interpreted as a modification of the Smith-Watson-Topper (SWT) parameter, in which a shear term (Δγ τ max ) has been added to the normal stress-strain term. This parameter can be considered to represent the strain energy acting on a plane, as it involves the multiplication of corresponding stress and strain components, if all stress components other than axial and shear stress are zero. This is true for axial-torsional loading of thin walled tubes or elastic solid round bars. For more generalized multiaxial loading, the interpretation must be that of a portion of the strain energy. In this study, a slight modification was made to the model by multiplying the second term (Δε σ max /2) by an empirical factor, k, to allow for an adjustment in the influence of this term. It was found that this modification significantly improved the accuracy of this model. Damage =Δγτ max + k Δε 2 σ max n The SWT parameter itself was also previously considered [1] as a critical plane parameter. However, comparison of this model with the smooth-bar multiaxial data showed it was highly non-conservative. The successful critical plane parameters were all based on alternating shear stress or strain terms, indicating that shear stresses/strains drive the formation of fatigue cracks in this material. Interpretation of the basic form of Equations (1-4) is debatable from a mechanistic viewpoint for non-proportional loading. In this research, the maximum damage orientation is defined as that which has the maximum value of the parameter. While for more complex loading, the identification of a cycle, and the ancillary terms associated with that event pose challenges, the current efforts consider the basic path (Fig. 1) to be the fatigue event or cycle. For more complex loading, the fact that the critical plane may differ from event to event has implications regarding cumulative damage, which will not be addressed in this investigation. Damage for the more statistically significant uniaxial data has been fit to the following base equation for each parameter. Damage = AN ( f ) b + C( N f ) d (5) (4) 2

It is emphasized that the coefficients and exponents are not the traditional strain-life constants. Uniaxial data at different load ratios were utilized to calculate the value of k or w for each parameter. The solid curves shown in any of the life figures are the fit of the uniaxial data, utilizing Eq. (5). The constants determined for each parameter are summarized in Table 1. Smooth specimen multiaxial results are summarized in Figures 2 through 5 for the aforementioned parameters. There are approximately eight additional data points not reported in reference [1], that are summarized in Table 2. Statistical comparison of the baseline multiaxial results is presented in Table 3 utilizing the following equation. Damageexperiment Damage predicted % Error = 1 (6) Damage experiment Using this method of comparison, the mean value of the percent error indicates the degree of offset of the predicted damage parameters for the multiaxial data, relative to the baseline uniaxial curve. The standard deviation provides a relative measure of the degree of scatter about the curve. An ideal model is one for which both the mean and standard deviation of the error are minimized. NOTCHED EXPERIMENTAL PROGRAM Five circumferentially notched round bars with a notch net section diameter of 12.5 mm, a notch depth of 3.25 mm, and a notch radius of 2.5 mm were tested using the load paths shown in Figure 6. Theoretical stress concentration factors of K σ = 1.85 and K τ = 1.35 are associated with this geometry [8]. Specimen preparation was identical to the uniaxial and multiaxial smooth specimens tested previously. Individual test data are summarized in Table 4. NOTCH ANALYSIS AND DISCUSSION Since strains could not be measured at the notch root, a finite element analysis was performed to determine the stress-strain state at the notch. ANSYS was utilized with a tetrahedral 1-node element. A sensitivity analysis was conducted to ensure convergence of the stress solutions, using meshes that varied from approximately 5, to over 17, elements for the entire specimen. In the final mesh, shown in Figure 7, the average element side length in the notch root was approximately.5 mm. The results of the elastic analyses indicated that one of the tests (test C in Table 4) experienced plastic deformation during the loading. Consequently, an elastic-plastic analysis was conducted in ANSYS for this case. In performing the elasticplastic analysis, a multilinear kinematic hardening rule was assumed. The Ramberg-Osgood relation was used to model the cyclic stress-strain curve, using the data obtained from uniaxial testing. The material properties used in this analysis are E = 116.3 GPa, ν =.349, K = 855 MPa, and n =.149. Figure 7 shows a typical von Mises contour stress plot for test A, while Figure 8 shows the stress components as a function of radial position through the notch section for tests A and C. These stress components were used to calculate a value for each damage parameter. Again, it was assumed any plastic deformation was restricted to the first reversal. Even with uniaxial notches, knowledge of the theoretical stress concentration factor often renders conservative life estimates. This led to the development of fatigue stress concentration factors, K f, and the concept of a notch sensitivity factor, q. These developments are summarized 3

in reference [9]. In general, the sharper a notch, the higher the stress concentration factor, and the steeper the elastic stress-strain gradient. In essence, these concepts assume the maximum principal stress or strain is the relevant damage parameter and serve to reduce the maximum value of the damage parameter. As previously mentioned, the damage parameters identified in this study that provided some measure of correlation for the multiaxial data can at least conceptually be considered shear based. As such, the constraint inherent in a circumferentially notched round specimen would lead the maximum principal stress concepts to overestimate the effect of the notch. Slavic [2] and Taylor [1] acknowledge that notches may not be as severe as anticipated, and postulate some critical region over which the damage should be considered. Figures 9 through 12 show the radial damage predicted for each damage algorithm utilizing the individual stress-strain components determined from the finite element analysis. The solid point on each curve shows the location of the damage parameter calculated using the experimental life and the baseline damage curve. The fact that these points are all located subsurface indicates that the damage predicted employing surface stresses and strains is conservative. They also render some insight as to a depth over which the damage needs to be assessed. Using.5 mm increments, the subsurface damage was calculated, and Eq. (6) was utilized to calculate a percent error for each test. The optimized depth, which best correlates all the notched specimen data to the uniaxial baseline curve, was chosen to minimize the percent error. In other words, the mean of the notched data was chosen to correspond to the baseline uniaxial data. These results, along with the surface damage calculations, are shown in Figures 13 through 16 (left plot) for each of the damage parameters under consideration. Notch acuity was shown to be influential in uniaxial experiments and theories previously presented. The observation that, in these tests, the torsional and axial stress concentrations differ could cause some of the depth discrepancy observed here. Depending on the parameter evaluated, the optimal damage depth varied from.25 mm to.45 mm. While these dimensions seem large compared to prior values assigned in the literature, it should be noted that the stress concentration factors in the current experiments were lower than most considered in the literature. The low value of K t also tends to cause the stress-strain gradient to shallow, but cover a significant portion of the net section of the specimen (Fig. 8). Hence, the dimensions cited here should be interpreted in the context of the geometry of notches tested. Rather than assign a specific depth value to the damage, a third methodology involves integrating the value of the damage parameter over a depth beneath the surface, r*, to obtain an averaged damage. D amage = 1 r * r o (Damage dr (7) r o r * where r o is the outer radius at the notch section. The results of this analysis are presented in Figures 13 through 16 (right plot), and closely resemble the results of the point-depth analysis. A similar increment in depth and methodology to place the mean notched life on the uniaxial baseline curve was employed in this analysis. The range of depths, or r*, ranged from.4 mm with the modified Manson-McKnight parameter to 1.2 mm for the Fatemi-Socie-Kurath parameter. Again these are somewhat higher than those reported in the literature. A statistical comparison of the point and line depth averaging techniques is summarized in Table 5. As before, lower values of the mean and standard deviation are desirable. ) 4

While the line averaged results look similar to the point results, the line average method has the conceptual advantage of somewhat accounting for the differences in axial and torsional stress concentration factors that result in different stress and strain gradients. The stress state of the geometry tested is two dimensional at the surface, but three dimensional below the surface. Neglecting the circumferential stress would have introduced significant errors in the calculations, but the effects of the radial stress below the surface were negligible. Due to the symmetry of the loading and stress state near the notch, the critical plane orientation remained approximately constant over the depths considered here. For more complex loading and a general geometry, an averaging scheme over a critical volume of dimension V* may be required. However, it seems that averaging or integrating the value of the damage parameter has inherent advantages over modifying the stress-strain state, especially for more complex loading. A major drawback at this juncture is that no overall methodology has been forwarded to determine the critical distance, line length or volume for a general multiaxial loading and geometry. CONCLUSIONS As with uniaxial loading, the theoretical stress concentration factors overestimate the effects of fatigue for multiaxial loading. The Findley critical plane approach provided the best overall representation of the smooth bar data, and also for the notched specimen data using the line averaging technique. Empirical determination of the depth or line length for averaging needs to be refined. ACKNOWLEDGMENT This work was supported under Air Force Contract No. F4962-99-C-7, Advanced High Cycle Fatigue Life Assurance Methodologies, with Dr. Jeffrey Calcaterra, AFRL/MLLM, Project Monitor. The contract is being administrated through the University of Dayton Research Institute under subcontracts RSC9913 with North Dakota State University and RSC987 with the University of Illinois. REFERENCES 1. Krgo, A., Kallmeyer, A. R., and Kurath, P., "Evaluation of HCF Multiaxial Fatigue Life Prediction Methodologies for Ti-6Al-4V," Proceedings of the 5 th National Turbine Engine High Cycle Fatigue Conference, Arizona, 2. 2. Slavic, D. C., Dunyak, T., Griffiths, J., and Kurath, P., "Fatigue Crack Initiation Modeling in Ti-6Al-4V for Smooth and Notched Geometries," Proceedings of the 5 th National Turbine Engine High Cycle Fatigue Conference, Arizona, 2. 3. McKnight, R. L. and Slavik, D. C., General Electric Aircraft Engines, personal communication, 2. 4. Findley, W. N., Fatigue of Metals Under Combinations of Stresses, Transactions, ASME, Vol. 79, 1957, pp. 1337-1348. 5. Fatemi, A. and Socie, D. F., "A Critical Plane Approach to Multiaxial Damage Including Out-of-Phase Loadings," Journal of Fatigue and Fracture of Engineering Materials and Structures, Vol. 11, No. 3, pp. 149-165, 1988. 6. Kurath, P. and Fatemi, A., "Cracking Mechanisms for Mean-Stress/Strain Low Cycle Multiaxial Fatigue Loadings," Quantitative Methods in Fractography, ASTM STP 185, B. 5

M. Strauss and S. K. Putatunda, Eds., American Society for Testing and Materials, Philadelphia, PA, pp. 123-143, 199. 7. Chu, C. C., Conle, F. A., and Bonnen, J. J. F., "Multiaxial Stress-Strain Modeling and Fatigue Life Prediction of SAE Axle Shafts," Advances in Multiaxial Fatigue, ASTM STP 1191, D. L. McDowell and R. Ellis, Eds., American Society for Testing and Materials, Philadelphia, pp. 37-54, 1993. 8. Peterson, R. E. Stress Concentration Factors, Wiley Interscience, 1974. 9. Peterson, R. E., Chapter 13, Metal Fatigue, Sines & Waisman, Eds., McGraw-Hill, 1959. 1. Taylor, D., Geometrical Effects in Fatigue: A Unifying Theoretical Model, International Journal of Fatigue, Vol. 21, pp. 413-42, 1999. TABLES AND FIGURES Table 1 Curve-Fit Constants to Uniaxial Data Model A (ksi) b C (ksi) d k or w Modified Manson-McKnight 7611 -.6471 65.39 -.3582.433 Findley 15,97 -.7368 69.2 -.3959.35 Fatemi-Socie-Kurath 7.54 -.6467.1264 -.2647 6. Modified Chu-Conle-Bonnen 424.4 -.6475.676 -.3152 2.7 Table 2 Additional Biaxial (Smooth Specimen) Data Spec. ID Test Description Δγ (%) Δε (%) N f (cycles) 142-5 R = 1 torsion 1.22 26,657 178-3 R = 1 torsion 1.12 843,44 142-9 R =.1 torsion 1.12 184,669 142-8 R ε = R γ = 1 box path.82.6 59,432 178-7 R ε = R γ = 1 box path.82.6 72,36 142-11 R ε = 3, R γ = 1 box path.82.6 182,149 178-8 R ε = 3, R γ = 1 box path.82.6 212,337 178-12 Check path 1.22.3 5,568 6

Table 3 Statistical Comparison of Multiaxial Models for Smooth Specimen Data Uniaxial Data (% Error) Biaxial Data (% Error) Model Mean St. Dev. Mean St. Dev. Modified Manson-McKnight -.54 7.51-5.91 11.75 Findley -.68 8.37 -.15 9.36 Fatemi, Socie, Kurath -2.42 16.27 2.86 12.25 Modified Chu, Conle, Bonnen -2.47 17.26 1.36 19.3 Table 4 Notched Specimen Test Conditions and Fatigue Lives Spec/Path P max (lb) P min (lb) T max (lb-in) T min (lb-in) N f (cycles) 132-4/A 2 2 75 75 369,7 132-3/B 2 2 75 75 >1.8 1 6 132-1/C 15 75 627, 132-2/D 9 1 75 9 123,8 132-5/E 65 65 55 55 416,2 Table 5 Statistical Comparison of Multiaxial Models for Notched Specimen Data Model DP calculated at notch root surface (% Error) DP calculated at critical depth (% Error) DP averaged over critical length (% Error) Mean St. Dev. Mean St. Dev. Mean St. Dev. Mod. M-M 8.84 8.47.51 11.24.43 1.45 Findley 16.3 6.39.47 7.44 1.1 8.6 FSK 32.4 12.77 -.67 2.93 -.75 19.39 Mod. CCB 31.6 12.45.3 18.84-1.37 21.69 τ τ τ τ σ σ σ 9 out-of-phase Triangle path Box path Check path σ Figure 1. Non-proportional load paths for smooth biaxial tests. 7

8 7 6 5 4 3 2 1 Torsion (R=-1) Torsion (R=.1,.5) Proportional 9 out-of-phase Triangle path Box path Check path 8 7 6 5 4 3 2 1 Torsion (R=-1) Torsion (R=.1,.5) Proportional 9 out-of-phase Triangle path Box path Check path 1.E+4 1.E+5 1.E+6 1.E+7 1.E+4 1.E+5 1.E+6 1.E+7 Figures 2 & 3. Damage parameter vs. life curves for modified Manson-McKnight (left) and Findley models (right), showing smooth biaxial data (curves represent uniaxial data). Damage Parameter.25.2.15.1.5 Torsion (R=-1) Torsion (R=.1,.5) Proportional 9 out-of-phase Triangle path Box path Check path 1.4 1.2 1.8.6.4.2 Torsion (R=-1) Torsion (R=.1,.5) Proportional 9 out-of-phase Triangle path Box path Check path 1.E+4 1.E+5 1.E+6 1.E+7 1.E+4 1.E+5 1.E+6 1.E+7 Figures 4 & 5. Damage parameter vs. life curves for Fatemi-Socie-Kurath (left) and modified Chu-Conle-Bonnen models (right), showing smooth biaxial data (curves represent uniaxial data). T T T B C A D E P P P Figure 6. Load paths for notched specimen tests. 8

Figure 7. Finite element mesh of notched specimen, and sample plot of von Mises stress distribution (test A). 5 45 4 35 Tangential Stress Axial Stress Shear Stress 7 6 5 Shear Stress Stress (ksi). 3 25 2 Radial Stress Stress (ksi). 4 3 15 1 5 2 1.5.1.15.2.25.5.1.15.2.25 Figure 8. Plots of stress distribution vs. depth under the notch for tests A (left) and C (right). 6 5 4 3 2 1 Test A Test B Test C Test D Test E Calculated Depth 6 5 4 3 2 1 Test A Test B Test C Test D Test E Calculated Depth.5.1.15.2.25.5.1.15.2.25 Figures 9 & 1. Damage parameter vs. depth under notch for modified Manson-McKnight (left) and Findley models (right), and depth at which damage parameters correspond with experimental lives. 9

Damage Parameter.24.21.18.15.12.9.6.3 Test A Test B Test C Test D Test E Calculated Depth 1.2 1.8.6.4.2 Test A Test B Test C Test D Test E Calculated Depth.5.1.15.2.25.5.1.15.2.25 Figures 11 & 12. Damage parameter vs. depth under notch for Fatemi-Socie-Kurath (left) and modified Chu-Conle-Bonnen models (right), and depth at which damage parameters correspond with experimental lives. 8 8 7 6 5 4 3 Notched (9.8 mil depth) 2 1 1.E+4 1.E+5 1.E+6 1.E+7 7 6 5 4 3 2 Notched (15.8 mil averaged) 1 1.E+4 1.E+5 1.E+6 1.E+7 Figure 13. Comparison of damage parameters from modified Manson-McKnight model calculated at surface, at depth of 9.8 mil (.25 mm), and averaged over 15.8 mil (.4 mm) depth. 1

7 7 6 5 4 3 2 Notched (13.8 mil depth) 1 1.E+4 1.E+5 1.E+6 1.E+7 6 5 4 3 2 1 Notched (35.4 mil averaged) 1.E+4 1.E+5 1.E+6 1.E+7 Figure 14. Comparison of damage parameters from Findley model calculated at surface, at depth of 13.8 mil (.35 mm), and averaged over 35.4 mil (.9 mm) depth..25.25 Damage Parameter.2.15.1.5 Notched (17.7 mil depth) 1.E+4 1.E+5 1.E+6 1.E+7 Damage Parameter.2.15.1.5 Notched (47.2 mil averaged) 1.E+4 1.E+5 1.E+6 1.E+7 Figure 15. Comparison of damage parameters from Fatemi-Socie-Kurath model calculated at surface, at depth of 17.7 mil (.45 mm), and averaged over 47.2 mil (1.2 mm) depth. 1.4 1.4 1.2 1.8.6.4 Notched (17.7 mil depth).2 1.E+4 1.E+5 1.E+6 1.E+7 1.2 1.8.6.4.2 Notched (39.4 mil averaged) 1.E+4 1.E+5 1.E+6 1.E+7 Figure 16. Comparison of damage parameters from modified Chu-Conle-Bonnen model calculated at surface, at depth of 17.7 mil (.45 mm), and averaged over 39.4 mil (1 mm) depth. 11