University of Dayton eCommons Graduate Theses and Dissertations Theses and Dissertations 2000 Development of nondestructive evaluation methods to characterize anomalous microstructures in Ti-6Al-4V Mark P. Blodgett University of Dayton Follow this and additional works at: https://ecommons.udayton.edu/graduate_theses Recommended Citation Blodgett, Mark P., "Development of nondestructive evaluation methods to characterize anomalous microstructures in Ti-6Al-4V" (2000). Graduate Theses and Dissertations. 1604. https://ecommons.udayton.edu/graduate_theses/1604 This Dissertation is brought to you for free and open access by the Theses and Dissertations at eCommons. It has been accepted for inclusion in Graduate Theses and Dissertations by an authorized administrator of eCommons. For more information, please contact mschlangen1@udayton.edu, ecommons@udayton.edu. DEVELOPMENT OF NONDESTRUCTIVE EVALUATION METHODS CHARACTERIZE ANOMALOUS MICROSTRUCTURES TO IN Ti-6AL-4V APPROVED BY: Danigf ( ) James A. Snide, Ph. D. Eylon, D. Sc. isory Committee Chairman Director and Professor, Graduate Materials Engineering Ofames 4. 8 Member Professor Emeritus, Graduate Materials Engineering Sathish Shamachary, Ph. D. Committee Member Adjunct Professor, Graduate Materials Engineering Joseph P. Gallagher, Ph. D. Committee Member Professor, Graduate Materials Engineering Muhammad N. Islam, Ph. D. Committee Member Associate Professor, Mathematics Department Blake Cherrington, aN Dean, School of WAN Donald L. Moon, Ph. D. Associate Dean, Graduate Engineering Programs & Research School of Engineering il ABSTRACT DEVELOPMENT OF NONDESTRUCTIVE EVALUATION METHODS TO CHARACTERIZE ANOMALOUS MICROSTRUCTURES IN Ti-6AL-4V Author: Blodgett, Mark, Patrick University of Dayton, 2000 Advisor: Dr. D. Eylon main The of this objective dissertation is to confirm research through the following hypothesis: “The use of nondestructive evaluation tools allows the detection of different microstructure allows the identification of microstructure types and anomalies (interior and surface) in metals and alloys.” The work was conducted on Ti-6Al-4V forged bar stock, presenting a case study for a high performance structural alloy. Ti-6Al-4V anomalies is model a good in demanding material, which cannot tolerate microstructure alloy is well established with extensive applications. The documentation on physical, chemical, and mechanical properties. It is also available in many different microstructures, addresses issues concerning readily generated by heat treating. characterization microstructure This dissertation and the identification of microstructural anomalies. Specifically, this work includes i) background research on the identification of ultrasonic and electrical characteristics of five different Ti-6Al-4V microstructures; ii) an application of ultrasonic backscattering measurements to detect ili diffusion bonded Ti-6AI-4V microstructure changes, to simulate locally isolated remnant cast structure for billet NDI; iii) original research on laser interferometric detection for ultrasonic phase mapping to characterize macroscopic and iv) texture in Ti-6Al-4V; original research on eddy current electrical conductivity mapping in titanium alloys. methods were developed to evaluate microstructure and Three original NDE microstructure anomalies in Ti-6Al-4V. a forward First, scattering measurement technique was developed to spatially map the incoherent grain scattering in the forward propagation direction. These results showed, for the first time, that mapping of the forward scatter provides a basis for characterization of texture in polycrystalline titanium alloys. Second, a laser interferometric system was developed to map the signal amplitude and phase of the transmitted acoustic crystallographic orientations. This new field in development samples with different preferred allows microscopic variations in phase (and amplitude) to be experimentally measured, which is vital to understanding statistical characteristics of NDE data due to microstructure. The method is original as this the first time the signal phase has been mapped with a sub-wavelength aperture for materials characterization. The third method was discovered serendipitously, as a result of applying an eddy current probe to the surface of different microstructures of titanium alloy samples. As a result, a new and original noncontacting approach was developed. The method surface characterization is based on the effects of grain anisotropy on electrical conductivity in titanium alloys. This electrical property imaging method allows for characterization of near-surface microstructure and processing anomalies. iv ACKNOWLEDGMENTS Special appreciation is in order to my advisor, Dr. Daniel Eylon for his generous support and direction of this dissertation topic, and to the committee members for following this work to conclusion. I would like to recognize the essential support of the Materials and Manufacturing Directorate for allowing me to pursue this project, especially to Dr. Walt Griffith, Mr. Tobey Cordell, and Dr. James Malas. A particular debt of gratitude goes to Dr. Peter Nagy of the University of Cincinnati for his skillful experimental support and also to Dr. Michael Gigliotti of the General Electric Company for his helpful assistance in the generation of diffusion bonded composite microstructure samples. I would also like to acknowledge the support given by Dr. Waled Hassan from the United Technologies Company for his help with the analytical calculations and finite element analysis of the eddy current data. Finally, my sincere thanks goes to my parents and family for their patience and support throughout this research project. TABLE OF CONTENTS sO 0) CO SO SOCOSOSOOOOOSS i EEE EERE EE EEE EEE ERE ETE il EEE EE EEE APPROVAL. .o. cece cnc cece cece eee nee nee e nee een ABSTRACT oc ccccec ccc cc eee e eee ene nen een een EEE ACKNOWLEDGMENTS. EEE EEE EEE EEE EEE EEE EE EE EEE il nena eats Vv ce eens e renee eens e nent eee e cece eee e eens cccc ........ LIST OF FIGURE CAPTIONS. ....... 0. ce cece cece een e nen e tence een n ene e eter e tenes Vili LIST OF TABLES 0....... cc cece ccc e cece erence eee nee en nee nnn r enn e eee gees X1V LIST OF SYMBOLS AND NOMENCLATURE eee e eens XV eee ee eeeee ..........:e:e CHAPTER I. INTRODUCTION ........... cece ccc e cee ee eee e nen een entero nena ee en nase 1 1.1 1.2 1.3 1.4 1.5 1.6 OVErvicW oo. cece ccc ccc cece ee eee eee nner e een n nen nee eee EEE EOE reer e EERE EEE EE EE 1 The Problem ......... 0.0... ccc ce nsec eee eee renee eee teen Een ene EEE EEE EE EEE Ean 4 neces aeenaes 5 eceec een ee een eeeennen eee eee nee eee eeeccecc eee ....c The Hypothesis .......... renee enn e aes 5 eee renee renee ... ccc cece cece cence eter eee The Ti-6AI-4V Alloy... e 7 anne eee nen eeaeaeags een een eee eece eee ee reece nentcccc cee ee .... Alloy Processing ........ 9 eeee eeneees ee neee ne nent ne .cecee eee sence ...... Ultrasonic Alloy Characterization ...... CHAPTER II. LITERATURE REVIEW eeereas 14 ee ere ne nee eeeene eeecececce ............:c ee tenes 15 eee ee eee ce ...cce eee etter 2.1 Fundamentals on Ultrasonic Scattering *........... eens 19 seeee nese 2.2 Wave Propagation in Inhomogeneous Materials ..............:sseee 2.3 Ultrasonic Backscattering for Structural Characterization .................:6665 20 2.4 Diffusion Bonding of Titanium Alloys e 22 eee ee nent eee eeees cece eee eeeeee ...........: CHAPTER III. MICROSTRUCTURAL EFFECTS ON ULTRASONIC PROPERTIES ...........c cece eee e cece cece eee ene en neeeeeseenaeeeeaenaes 27 Microstructures Evaluated ccsees 27 ence essen ee eeneneeeeeee cece cece eee eee e ...cc .......... 3.1 scenes 34 PART I: Conventional Ultrasonic Evaluation Methods ................:cceceee 3.2 Ultrasonic Velocity ........... ccc cc cece cece nee tenner n erent eet et teen eae 34 3.3 Ultrasonic Attenuation eee eneeees 46 e eee eens een ee cce ence teen ccecec ...........c Vi 3.4 Ultrasonic Grain Scattering ............ ccc cece cece e eee eee e eee n eee ee ee tens 54 eet eeeees 62 eenseeeeeee PART II: The New Forward Scattering Method ..............:cce eee eee e ene nee 62 cece ence nee ee.... .... cece 3.5 Ultrasonic Forward Scattering .... CHAPTER IV. SUBSURFACE MICROSTRUCTURE ANOMALY DETECTION IN BULK TITANIUM ALLOY USING UT BACKSCATTERING 74 4.1 Introduction ..........ccceccc cece eect e eee eee nee e ene e eee ene 4.2 Sample Development een E EEE EEE EEE Se nee es 74 ee es 76 anrenenetin cceee e eect eet eaeeenneeeenet reece eee....c .......... enna eeeees 80 .cc cece eee eee ..... 4.3 Detection of Anomalous Microstructure .....cece e nena eee e ene eee ee 85 ene ee eee eeecccc ...c cece ence eee e near 4.4 Results and Discussion ........ 4.4.1 Axially Bonded Samples 4.4.2 Radially Bonded Samples eeeeneaes 86 eee nese ee eeeeee eee eee eee..c:ec ............ eee ecees 92 senses en ee eec :seceee sense .............: AS SUMMALY .eecccceccsesecsccesseesecsceceeesecseceeceseeeeesscceseesesectecseeenesaeeass 101 CHAPTER V. THE INFLUENCE OF TEXTURE AND PHASE DISTORTION ON ULTRASONIC ATTENUATION IN Ti-6AI-4V 5.1 5.2 5.3 5.4 5.5 Introduction ..........-604- 104 .......ccccccee cece cee e cnet cee ee ee en eee rene ee ee eee eee ene e ee en eee eee ene Ee ee es 104 teens 109 ence eee eeees Ultrasonic Properties of Mill Annealed Ti-6AI-4V ..........eece es 121 neeeeene .:ceceeeeenee ............. Laser Detection of Ultrasonic Phase Distortion eees 125 ece enereeeeereeen cceee ce eneeeneeeeeeen .cccc cence Results and Discussion .......... eats 134 Ener e een TEE n eee EEE nee n eee e een eee e eee eee SUMMAry oo... cece CHAPTER VI. DEVELOPMENT OF AN EDDY CURRENT MATERIALS CHARACTERIZATION METHOD FOR TITANIUM ALLOYS..........-..-50005 138 6.1 6.2 6.3 6.4 6.5 Introduction 22.0.0... cece ccc cece eee e ence eee eee ee eee e eee e een ee eens nese renee enna ges 138 teense nes caies 146 cette eee e tennereeeeee ees ee:::cee Eddy Current Experiments ............ eeenenes 155 e eeeeene nee ee :ceecee renee ......: Resolution of Eddy Current Imaging ....... e eee ee ene n enaetets 164 eee n een ee enecceee erence ence..cce Results and Discussion .......... SUIMMALY 2... cece cece cece eee nee ee eee eee eee ee neste eee ee eee eeeeer eee eeceee eee ttty 175 CHAPTER VII. CONCLUDING REMARKS. ....... 00. ccece cee eee cnet een e eens 06 9 S100 7 7.2 CONCIUSIONS 179 SOSOOSOROOOOOOOCOOOOOOOCOOOOOOD 179 EEOC 1.0... cece cee cc cece eee e eee eee eee Eee Eee EEE ene DETTE EEE EEE EEE EERE 182 7.3 Comparison of Results for Microstructure Anomaly Characterization ......... 185 7.4 Future Nondestructive Materials Characterization Research .............-...5+5+ 189 7.4.1 Application and Development of Array Transducers ..............+++- 189 eee ee ee eeneees 190 esses cceeeee 7.4.2 Geometry Insensitive Techniques ............:s 7.4.3 Residual Stress Gradient Measurement ..............csceeeeeeeeeeeeeeees 191 7.4.4 Forward Scattering Using Dual-Transducer Articulation .............. 192 APPENDICES. EE EEE 193 ener eee eset Eee ne nee n enna e eensecen ee ee ee ence cee ceec eee ne eee ccce ......cc A. Data Tables .........c cece ccc cee ee eee cnet eee e een nee ne seas een eneeeeeeneens 193-203 Vii LIST OF FIGURE CAPTIONS 3.1. The as-received condition (mill annealed) of Ti-6Al-4V bar stock in the transverse (a) and axial (b) orientation. Both microstructures shown at 200X. 3.2. Microstructures resulting from the duplex anneal (a) and the recrystallization anneal (b). Both microstructures shown at 200X and both have nominally the same primary o grain size of approximately 20 pm. 3.3. The beta annealed microstructures: (a) fine beta annealed structure shown at 200X, exhibiting approximately 300-500 ym prior beta grain size, and (b) coarse beta annealed structure shown at 25X, with extremely large approximately 2-3 mm prior beta grain size. 3.4. The geometric configuration of the birefringence measurement with different polarization and wave propagation directions. 3.5. Results showing the variation in average longitudinal wave velocity for the various titanium alloy samples. The scatter on these measurements is generally about + 0.5%. 3.6. Diagram showing the different Ti-6A1-6V sample sections cut from the 2.5" diameter cylindrical bar (i.e., five samples having varying surface normals relative to bar axis). 3.7. The hexagonal principal directions preferentially rotate to the radial direction (x, y plane) in the as-received titanium bar stock, based on longitudinal velocity and x-ray analysis. 3.8. Shear wave velocity data taken at different polarization angles consistent with the slow and fast pure shear modes. 3.9. This diagram shows the fast and slow pure mode polarization direction in the 90° samples. Note the fast mode occurs when particle displacement is perpendicular to the alignment bands in the axial direction of the bar. 3.10. Illustration of the ultrasonic wave propagating back and forth inside the material and the associated waveform showing the multiple coherent echoes. 3.11. The acoustic field of a circular piston source variations at discrete distances from the transducer. Vili showing the pressure 3.12. Frequency dependent attenuation losses for the titanium alloy samples in the (a) radial direction, and (b) axial direction. 3.13. signal traces showing Oscilloscope the difference in amplitude of the microstructure scatter from the mill annealed samples. 3.14. Schematic of the two different experimental set-ups used for microstructure characterization using backscattering. 3.15. Spatial averaging of backscattered longitudinal waves in the time domain, using broadband excitation with a transducer having a center frequency of approximately 8.5 MHz. 3.16. Schematic of the forward scattering experiment. 3.17. Backscattered signals from sample AR(90) showing the effect of increasing the amplification by 40 dB to reveal the backscatter signals. 3.18. Spatially averaged backscattered signals from the axial and transverse orientations, using flat entry surfaces with beam focused on front surface. 3.19. Time-lapsed C-scan images mapping the amplitude and divergence of forward scattering in mill annealed Ti-6Al-4V, showing clearly distinguishable differences between the axial versus radial directions. 3.20. Time-lapsed C-scan images mapping the amplitude of and divergence forward scattering in coarse beta annealed Ti-6A1-4V, showing indistinguishable differences between the axial versus radial directions. 3.21. Notional scattering cross-sections in the axial and radial directions for anisotropic and random structures, which were generated based on the backward and forward scattering results. 4.1. Schematic for the axial diffusion bonded Ti-6A1-4V samples. 4.2. The five axially diffusion bonded samples. 4.3. Schematic for the radial diffusion bonded Ti-6AI-4V samples. 4.4. The nine radially diffusion bonded samples. 4.5. Metallographic results comparing the original as-received microstructure with that of the HIP'ed sample. Both images were taken at 200x from the same orientation of the bar. 4.6. Metallographic AR/CBA) comparing results from an alpha dissimilar contaminated microstructures bond samples with a clean bond. (i. e., Both images were taken at 100x from the same orientation of the bar. 4.7. A backscattered ultrasonic B-scan shown without the advent of color coding. 4.8. An illustration showing the relationship between the B-scan image, the backscattered waveform, and the associated color scale. 4.9. An ultrasonic waveform which was collected through the curved surface on the side of the radially bonded AR/CBA/0.9" sample. 84 bonded 85 4.11. An illustration showing how the transducer reference signal is removed from the B-scan, leaving primarily only the material response. 4.12. Enhanced ultrasonic B-scans using equalized histograms from the axially 87 4.10. The B-scan configurations used for the two sets of diffusion samples. 88 bonded sample series. AR/AR shows a clearly evident bondline signal. AR/BA clearly demonstrates the typical change in the scattering characteristics between two different microstructures. 4.13. Enhanced ultrasonic B-scans from the AR / X / AR sample series. Both the FBA and CBA implant images show indistinguishable characteristics, with clear 90 indications of scattering changes between the joined microstructures. 4.14. B-scans in the axial direction through the curved surface on the side of the 93 AR/AR/X bonded cylinders focused at surface at 10 MHz. 4.15. Enhanced B-scans from the AR / AR / X / x sample 94 anomaly, designates the diameter of the implanted series, where X and x indicates the scan orientation (r - radial). 4.16. Enhanced ultrasonic B-scans from the AR / FBA / X / x sample series, where X designates the diameter of the implanted anomaly, and x indicates the 96 scan orientation (a - axial). 4.17. Enhanced ultrasonic B-scans from the AR / FBA / X / x sample series, where X designates the diameter of the implanted anomaly, and x indicates the 97 scan orientation (r - radial). 4.18. Enhanced ultrasonic B-scans from the AR / CBA / X / x sample series, where X designates the diameter of the implanted anomaly, and x indicates the 99 scan orientation (a - axial). 4.19. Enhanced ultrasonic B-scans from the AR / CBA / X / x sample series, 100 where X designates the diameter of the implanted anomaly, and x indicates the scan orientation (r - radial). 4.20. Correlation lengths of the radially bonded samples in the axial direction. 101 5.1. The as-received (mill annealed) condition of Ti-6Al-4V bar stock from 110 surfaces of the (a) axial (0°) and (b) transverse (90°) samples. Both microstructures are shown at 200x. 5.2. Longitudinal wave velocity as a function of orientation for 1.5"-thick mill annealed Ti-6A1-4V samples. 111 5.3. Shear wave velocity as a function of orientation for 1.5"-thick mill annealed 112 Ti-6Al-4V. 5.4. Orientation dependence of the average attenuation loss versus frequency for 115 1.5"-thick mill annealed Ti-6Al-4V samples. different 118 structure on 119 5.7. C-scans of the forward scatter in the axial (a) and radial (b) directions, taken over a 1" x 1" area at 10 MHz with a 0.5"-diameter, 3"-focal length receiver, 120 5.5. backscattering Average intensity a as function of time in orientations of mill annealed Ti-6AI-4V. 5.6. Schematic diagrams showing the effect of the macroscopic scattering for the 0 and 30 degree samples. focused on the surface. 5.8. Schematic diagrams of the experimental set-ups used to map phase (a) and 122 magnitide (b). 5.9. Schematic diagrams showing (a) three regions on the specimen over which the phase map is repeatedly folded back by 2n and (b) the variation in arrival 127 time at different positions across the wavefront. 5.10. Phase through (a) and corresponding 0.5"-thick samples magnitude of the mill (b) images taken at 9.7 MHz, annealed Ti-6Al-4V from 128 the axial direction, covering a 1" x 1" area. Same images are shown for the radial direction (c), (d). 5.11. Phase maps taken through a 0.2"-thick Ti-6Al-4V sample which had been heat treated to generate a coarse widmanstatten microstructure with very large lamellar « + B colonies. Phase images taken at (a) 7.5 MHz and the same region 129 at (b) 9.7 MHz. 5.12. Phase (a) and corresponding magnitude (b) images over a 1" x 1" area through a 1.5"-thick sample of mill annealed Ti-6A1-4V in the axial direction at 9.7 MHz. Same images are shown for the radial direction (c), (d). 5.13. Phase images taken through 1"-thick samples of the mill annealed Ti-6AI4V covering a 0.4" x 0.4" area using the 0.25"-diameter transmitter at 9.7 MHz. 130 132 Raw data from the axial direction (a) after removing the first 2n foldback, same for the radial direction (b). High-pass filtered data from axial (c) and radial (d) directions. 5.14. Comparison of attenuation measurements from a 0.5"-diameter unfocused 133 immersion transducer in pulse-echo, and from laser interferometric detection using a spot size of approximately 50 ym in through-transmission. 6.1. Eddy current images of small fatigue cracks in 2024 aluminum and Ti-6Al4V specimens (0.5" x 0.5", 2 MHz, 0.060"-diameter coil). Xl 144 6.2. Comparison of (a) optical, (b) eddy current, and (c) acoustic microscopic images of a coarse-grained Ti-6Al-4V sample (1" x 1") from nearly the same 146 area on the sample. 6.3. Electrical resistivity probability distributions for three single crystal surface 151 orientations in a) aluminum, b) copper, and c) cadmium and d) on the surface of polycrystalline Ti-6V-4V (solid lines are best fitting Gaussian distributions). 6.4. Scanned eddy current images of different Ti-6Al-4V microstructures; a) 154 sample containing a severe microstructure anomaly (right side, middle); b) the billet microstructure showing texture related features in the horizontal direction; c) a large grained sample; and d) equiaxed beta annealed microstructure (dimension 1" x 1"). 6.5. Magnetic field distribution produced by a small pancake coil in titanium at 158 four different frequencies. 6.6. Eddy current distribution produced by a small pancake coil in titanium at 159 four different frequencies. 6.7. Axial penetration depth versus frequency for a 1-mm-diameter pancake coil in titanium. The symbols represent the numerical results calculated by finite 161 element (FE) simulation, the solid line represents the general trend of the FE data, and the dashed line is the plane wave asymptote calculated from the standard penetration depth according to Eq. (6.3). 6.8. Radial penetration p versus frequency for a 1-mm-diameter pancake coil in 162 titanium. The solid circles represent the numerical results calculated by finite element (FE) simulation, the solid line illustrates the general trend of the FE data, the empty circles represent the analytical results calculated by Dodd and Deed's method, and the dashed line illustrates the general trend of the analytical data. 6.9. Corrected radial penetration Pop, = P — dep /2 (deg © 119 douter) 163 versus frequency for a 1-mm-diameter pancake coil in titanium. The solid circles represent the numerical results calculated by FE simulation, the solid line illustrates the general trend of the FE data, and the dashed line is the standard penetration depth calculated from Eq. (6.3). 6.10. Experimental impedance diagram in titanium at 2 MHz. The rotation angle 167 was chosen so that the lift-off curve is horizontal. 6.11. Experimentally determined lateral resolution versus inspection frequency by a commercial pencil-probe in titanium. The adjusted FE prediction for the radial penetration is plotted to indicate the trend of the data. Xii 168 6.12. Eddy current images (0.5" x 0.5") taken at three different frequencies to demonstrate the effect on lateral resolution. These images were scanned from an extremely large grained polycrystalline titanium alloy. 6.13. Impedance diagrams and resolution profiles at three different rotational angles relative to the horizontal lift-off angle in titanium at 2 MHz. 6.14. Experimentally measured lateral resolution versus rotation angle curves for three different frequencies in titanium. 6.15. Eddy current c-scan images of a coarse-grained Ti-6AI-4V specimen at 5 MHz and two different rotational angles (0.5" x 0.5"). CHAPTER I INTRODUCTION 1.1 Overview Many different nondestructive materials characterization techniques have been developed and applied in recent decades to address grain scattering, texture, and the detection of discrete manufacturing and service related flaws inside the bulk of different forged and cast structural components [1, 2, 3, 4, 5, 6, 7, 8]. Likewise, a great deal of work has been done to understand the physical and mechanical properties of titanium and its many different alloys and microstructures for high performance applications [9, 10, 11, 12, 13, 23]. The primary objective of traditional NDE is the detection and quantification of defects. However, the same nondestructive inspection approaches used to find flaws can also provide information on microstructure and microstructural variations inside metals, based on the characteristics of the NDE signals. Unfortunately, this NDE subject is not well developed nor used. The motivation for this research is based on the need to develop new NDE methods to address problems associated with the detection and quantification of interior and surface-connected microstructural anomalies originating from alloy production and processing. The main goal of this research is to develop new NDE methods. Structure related background noise is often a source of frustration to traditional nondestructive inspection as these signals can mask over other genuine flaws, such as voids or porosity [14, 15, 16]. This is especially true in cast and forged components due to the nature of the solid in terms of local elastic property variations. What makes the evaluation of structural damage such a difficult problem for cast and wrought components in general and titanium alloys in particular stems from effects the the material has on the NDE signals. Background signal noise is a problem that clearly hinders our ability to detect tiny, or worse, microscopic damage due to incoherent grain scattering [17, 18, 19]. The innovative and unique aspect of this work is the development of new NDE methods (not based on ultrasonic backscattering) to, rather than suppress the background noise signals, magnify and evaluate them as a source of information to characterize microstructure. NDE signals require careful collection and interpretation in order to relate to the microstructure anomalies. unique Moreover, probe, used and identify the presence these signals are the result of complex to sense elastic or electrical property of microstructure interactions between a variations, and a locally changing grain structure. The dissertation outline initially parallels the progress of sample development, starting with the evaluation of simple individually categorized microstructures in Chapter 3 and leading to the study of more complex, joined microstructures in Chapter 4, where standard ultrasonic backscattering was used as a basis for the detection and evaluation. Since from the results simulated anomaly the (implanted) were samples somewhat inconclusive (from the standpoint of actually discriminating the different anomaly types) further original research was conducted to develop new NDE methods. In Chapter 5,a new NDE method was developed to spatially map the signal phase of ultrasonic waves propagating through different orientations of textured titanium bar stock. In Chapter 6, an eddy current method was developed and applied for the characterization of grain structure in titanium alloys. The of development nondestructive inspection, and testing, materials characterization techniques is essential to evaluate the structural behavior, performance, and life span of components, which are known to accumulate damage throughout the [20]. This is true in titanium alloy components, course of service since they cannot tolerate internal defects or microstructural anomalies due to their high level of loading (21, 22] and the critical performance requirements in airframes and engines. Hence, these structures require periodic maintenance and nondestructive careful inspection to minimize the potential for defects, which could ultimately lead to catastrophic failures. The detection and quantification of microstructure anomalies is a problem of major importance to industry, but there are presently no definitively effective NDE methods nor calibration standards available to do this. New phenomena. NDE methods Experiments have been developed based on two independent physical were designed and conducted to examine both elastic and electrical property variations in titanium alloy samples. Results have been collected from conventional NDE However, techniques to benchmark the samples and from the new methods. results from the different experimental of experimental the combination categories into a single chapter did not make sense without a sound theoretical foundation for doing so. Therefore, starting with Chapter 3, the organization of the dissertation is its own chapter has such that each independent sections covering the introduction, experiments, results, and discussion. This approach was taken in order to avoid confusion regarding what versus is new, what been has done before in each of the areas investigated. 1.2 The Problem Many wrought alloys are prone to the development of interior microstructural irregularities, remnant of the original ingot cast structure that did not get globularized during the ingot breakdown stage. These may ultimately compromise the material’s structural integrity [23]. Such processing related problems could be easily overlooked due to the difficulty, time, and cost of developing and applying nondestructive inspections geared towards locating these, often subtle, microstructure signals. These anomalies could lead to a number of different problems including: i) the introduction of processing defects in subsequent operations due to stress induced porosity; ii) early introduction of damage due to mechanical cycling; and iii) difficult challenges for flaw detection due to the generation of background noise. The key reasons for the limitations of conventional NDE approaches are : a) microstructure anomalies are generally weak scatterers and are easily overlooked [24, signals from dependent and often lack uniqueness [16, 25]; b) the microstructure anomalies are frequency 17, 18]; c) discriminating between benign microstructure noise and genuine defects or damage is difficult due to lack of suitable calibration samples; and d) most NDE methods are simply not capable of looking deep inside materials for subtle microstructure anomaly indications. 1.3. The Hypothesis The hypothesis for this dissertation is: “The use of nondestructive evaluation tools allows identification the detection of of microstructure different microstructure anomalies (interior alloys.” This work centers on the measurement types surface) and allows the in metals and and of elastic and electrical properties in samples processed to different microstructures, textures, and orientations. The work is supported by experimental of conventional application methods, carried out on numerous the and techniques samples development of new of uniform comprised microstructures and composite microstructures, generated via diffusion bonding. 1.4 The Ti-6AI-4V Alloy Ti-6Al-4V, one of the first titanium alloy's to be developed, is an all-purpose alpha + beta alloy commonly used in high performance structural applications such as aircraft gas-turbine engine compressor blades and disks. The alloy also has an excellent combination of strength and toughness, excellent corrosion resistance, and biocompatibility for orthopedic applications [10]. The mechanical properties of titanium alloys depend on the chemistry, microstructure, and proper design and selection. These parameters influence the strength, toughness, environmental resistance, and the fatigue crack initiation and propagation characteristics of the material. Microstructure related noise is an important NDE consideration because it largely determines the detectability limits of conditions anomalous the within structure. Ultrasonic grain noise stems from scattering in the polycrystalline structure as the wave propagates through the material [26]. Scattering sites originate from the crystal-to-crystal variations in density and elastic constants. For single phase materials, each crystal has the same density and the same crystalline structure. Hence, scattering in these materials arises mainly from the variation in velocity resulting from the anisotropy of the single crystal elastic constants microstructural and from the random noise arises due to the of nature constructive grain and orientation. destructive Polyphase interference of scattered signals from thousands of different combinations of individual grains and grain colonies [27]. Grain scattering signals in such complex microstructures are generally used only as an indication of structural differences, since the signals are not unique in terms of relating quantitative features of the microstructure. Ti-6Al-4V was developed at the U. S. Army's Watertown Arsenal Laboratories in 1954 [28] and is still the most commonly used titanium alloy all around the world. The alloy is high strength, lightweight, corrosion resistant, relatively inexpensive, and can be easily acquired from a number of vendors in various grades of purity. In stock materials, the problem of microstructure anomalies is minimal because the sections are generally small, allowing fabrication processes and heat treatments to refine all remnant microstructure anomalies. In titanium components requiring large cross-sections like in gas-turbine engine fan disks or forged bulkheads, complete microstructural uniformity is due to processing difficult to achieve complexities, but advanced techniques geared toward more complete homogenization and refinement are under consideration [29}. 1.5 Alloy Processing All structural metals begin by undergoing generally chemical treatments directed towards removing composition to allow best the combination a series of thermal and impurities and adjusting the chemical of performance, reliability, and cost effectiveness for a specified application. Cast metal ingots derived from the refining and alloying processes ultimately yield stock materials (billet, bar, plate, sheet, etc.) via the application of additional fabrication processes like rolling, press cogging, and extrusion [9, 10]. The ingot material is generally not used in its as-cast state due to obvious performance limitations originating from the nonuniform, highly inhomogeneous nature of the coarse cast structure both locally and throughout the ingot volume [13, 23]. The presence of localized microstructure flaws resulting from impurities, inclusions, voids, or incomplete processing of the cast structure can compromise the ingots are strength related properties of stock materials. As such, titanium thermomechanically transformed into billet, bar, and plate. Hot working is the only option for the hexagonally symmetric Ti-6AI-4V alloy as lower temperature or room temperature working is difficult and leads to cracking. Hot working is more forgiving than cold working in terms of processing related defect generation due to the ease of deformation, a but small inaccuracies in the processing time or temperature can result in structural inadequacies developed subsequently in the to due components, presence of manufacturing defects. For Ti-6AI-4V, the first breakdown of the large cast ingot structure is generally a press cogging, performed above the beta transus (~1000 °C) which is geared towards reducing the cross-section and grain refinement [9, 13, 23]. Critical aircraft components such as aircraft gas-turbine engine fan and compressor disks demand specially developed billet stock with higher and uniformity of the to allow long-term reliable performance of the refinement alloying purity and more microstructure throughout its volume component. Nondestructive evaluation of premium grade rotor stock material is crucial to determine the suitability of material for further fabrication to avoid the generation of manufacturing defects. This incompatibilities and voids is especially can form, true due in forging operations to microstructural where irregularities strain in high deformation regions. The processing industry is familiar with the pitfalls of titanium processing and the current trend is to develop smaller [approximately 10" (250 mm) diameter] billet sections in order to achieve structural uniformity throughout the volume [29, 30]. A component may contain conglomerations of grains with undesirable extremes in the mechanical properties due to localized texture, which could lead to faster damage initiation in dwell applications [31]. Similarly, voids, stress-induced irregularities [30]. a component may contain manufacturing defects such as porosity, or cracking related to hard alpha compositional 1.6 Ultrasonic Alloy Characterization A variety of microstructural features can account for scattering losses from both ultrasonic and eddy current methods. Grain size, shape, and anisotropy for example are key characteristics of the microstructure and play a strong role in acoustic [32] and electrical grain scattering [33, 34]. Diffraction losses are due to interference effects from edges, corners, and other features that generate wave fringes, phase shifts, or frequency shifts. Beam spreading involves the widening distribution of energy in the propagating wave and the transition from planar waves to spherical waves. To further complicate the scattering problem, it is known know Ti-6AI-4V has a relatively complex microstructure. A significant obstacle to detecting and characterizing incipient or low-level damage is grain noise, which evolves due to random local variations in the crystallography and density of the microstructure. In Ti-6Al-4V, the microstructural complexity poses significant challenges for materials characterization, stemming from the alloys two phase structure with different crystallographic symmetry, density, and stiffness between the alpha and beta phases. Moreover, the range useful titanium alloy microstructures is diverse and widely distributed in terms of the types of features, which can generate scatter. Grain scattering of ultrasonic waves occurs because most metallic materials are simply not homogeneous. Virtually all structural metal components are polycrystalline, being comprised of a large number of randomly oriented individual crystals. Each grain is individually a homogeneous single crystal, which in general has anisotropic properties. In some alloys, fabrication processes tend to impart long-ranging crystallographic texture. Crystal discontinuities at grain and twin boundaries tend to deflect small amounts of Undeformed energy out of the main ultrasonic beam. single crystals have a uniform lattice structure that is generally characterized by three orthogonal axes, along which the various properties, like elastic moduli, are unique. For the case of Ti-6A1-4V, each grain of the primary constituent phase has hexagonal symmetry and is characterized by a principal direction, i.e. normal to the basal plane (plane of isotropy). However for Ti-6Al4V, the five single crystal elastic constants (alpha) are not available in the open literature. An interesting aspect of attenuation in polycrystalline materials has to do with the influence of microstructure and texture on phase distortion of the propagating wave [35, 36, 37]. The phase distortion is related to the arrival time variations of rays on different acoustic paths, through different grains and has a major, yet uncommon, role in attenuation, especially in anisotropic materials. This attenuation factor is not very well understood because the vast majority of experimental methods are based on amplitude detection and the detection devices generally are phase sensitive [38]. Usually, attenuation is considered in the context of a loss of energy due to scattering, but losses associated with phase distortion represents a separate attenuation mechanism, which is addressed in detail by this work in Chapter 5. 10 Chapter 1, References: 1. E. P. Papadakis, “Ultrasonic Attenuation due to Grain Scattering in Polycrystalline Metals,” Journal of the Acoustical Society of America, Vol. 37, p. 711 (1965). 2. E. P. Papadakis, “Ultrasonic Attenuation Caused by Scattering in Polycrystalline Media,” Physical Acoustics Principles and Methods, Vol. IV, Part B, ed., Mason, W. P., (Academic Press, New York, 1968) p. 269. 3. J. Lewandowski, “Evaluation of the Texture of Polycrystalline Aggregate from Ultrasonic Measurements,” Ultrasonics, March, p. 73 (1986). 4. C. B. Guo, P. Holler, and K. Goebbels, "Scattering of Ultrasonic Waves in Anisotropic Polycrystalline Metals," Acoustica 59, pp. 112-120 (1985). 5. D. Beecham, “Ultrasonic Scatter in Metals, Its Properties and its Application to Grain Size Determination,” Ultrasonics, April, p. 67 (1966). 6. J. Szilard, and G. Scruton, “Revealing the Grain Structure of Metals by Ultrasonics,” Ultrasonics, May, p. 114 (1973). 7. J. H. Rose, “Ultrasonic Backscatter from Microstructure,” in Review of Progress in ONDE, Vol. 11, eds., D. O. Thompson, D. E. Chimenti (Plenum, New York 1992) p. 1667. 8. K. Goebbels, “Structure Analysis by Scattered Ultrasonic Radiation,” Research Techniques in Nondestructive Testing, ed., by R. S. Sharpe, (Academic Press, New York, 1980), Vol. IV, pp. 87-157. 9. S. L. Semiatin et. al., “Hot working of titanium alloys — An overview,” Advances in Titanium Alloy Processing, eds., 1. Weiss, et. al., Proceedings from the 125" TMS Meeting held in Anaheim, California, 5-8 February, 1996, (TMS, 1997) pp. 4-26. 10. Materials Property Handbook on Titanium Alloys, eds., R. Boyer, G. Welsch, and E. W. Collings, (ASM International, 1994) pp. 483-500. 11. W. F. Hosford, "The Mechanics of Crystals and Textured Polycrystals," (Oxford University Press, New York, 1993) pp. 67-112. 12. J. F. Nye, "Physical Properties of Crystals - Their Representation by Tensors and Matrices," (Clarendon Press, Oxford, 1985) pp. 12-17. 13. J. C. Williams and E. A. Starke, Jr., "The Role of Thermomechanical Processing in Tailoring the Properties of Aluminum and Titanium Alloys," Deformation, Processing, and Structure, G. Krauss, ed. (ASM International, 1984) pp. 279-300. 14. F. J. Margetan, R. B. Thompson, and I. Yalda-Mooshabad, "Backscattered Microstructural Noise in Ultrasonic Toneburst Inspections," Journal of Nondestructive Evaluation, Vol. 13, pp. 111-136 (1994). 15. J. H. Rose, “Theory of Ultrasonic Backscatter From Multiphase Polycrystalline Solids,” in Review of Progress in ONDE, Vol. 12, eds., D. O. Thompson, D. E. Chimenti (Plenum, New York 1993) p. 1719. 16. H. Willems and K. Goebbels, "Characterization of Microstructure by Backscattered Ultrasonic Waves," Metal Science, Vol. 15, pp. 549-553 (1981). 17. C. B. Guo, P. H-ller, and K. Goebbels, "Scattering of Ultrasonic Waves in Anisotropic Polycrystalline Metals," Acoustica, 59(2), pp. 112-120 (1985). 18. K. Goebbels, "Evaluation of the Structure of Steels by Ultrasonic Scattering,” Materials Testing. 77(7), p. 231-233 (1975). 11 19. W. Hassan and P. B. Nagy, "Experimental Investigation of the Grain Noise in Interferometric Detection of Ultrasonic Waves," Journal of NDE (submitted for publication). 20. J. P. Gallagher, et. al., “USAF Damage tolerant design handbook: Guidelines for the analysis and design of damage tolerant aircraft structures,” AFWAL-TR-82-3073, pp. 1.3.1-1.6.8, (1984). 21. D. Eylon and J. A. Hall, “Fatigue behavior of beta processed titanium alloy IMI 685,” Met. Trans A, Vol. 8, pp. 981-990 (1977). 22. D. Eylon, “Fatigue crack initiation in hot isostatically pressed Ti-6Al-4V castings,” J. Met. Sci., Vol. 14, pp. 1914-1922 (1979). 23. Titanium, A technical guide, ed., M: J. Donachie, Jr., (ASM International, 1988) p. 44. 24. F. J. Margetan and R. B. Thompson, "Microstructural Noise in Titanium Alloys and Its Influence on the Detectability of Hard-Alpha Inclusions," in Review of Progress in Quantitative Nondestructive Evaluation, Vol. 11, eds., D. O. Thompson, D. E. Chimenti (Plenum Press, NY 1991) pp. 1717-1724. 25. P. J. Howard and R. S. Gilmore, "Ultrasonic C-Scan Imaging for Hard-Alpha Flaw Detection and Characterization," in Review of Progress in Quantitative Nondestructive Evaluation, Vol. 13, eds., D. O. Thompson, D. E. Chimenti (Plenum Press, New York, New York 1994) pp. 763-770. 26. J. H. Rose, “Ultrasonic Backscatter from Microstructure,” in Review of Progress in ONDE, Vol. 11, eds., D. O. Thompson, D. E. Chimenti (Plenum, New York 1992) p. 1677. 27. J. H. Rose, “Ultrasonic Backscattering from Polycrystalline Aggregates Using TimeDomain Linear Response Theory,” Review of Progress in QNDE, Vol. 11, eds. D. O. Thompson, D. E. Chimenti (Plenum, New York 1992) p. 1677. 28. R. J. McClintick, G. W. Bauer, et al., “A New Titanium Alloy,” Materials and Methods, Aug, pp. 90-92, (1955). 29. G. A. Salishchev, et. al., “Fine grained billet processing of titanium alloys,” Proceedings, gt" World Conference on Titanium, 7-11 June 1999, St. Petersburg, Russia. 30. B. P. Bewlay and M. F. X. Gigliotti, “Hard alpha phase stability in nitrided Ti-64 and Ti-17," Proceedings, 9 World Conference on Titanium, 7-11 June 1999, St. Petersburg, Russia. 31. J. C. Chesnutt and N. E. Paton, “Hold Time Effects on Fatigue Crack Propagation in Ti-6Al and Ti-6AI-4V,” in Titanium Science and Technology, ed. H. Kimura and O. Izumi (The Metallurgical Society of AIME, Kyoto Japan, 1980), pp. 1855-1863. 32. F. E. Stanke and G. S. Kino, “A Unified Theory for Elastic Wave Propagation in Polycrystalline Materials,” Journal of the Acoustical Society of America. 75(3) p. 665 (1984). 33. M. P. Blodgett, P. B. Nagy, “Anisotropic Grain Noise in Eddy Current Inspection of Noncubic Polycrystalline Metals,” Appl. Phys. Let., 72(9), pp. 1045-1047 (1998). 34. M. P. Blodgett, W. Hassan, P. B. Nagy, “Theoretical and Experimental Investigations of Lateral Resolution in Eddy Current Imaging,” Materials Evaluation, 58(5), 2000. 35. M. P. Blodgett and D. Eylon, “The influence of texture and phase distortion on ultrasonic attenuation in Ti-6AI-4V,” J. of NDE, (approved for publication). 12 36. P. D. Panetta, F. J. Margetan, I. Yalda, and R. B. Thompson, "Observation and Interpretation of Microstructurally Induced Fluctuations of Back-surface Signals and Ultrasonic Attenuation in Titanium Alloys," Review of Progress in Quantitative Nondestructive Evaluation, Vol. 16, Eds., D. O. Thompson and D. E. Chimenti, (Plenum Press, NY, 1997) pp. 1547-1554. 37. T. Seldis and C. Pecorari, "Scattering-induced Attenuation of an Ultrasonic Beam in Austenitic Steel," Journal of the Acoustical Society of America (submitted for publication). 38. M. R. Hollard and J. G. Miller, "Phase-insensitive and Phase-sensitive Quantitative Imaging of Scattered Ultrasound Using a Two-dimensional Pseudo-array," Ultrasonics Symp. Proc. IEEE Cat. No. 88CH2578-3, pp. 815-819 (1988). 13 CHAPTER If LITERATURE REVIEW This research draws heavily upon the works of previous researchers. In fact, there exists an enormous wealth of information in archival journals and conference proceedings dealing with subject matter similar to this research. Not surprisingly, the search for "new ground" in the subject area has required a substantial investment of time and effort in reviewing the literature. Much of the documented literature shows similarities to activities undertaken during this project. This is especially true in Chapter 3, which reviews the conventional ultrasonic measurements, including longitudinal and shear wave velocities, attenuation, and backscattering. While similarity to previous works clearly exists in Chapter 3 and 4, Chapters 5 and 6 are completely original and based on refereed publications, resulting from this dissertation. The results from Chapter 4 are based on the well known ultrasonic backscattering approach. All of the samples developed are original and unique to this dissertation. To demonstrate originality, this research has yielded the publication of three refereed journal papers, with at least one more to follow. To the best of my knowledge, this work is not duplicated nor examined as such in any of the open literature. 14 2.1 Fundamentals on Ultrasonic Scattering The fundamental scattering problems have been around since the days of Lord Rayleigh [1] in the study of sound wave interaction with discrete inhomogeneities, which laid the groundwork for developing the single-source scattering theory. This is an important starting point and it is well understood that when an oscillating wave traveling in a homogeneous medium like pure water impinges on an inhomogeniety, the response of the wave to the obstruction will depend on certain physical and mechanical aspects of the scatterer and propagating medium. The disturbance presented by the inhomogeniety will result in scatter that will inside characteristics propagation depend on the the medium direction and the of the incident wave, the on the and inhomogeniety, boundary conditions across the interface of the scatterer, which must satisfy continuity of stress and displacement across the boundary. While scattering from a discrete, single source of a certain characteristic geometric shape is an important fundamental acoustical problem, the main issues for the current research stem from the consideration of a material whose scatterers. Here, incoherent scattering arises from the elastic volume anisotropy is full of and phase heterogeneity of the generally microscopic single crystals which comprise the material. An abundance of analytical work has been conducted in the recent past to characterize the grain-to-grain variations of elastic stiffness, relative to the average material stiffness, for untextured single phase materials with cubic symmetry [2, 3, 4]. Past research has also delivered expressions for the effective wave speed and attenuation in what are now called the Rayleigh and Stochastic scattering regimes, where the wavelength is much larger and 15 comparable size of the to the nominal scatterer, respectively. work Subsequent by Huntington [5] in the Diffusive (or geometric) scattering regime, where the wavelength is much smaller than the scatterer, completed the fundamental theoretical development of the problem. Microstructural disturbances generate ultrasonic scattering and mode conversion at discontinuities associated with changes in phase, density, or stiffness because of slight differences sound in acoustic travels in a impedance velocity and acoustic given direction across the boundary. When there is a a polycrystalline through material, discontinuity in the wave speed at each grain boundary, due in the simplest case, to local elastic property new each With differences. unique , entity crystallographically encountered by the propagating wave, scattering takes place at the boundaries when the mean linear wavelength. dimension, D, of the grains is comparable with, or smaller than the The scattering represents a loss of energy to the propagating wavefront, generally providing the main source of attenuation in metals. In the case of grain sizes of 1/1000th to 1/100th of the wavelength, scatter is considered negligible [6]. However, when the dimensions of the scatterer are 1/10th to the full value of the wavelength the scattering increases very rapidly, approximately as the third power of the grain size. Three different regimes for scattering of ultrasonic waves can depending on the ratio of the mean grain size to the wavelength (i.e. D /): be specified @) the low frequency (Rayleigh) region with scattering induced attenuation proportional to the fourth power of the frequency and to the cube of the mean grain diameter, (ii) the medium 16 frequency (stochastic) region with attenuation proportional to the square of frequency and to the mean grain diameter, and (iii) the high frequency region (geometric) with attenuation independent of frequency. For the geometric scatter case, conditions suitable for testing may no longer exist, due to the enormous degree of scatter, especially if the test material is anisotropic. Other theoretical studies have been conducted to model more complicated two phase material systems [7, 8]. A simplified approach was used to model the material as if it were composed of spherical scatterers in a matrix with different density and stiffness between the two phases. This simple model cannot however, predict the attenuation and velocity in complicated structural alloys of titanium. Titanium microstructures, owing to their elastic anisotropy, the nature of the slip, and the effects of deformation, are complex and wave propagation and scattering cannot be described in simple mathematical terms. This complication is normal processing, for materials subjected to thermomechanical which often result in the alignment of the second phase particles and grain boundaries. In addition, preferred crystallographic orientation adds another of level analytical complexity due to the directionally dependent macroscopic elastic properties. Multiple scattering is an important subject that has been investigated in the recent past [9, 10]. Single scattering theories assume there is no interaction between scattered signals. scattering However, is of major titanium alloys, types of materials including importance as the neglect of this phenomenon in many multiple can result in significantly misleading interpretations of the materials ultrasonic characteristics. The 17 microstructure dependent ultrasonic attenuation of a material can be generally resolved by isolating the energy carried off by scattering, and correlating this fractional quantity of the incident power to grain structure. This correlation works nicely when the assumption of single scattering holds (as for an equiaxed media), but tends to breakdown more becomes microstructure scatter becomes Multiple complex. a more as the significant problem in highly scattering materials. In multiple scattering materials, the influence of on ultrasonic microstructure energy attenuation is generally overestimated scattered again and again. Moreover, is subsequently as the scattered scattering can multiple interfere with determining the origination of the scatter source inside the material, which can lead to anomalous defect indications where in fact there is no defect of significance. The grain scattering problem occurs with many different types of structural alloys, due to the natural solidification process and other changes that take place in the solid as a result of fabrication and heat treatment processes. Papadakis [11, 12, 13] has conducted research in developing extensive applied to different types of ultrasonic the understanding media. of inhomogeneous His work grain scattering as concentrated on understanding the frequency dependence of attenuation, the influence of microstructure on velocity and attenuation, and on developing a correlation between the ultrasonic response and grain size in cubic, monophase, polycrystalline materials having equiaxed grains with a narrow distribution of grain sizes and shapes. This work also comprehensively reviews related research going back to the late 1950’s. Equiaxed single phase aluminum and steel alloys have been the subject of a great deal of research in the ultrasonic grain scattering arena. These materials are nicely suited for studies on the 18 effect of grain size on ultrasonic properties, due to the ease of generating samples with uniform, narrowly distributed grain sizes. Many made researchers [14, 15, 16, 17, 18] have long lasting contributions to this field, having concentrated on modeling the interaction of the ultrasonic waves with the solid. Others [19, 20, 21] have taken the more practical approach by empirically correlating ultrasonic measurements with microstructural features like grain size. 2.2 Wave Propagation in Inhomogeneous Materials The literature on wave propagation in inhomogeneous or heterogeneous media is extensive. Composite materials are considered heterogeneous and wave propagation is dramatically influenced by the structure of the material. Many general studies concerning ultrasonic propagation in heterogeneous material have been made in the recent past (22, 23]. Wave propagation and scattering theory in weakly anisotropic, equiaxed polycrystals in the absence of preferred orientations has been extensively studied [24, 25] for both single and multiphase materials. A unified approach to the solution to the wave equation, based on ultrasonically detected variations of elastic properties, has been developed [26]. The influence of texture and elastic anisotropy on wave propagation in polycrystalline metals has been also the subject of extensive research [27, 28, 29, 30]. Here, the theoretical work has concentrated on allowing the prediction of scattering coefficients, longitudinal and shear wave velocities, and attenuation as a function of frequency in textured polycrystals. However, these models only apply to cubic metals. Much of the work on theoretical wave 19 propagation is mathematically intensive and goes well beyond the scope of this research. The wave propagation in polycrystalline metals involves solving the wave equation, taking into account for calculations scattering, directional dependencies, frequency variations, elastic properties, different types of waves, mode conversion and so on. It is important to realize that much work has been conducted in this research subject; however, most of these studies are based on assumptions and criteria that are simply not supported for titanium alloys. For example, the fact that the single crystal elastic constants for Ti-6AI-4V do not exist in a rigorously determined and reproducible manner makes it difficult to apply any of these extensively developed theoretical models to the material. Moreover, wave propagation in Ti-6AI-4V is complicated its by hexagonal symmetry and differences in density, elastic constants, and lattice structure of its two phase microstructure. Likewise, developing an understanding of the influence of microstructure on wave propagation in the absence of a preferred orientation is very difficult, due to the relative ease in imparting texture and the difficulty in getting rid of it without otherwise affecting the microstructure. 2.3 Ultrasonic Backscattering for Structural Characterization The occurrence of grain structure in metal alloys gives rise to randomly occurring ultrasonic signals, which can be in some Microscopically homogeneous but cases easily mistaken randomly oriented individual for genuine grains make defects. up a macroscopically isotropic but inhomogeneous medium which produces incoherent wave scattering commonly referred to as "grain noise." This ultrasonic scatter generated in structural alloys is basically the result of local variations of either stiffness or density in the material. Features which promote scatter are both microscopic, due to grain structure, second phase particles, precipitates, crystal defects, microporosity, and microcracks, and macroscopic, due to the alignment of second phase particles or grain boundaries, the formation of large colonies of similarly oriented grains, and deformation flow lines. In Ti-6Al-4V, the microstructural complexity is such that the ultrasonic scatter originates from grain-to-grain differences in crystallographic orientation, density, elastic properties, the formation of macroscopic structure; and the formation of features originating from a prior phase state (i.e. prior beta grain boundaries). In light of the multitude of scattering sources and microstructural variations, titanium alloys represent a truly challenging class of materials for ultrasonic characterization. Acoustic backscattering. ultrasonic flaw observed in many materials using ultrasonic grain noise has an obvious adverse, often prohibitive, effect on and it can be grain noise The detection is readily [31, 32, 33, 34] exploited for ultrasonic characterization of the grain structure (35, 36]. A major source for the generation of ultrasonic scatter is grain boundaries. The factors primarily responsible for the ultrasonic scatter in most titanium alloys are the alignment of grains and grain boundaries, second phase particles, and grain colonies. Changing the ultrasonic inspection parameters, such as reducing the frequency or increasing the transducer aperture, causes a reduction in the generation of grain noise, but these tradeoffs often reduce the detectability threshold of the inspection. Likewise, a variety of different averaging techniques can be employed to reduce grain noise and improve the microstructure discrimination capability, but these techniques are not favorable for detection of localized defects [37, 38]. 2.4 Diffusion Bonding of Titanium Alloys While the Ti-6Al-4V alloy not is ideal for analytical microstructure characterization from the standpoint of its many microstructural features responsible for scattering and its diverse array of useful microstructure types, the alloy does have some uniquely attractive properties. One of the most useful aspects of this alloy, aside from its extensive processability, is its joining capability given the right combination of pressure, time, and temperature [39, 40]. A beneficial feature of Ti-6AI-4V is that it is capable of being welded and diffusion bonded. In diffusion bonding, the mating parts are heated to about one half of the melting point, then pressurized to a stress level below the macroscopic yield point, and held at temperature and pressure for a given time period. The procedure commonly practiced, and used for development of bonded samples for this research, was conducted in a HIP (Hot Isostatic Pressure) chamber with temperature, pressure, and dwell time conditions of 950 °C, 30 ksi, for4 hours, respectively. The diffusion bonding process involves three basic steps: (i) local yielding of initially contacting surface asperities, (ii) creep deformation of the bond plane yielding discontinuous voids, and (iii) closure of the voids by vacancy diffusion [41]. In practical situations, oxide free surfaces are rare. Oxidized surfaces generally interfere with the bonding process due to impedance of the rate and extent of diffusion across the bond plane. This is generally not the case for titanium alloys, which at temperatures above 850 °C readily dissolve minor amounts of absorbed gases and thin surface oxides, allowing 22 diffusion to carry them away from the bond interface. Another aspect of diffusion bonding unique to this research is the joining of dissimilar microstructures of the same chemical composition. While many researchers have investigated the bonding and characterization of dissimilar metal compositions, it is evident from the limited number of publications found that bonding of dissimilar microstructures has been given very little attention in the literature [42]. A great deal of research has been conducted to better understand how to nondestructively assess diffusion bond characteristics and to address the determination of processing parameters that provide optimal bond strength (43, 44, 45, 46]. Here, considerable effort has gone into the development of theories to predict the relationship between ultrasonic scattering from diffusion bonds and the bond conditions, based on experimental NDE measurements. 23 Chapter 2, References: 1. L. Rayleigh, “The Theory of Sound”, (Macmillan, 1945) pp. 147-152. 2. I. M. Lifshits and G. D. Parkhomoskii, “On the Theory of the Propagation of Supersonic Waves in Polycrystals,” Zhur-Experim. I. Theoret. Fiz., Vol. 20 (1950). 3. A. B. Bhatia, “Scattering of High Frequency Sound Waves in Polycrystalline Materials,” Journal of the Acoustical Society of America, Vol. 31, pp. 16-23 (1959). 4. A. B. Bhatia and R. A. Moore, “Scattering of High Frequency Sound Waves in Polycrystalline Materials II,” Journal of the Acoustical Society of America, Vol. 31, pp. 1140-1144 (1959). 5. H. B. Huntington, “On Ultrasonic Scattering by Polycrystals,” Journal of the Acoustical Society of America, Vol. 22, p. 362 (1950). 6. Ultrasonic Testing of Materials, eds., J. Krautkramer and H. Krautkramer, (SpringerVerlag, NY, 1990), pp. 425-430. 7. C. F. Ying and R. Truell, “Scattering of a Plane Longitudinal Wave by a Spherical Obstacle in an Isotropically Elastic Solid,” Journal of Applied Physics, Vol. 27, pp.10861097 (1956). 8. P. C. Waterman and R. Trvell, “Multiple Scattering of Waves,” Journal of Mathematical Physics, Vol. 2, pp. 512-537 (1961). 9. Ultrasonic Methods in Solid State Physics, eds., R. Truell, C. Elbaum, and B. B. Chick, (Academic Press, New York, 1969). 10. R. H. Latiff, and N. F. Fiore, “Ultrasonic Attenuation and Velocity in Two-Phase Microstructures,” Journal of the Acoustical Society of America, Vol. 57, pp.1441-1447 (1975). 11. E. P. Papadakis, “Ultrasonic Attenuation due to Grain Scattering in Polycrystalline Metals,” Journal of the Acoustical Society of America, Vol. 37, p. 711 (1965). 12. E. P. Papadakis, “Ultrasonic Attenuation Caused by Scattering in Polycrystalline Media,” Physical Acoustics Principles and Methods, Vol. IV, Part B, Ed. Mason, W. P., (Academic Press, New York, 1968) p. 269. 13. E. P. Papadakis, “Influence of Preferred Orientation on Ultrasonic Grain Scattering,” Journal of Applied Physics, 36(5), May, p. 1738 (1965). 14. F. J. Margetan, R. B. Thompson, and I. Yalda-Mooshabad “Modeling Ultrasonic Microstructural Noise in Titanium Alloys,” Review of Progress in Quantitative Nondestructive Evaluation, Vol. 12, eds., D. O. Thompson and D. E. Chimenti, (Plenum, New York, 1993), p. 1753. 15. B. Fay, “Theoretical Considerations of Ultrasonic Backscatter,” Acoustica 28, p. 354 (1973). 16. J. Lewandowski, “Evaluation of the Texture of Polycrystalline Aggregate from Ultrasonic Measurements,” Ultrasonics, March, p. 73, (1986). 17. C. M. Sayers, “Ultrasonic Velocities in Anisotropic Polycrystalline Aggregates,” Journal of Applied Physics D, 15, p. 2157, (1982). 18. C. B. Guo, P. Holler, and K. Goebbels, "Scattering of Ultrasonic Waves in Anisotropic Polycrystalline Metals," Acoustica 59, pp. 112-120 (1985). 19. D. Beecham, “Ultrasonic Scatter in Metals, Its Properties and its Application to Grain Size Determination,” Ultrasonics, April, p. 67 (1966). 20. J. Szilard, and G. Scruton, “Revealing the Grain Structure of Metals by Ultrasonics,” Ultrasonics, May, p. 114 (1973). 24 21. R. L. Smith, “The Effect of Grain Size Distribution on the Frequency Dependence of the Ultrasonic Attenuation in Polycrystalline Materials,” Ultrasonics, September, p. 211 (1982). 22. J. M. Perdigao, A. Ferreira, and C. Bruneel, “Ultrasonic Anomalous Behavior in Composite Samples,” Acoustica, Vol. 63, pp. 106-110 (1987). 23. D. Sornette, “Acoustic Waves in Random Media, Experimental Situations,” Acoustica, Vol. 68, pp. 15-25 (1989). 24. §. Hirsekorn, “The Scattering of Waves by Polycrystals,” Journal of the Acoustical Society of America, 72(3) (1982). 25. S. Hirsekorn, “The Scattering of Ultrasonic Waves by Multiphase Journal of the Acoustical Society of America, 83(4) (1988). Polycrystals,” 26. F. E. Stanke and G. S. Kino, “A Unified Theory for Elastic Wave Propagation in Polycrystalline Materials,” Journal of the Acoustical Society of America. 75(3) p. 665 (1984). 27. S. Hirsekorn, “The Scattering of Ultrasonic Waves in Polycrystalline Materials with Texture,” Journal of the Acoustical Society of America, 77(3), (1985). 28. S. Hirsekorn, “The Dependence of Ultrasonic Propagation in Textured Polycrystals,” Journal of the Acoustical Society of America, 79(5), (1986). 29. C. M. Sayers, “Angular Dependent Ultrasonic Wave Velocities in Aggregates of Hexagonal Crystals,” Ultrasonics, Vol. 24, pp. 289-291 (1986). 30. S. Ahmed and R. B. Thompson, “Propagation of Elastic Waves in Equiaxed Stainless-Steel Polycrystals with Aligned [001] Axes,” Journal of the Acoustical Society of America, 99(4), (1996). 31. J. H. Rose, “Ultrasonic Backscattering from Polycrystalline Aggregates Using TimeDomain Linear Response Theory,” Review of Progress in QNDE, Vol. 11, eds D. O. Thompson, D. E. Chimenti (Plenum, New York 1992) p. 1677. 32. J. H. Rose, “Ultrasonic Backscatter from Microstructure,” in Review of Progress in ONDE, Vol. 11, eds D. O. Thompson, D. E. Chimenti (Plenum, New York 1992) p. 1677. 33. J. H. Rose, “Theory of Ultrasonic Backscatter From Multiphase Polycrystalline Solids,” in Review of Progress in QNDE, Vol. 12, eds D. O. Thompson, D. E. Chimenti (Plenum, New York 1993) p. 1719. 34. F. J. Margetan, R. B. Thompson, and I. Yalda-Mooshabad, “Backscattered Microstructural Noise in Ultrasonic Toneburst Inspections," Journal of Nondestructive Evaluation, Vol. 13, 111-136 (1994). 35. K. Gobbles, “Structure Analysis by Scattered Ultrasonic Radiation,” Research Techniques in Nondestructive Testing, Vol. IV, edited by R. S. Sharpe, (Academic Press, New York, 1980), pp. 87-157. 36. A. Hecht, R. Thiel, E. Neumann, and E. Mundry, "Nondestructive Determination of Grain Size in Austenitic Sheet by Ultrasonic Backscattering," Materials Evaluation, Vol. 39, pp. 934-938 (1981). 37. H. Willems and K. Goebbels, "Characterization of Microstructure by Backscattered Ultrasonic Waves," Metal Science, Vol. 15, pp. 549-553 (1981). 38. Y. K. Han and R. B. Thompson, “Ultrasonic Backscattering in Duplex Microstructures: Theory and Application to Titanium Materials Transactions A, Vol. 28A pp.91-103 (1997). 25 Alloys,” Metallurgical and 39. H. G. Kellerer, and L. H. Milacek, “Determination of Optimal Diffusion Welding Temperatures for Ti-6Al-4V,” Welding Research Supplement, 49(5), pp. 219s-224s, (1970). : 40. R. J. Rehder, and D. T. Lovell, “Process Development for Diffusion Welding Ti-6Al4V Alloy,” Welding Research Supplement, 49(5), pp. 213s-218s, (1970). 41. C. H. Hamilton, “Pressure Requirements for Diffusion Bonding of Titanium,” in Titanium Science and Technology, eds., R. I. Jaffee and H. M. Burte (Plenum Press, 1972), pp. 625-647. 42. A. A. Gelman, A. A. Kotelnikov, and N. I. Kolodkin, “The Effect of Diffusion Bonding Variables on Structure and Properties of the Bonded Alpha + Beta Titanium Alloy,” in Titanium and Titanium Alloys, Eds. J. C. Williams and A. F. Belov, (Plenum Press, 1976), 1209-1220. 43. D. D. Palmer, C. D. Roberts, et al., “Strength and Ultrasonic Characterization of Metallic Interfaces,” in Review of Progress in Quantitative Nondestructive Evaluation, Vol 7, Eds. D. O. Thompson and D. E. Chimenti, (Plenum Press, 1988), pp. 1335-1342. 44. J. H. Rose, “Ultrasonic Characterization of Solid-Solid Bonds from Microstructural Changes,” in Review of Progress in Quantitative Nondestructive Evaluation, Vol. 7, Eds. D. O. Thompson and D. E. Chimenti, (Plenum Press, 1988) pp. 1311-1318. 45. G. T. Thomas and J. R. Spingarn, “Ultrasonic Determination of Diffusion Bond Strength,” in Review of Progress in Quantitative Nondestructive Evaluation, Vol. 3, Eds. D. O. Thompson and D. E. Chimenti, (Plenum Press, 1984) pp. 1243-1250. 46. B. H. Hosten, L. A. Ahlberg, B. R. Tittman, and J. R. Spingarn, Ultrasonic Characterization of Diffusion Bonds,” in Review of Progress in Quantitative Nondestructive Evaluation Vol. 6, Eds. D. O. Thompson and D. E. Chimenti, (Plenum Press, 1986) pp. 1701-1706. 26 CHAPTER III MICROSTRUCTURAL EFFECTS ON ULTRASONIC PROPERTIES 3.1 Microstructures Evaluated One of this research was influence the to determine of of the main goals on the ultrasonic response in investigated, including: (i) the initial as-received mill annealed microstructure microstructures were material, (AR); (ii) coarse beta annealed, (CBA); and (RA); (v) duplex Hence, Ti-6Al-4V. annealed, (DA) annealed, Microstructures Evaluated). microstructures of interest since the bar is highly anisotropic. as-received different (iv) (ili) fine beta annealed, (FBA); recrystallization The five material really As (see Table has two 3.1 — different such, each of the different microstructures were examined in both the axial and transverse orientations. However, only the mill annealed Ti-6A1-4V clearly showed a distinction between the two orientations studied, as shown in Figure 3.l.a and 3.1.b. The commonly produced in the mill annealed condition, where it has strength, toughness, ductility, and fatigue properties, Ti-6Al-4V alloy is a good combination of but the properties are strongly orientation dependent [1, 2]. Approximately 2 meters of 2.5” (63.6 mm) diameter forged bar stock was acquired as a basis for the different microstructure samples, which were sectioned, heat treated, and polished for flat and parallel surfaces. 27 87 ‘SOINJOMASOIOIU UOTIPPe OJ pue ‘UOTIPUOD poataool-se oY} JO SUOTIEIUSLIO a[dryNUI Jo pajstsuod YOrYM porprys OE-ST Or-0€ 06 ‘09 ‘Sb “LE “OE “0 a Te e198L UleIS [BUIWION (ur) 9218 pure poyersue8 osom sopdueg ‘sydesSoso1w Surpnyjout ‘Apmis WoNeyossip Sty} SULINP pojenyead SoMjon.ysOIONW sy YL, 00S-00€ 06 0 bes a 000€-0007 , SUOTJCJUSTIO _, | (seeudap) POTPTS . 06 0 soiniee fe « 06 0 saluojoo , , Jo suorSau Je]/auey urexd ‘sures Jo —ures3 pu ‘saepunoq , 06 0 q+ Jo Sunsisuoo ‘ejaq]} 3 pouojsuey paynqlsysip Joiid yo aouapiaa | Jord Jo souaplAa yuOTUeO yeOHT SulYysInsurjsiq nyeoj sured eydje pexembs| Sulels 8}0q Bjoq IBjNUEIsIO}U! surels B}9q eydye poxeinbs surly OV ‘OD 0€2 “Ub dINJON.YSOISTYA SITYepOUSUION, dv (PSAT999I-SB) poereouue [TIA yuowustye jerxe Suoys DAROMO¥sSeq Jeg] pue satucjoo | = — SWOs ‘anjon43s uieId JEppaweE] 93Ie] Smid ‘OW ayIsusTeU SAOge JOY ‘OD geo “AY | “dutay els OV ‘D POL “84 DUISUSLEU MOfeq JeaY ‘9.006 TY p ‘dus eis | ov ‘ovo. | ov ‘DO rol UN zsnid SOV] sy zsnd oa} OV ‘DO H0L ‘> 000154 S‘0 vd peyTeouue xojdnq poyenjeay SompoNQSOIJAl UOZT[VISAINIY Va poyeouue ‘ay Z snd ‘oq Vdd peTeouue ejoq Our] “> LEOLIY 9 Vado peyeouue BIJ9q BSIBO7) Figure 3.1 The as-received condition (mill annealed) of Ti-6Al-4V bar stock in the transverse (a) and axial (b) orientation. Both microstructures shown at 200X. Thermomechanical breakdown of the original ingot is generally done via cogging, upsetting, or (for smaller sections) extrusion [3]. In cogging, the initially round ingot is pressed on its sides between flat dies and incrementally rotated and indexed forward, resulting in a long, relatively thin (billet) product. Upsetting cogging to introduce additional strain energy to promote is often a precursor to recrystallization upon hot working in the a +B phase field where the self diffusion coefficient is quite low (e.g. at 500 °C, Dar © 107!9m? /s). As a result, the grain structure remains fine upon cooling, with nominally 15-20 pm primary a grains [4] and approximately 5-7% Vr of intergranular B in either orientation. The mill annealed microstructure in the transverse orientation takes on a distinctive "wavy" lamellar appearance sometimes called a partially transformed widmanstiatten structure. In the axial orientation, the primary alpha grains of the mill annealed structure have nominally a 5:1 aspect ratio and large colonies of similarly oriented alpha grains tend to agglomerate into distinctive bands, which are thought to behave essentially like large single crystals. The mill anneal conditions the 29 alloy into a relatively soft (Rc ~ 34), easily machinable state. Typical mill annealing involves heating the alloy to approximately 730 °C (in the a + B phase field) where it is held for approximately 4 hours, followed by air cooling to room temperature. This heat treatment is generally given to all titanium mill products and often the annealing is intentionally left incomplete, leaving telltale traces of the heavily worked structure as seen in Figure 3.1b. This conditioning leaves the billet in a highly anisotropic state. The duplex and recrystallization annealed microstructures were also investigated. It is important to point out that neither of these samples demonstrated any significant recognizable difference in microstructural features between the axial and transverse orientations, except at low resolution (~25X) optical microscopy images, which reveals a banded structure similar to that pbserved in the mill annealed condition. The duplex microstructure results from heat treating the mill annealed alloy above the martensite start temperature. Different variants of the duplex anneal process are available to vary the relative amounts of « versus B, but the process chosen for this work involved heating the alloy to 938 °C for 1 hour, followed by air cooling, then reheating to 704 °C for 2 hours and air cooling to room temperature. The duplex microstructure dramatically transforms the mill anneal to appear as is shown in Figure 3.2.a. This microstructure is comprised of equiaxed a grains of nominally 30-40 pm with distributed transformed 8 appearing as fine lamellar regions throughout the structure. The annealing temperature determines the amount of a and transformed B according to the phase diagram lever rule. During annealing, the primary o phase retains the morphology of the originating structure from thermomechanical processing. On the other hand, the transformation of the beta phase 30 depends on the cooling rate, whereas slower cooling results in the commonly observed lamellar packets of secondary alpha, shown as the dark regions. The recrystallization annealed (RA) microstructure is also result of heating the mill annealed material below the martensite start temperature, leading to the evolution of fine equiaxed a and B grains as shown in Figure 3.2.b. This treatment involves heating the alloy to 900 °C for 4 hours followed by air cooling, then reheating to 704 °C for 2 hours and air cooling to room temperature. The microstructure consists of equiaxed primary o grains of approximately 25-30 jm and intergranular 8. The recrystallization annealed condition results in high fracture toughness, good hot formability, and good fatigue initiation resistance. Beta annealing relies on slow cooling from above the beta transus where the alloy is transformed to a cubic lattice structure. As the hot worked alloy is cooled from the B phase region, « crystallizes in the form of plates with a known crystallographic regularity ‘in which the o plates form with their basal planes parallel to the {110} planes in the B phase. Similarly, the body diagonal in the B phase forms parallel to the basal plane diagonal in the a phase. These orientation rules are relationship [5, 6], expressed in terms of Miller indices as: {110}, //(0001), and (111), / (1120) consistent with the Burgers Eso yc + eb ete!¢ Bee aura Figure 3.2 Above Gora a Se ub Bhd t. (a) and_ the anneal the duplex resulting from Microstructures recrystallization anneal (b). Both microstructures shown at 200X and both have nominally the same primary « grain size of approximately 30 jm. the beta transus (~1000 °C) each B grain has six possible combinations of independent {110} planes on which a plates can form, resulting in the now commonly known Widmanstatten microstructure. This microstructure consists of a plates separated by thin films of the B phase (approximately 10-15 ym thick), which form with a known crystallographic relationship to the prior B phase from which they originated. The fine beta annealed microstructure examined in this work is the result from heating the alloy to 1000 °C for 0.5 hours followed by furnace cooling, then reheating to 704 °C for 2 hours and air cooling to room temperature, resulting the microstructure as shown in Figure 3.3.a. This heat treatment drastically changes the appearance of the microstructure from the highly anisotropic mill annealed condition. The banded structure evident at low magnifications in the AR, DA, and RA samples is almost completely removed, leaving a structure with a relatively small prior beta grain size of 300 to 500 um, but broad lamellae because of the slow furnace cooling rate (~100 °C/h) 32 from approximately 1000 °C. The relatively slow (air) cooling rate results in diffusion controlled partitioning between the « and B phases as the temperature falls below the B transus. The microstructure is lamellar and the size of the microcolonies depends on the prior beta grain size and on the cooling rate. To reiterate, for this FBA microstructure the transverse orientation is virtually indistinguishable from the axial orientation. The coarse beta annealed microstructure is heated to 1037 °C for 6 hours followed by furnace cooling, then reheated to 704 °C for 2 hours and air cooled, as shown in Figure 3.3.b. The prior beta grain boundaries are clearly evident in the coarse beta annealed microstructure. The prior beta grain size is determined by the dwell time above the beta transus and the morphology of the « /B combination is strongly dependent on the cooling rate. Quenching from above the beta transus results in a martensitic transformation where the beta phase becomes needlelike. Slower cooling allows diffusion between the phases, resulting in a lamellar structure with broad o plates alternating with fine B phase. This lamellar structure partitions into colonies of similar orientation. The size of the colonies depends on the prior beta grain size and the cooling rate. Another interesting feature, often occurring in alloys that have been cooled from boundary « (GBa), which is clearly evident in the CBA above the B transus is grainmicrographs of Figure 3.3.b. GBa nucleates at prior B grain boundaries and its thickness and continuity depends on the cooling rate and on alloy composition. Again, no difference was observed in the appearance between the two CBA orientations studied. However, a huge difference is seen between the fine versus coarse beta annealed samples in terms of both the prior beta grain size and microcolony size. (b) Figure 3.3 The beta annealed microstructures: (a) fine beta annealed structure shown at 200X, exhibiting approximately 300-500 1m prior beta grain size, and (b) coarse beta annealed structure shown at 25X, with extremely large approximately 2-3 mm prior beta grain size. PART I: Conventional Ultrasonic Evaluation Methods 3.2 Ultrasonic Velocity This section centers on an assessment of the influence of texture, orientation, and wave microstructure on ultrasonic microstructures. For fundamental in propagation understanding of the several common interaction of titanium propagating ultrasonic waves within a titanium solid, many reliable techniques have been developed in the past for measurement of ultrasonic velocity, attenuation, and backscattering characteristics. While these fundamental techniques are not new, they provide a baseline capability to evaluate materials and provide useful information on the elastic properties and microstructure characteristics. This section relies on NDE experimentation to study different structures, especially since theoretical models have not been developed which are capable of predicting the complex interactions that take place as an ultrasonic wave passes through a complex material. Moreover, the theoretical understanding of elastic 34 wave propagation in an inhomogeneous anisotropic solid is mathematically very intensive this research. and goes beyond an objective of this section is to develop The main understanding of the basic ultrasonic response characteristics of different microstructures of the titanium alloy and to determine what parameters primarily influence the behavior waves. ultrasonic of propagating into three basic is broken section The categories: velocity, attenuation, and backscattering. The longitudinal velocity of sound, including ultrasound, depends on the elasticity and density propagation of the The medium. where compressional wave of a small amplitude is c= JE/p, and p_ is the density. For fluids, E of a for velocity expression general E is the elastic modulus is the bulk modulus of elasticity, but is equal to Young's modulus for the case of solids. Crystallographic texture and anisotropy integrally linked to the elastic response ultrasonic velocity due to directional are of the material and significantly influence [7, 8, 9]. The elastic property dependencies variations in polycrystalline solids are evident from the ultrasonic velocity and depend fundamentally on a number of different factors including: (i) the single crystal elastic properties of the metal, (ii) the crystallographic structure, and (iii) the effects (if any) of deformation and processing, which may bring about residual stresses or crystallographic texture. Acquiring single crystal elastic properties of metals has been a subject of interest to many researchers [10, frequencies to the sample algorithms to deduce the 11]. Resonance techniques introduce a broad range of and record the response which feeds different computer independent elastic moduli of anisotropic single crystal materials [12, 13, 14]. However, fundamental technical challenges still inhibit our ability to experimentally determine the independent elastic constants of Ti-6Al-4V. Accurate velocity measurements require sophisticated hardware and the use of a as pulse-overlap excited via tone-burst, which results in an reliable technique such accuracy of +0.5%. Also, samples are required to be flat and parallel with a sample diameter to transducer aperture ratio of approximately 3:1, to minimize errors introduced via sidewall interference. An extensive effort has been taken to satisfy the sample requirements, consistent with guidelines specified in the literature for the practice of measuring ultrasonic velocity via the pulse overlap technique [15, 16]. All samples were measured in both the axial and transverse directions to assess the effects of texture, orientation, and microstructure on the velocity. The nomenclature used to describe the samples is shown in Table 3.2 -Sample Nomenclature, Appendix A. Anisotropy can be observed through a variety of phenomena such as orientation dependent acoustic velocity changes and birefringence. Experimentation involving orientation dependence of longitudinal and shear wave propagation provides a method to assess texture and microstructural influence in bulk materials. Birefringence means the refraction of the ultrasonic shear wave into two waves of slightly different velocity. Figure 3.4 shows the geometrical configuration for the birefringence measurement. At the pure mode polarization directions both fast and slow modes can simultaneously travel at two separate, slightly different times. However, the phase interference resulting from the difference in arrival times often causes significant amplitude reductions in the received radial : ‘=> “~ AR(90) Figure 3.4 axial polarization The geometric configuration of the birefringence measurement different polarization and wave propagation directions. with signal in-between the pure mode directions. Velocity measurements have revealed only minor variations due to microstructure, except in the coarse beta annealed sample. However, some dramatic velocity variations results were observed due to the presence of a preferred crystallographic orientation, shows the average stock. Figure 3.5 velocities (see Table 3.3 — Longitudinal Wave Velocity, Appendix especially in the as-received bar longitudinal A). To minimize experimental error, the thickness and time delays between signals on each sample was measured several times and averaged. Thickness calibration blocks were also used to make certain the delay times were measured accurately. The data clearly shows the average longitudinal wave velocity is highest in the transverse direction of the mill annealed sample, AR(90). A difference of about 2% is observed in the wave velocity between the axial versus the transverse direction of the as- received bar stock. The longitudinal data also shows a linear decrease in velocity as a 37 6300 7 6050 +t | 90 degree L] 0 degree 6200 -- _ — [| 6150 ++ 6100 4 6050 | + cog JM yt tit AR(30)ttt AR(0) DA(0) DA(90) RA(O) RA(90) FBA(0)FBA(90)CBA(0) CBA(90) AR(90) AR(60) AR(45) AR(37) sample Figure 3.5 Results showing the variation in average longitudinal wave velocity for the various titanium alloy samples. The scatter on these measurements is generally about + 0.5%. function of the angle between the surface normal and the axial direction in the six asreceived samples. Special samples were developed to examine the influence of macroscopic texture on the ultrasonic response. These as-received samples were cut such that the surface normals were at known angles (0, 30, 37, 45, 60, and 90 degrees) relative to the axis of the mill annealed bar stock, shown in Figure 3.6. A steady reduction in velocity of about 22 m/s is observed as the surface normal changes from transverse (90°) to parallel (0°), relative to the bar axis. Beyond this uniform reduction in velocity as the incidence angle changes in the as-received samples, we see measurable variations in velocity between the other the axial and transverse directions for samples AR, DA, and RA, all showing approximately a 2% difference. Samples FBA and CBA also show this 38 - ay Lo) ON rom) o ‘© Cnt: ee it ~ ~, Figure 3.6 Diagram showing the different Ti-6Al-6V sample sections cut from the 2.5" diameter cylindrical bar (i.e., five samples having varying surface normals relative to bar axis). directional dependence effect, but to a much lesser degree. The reason for this behavior is because of recrystallization, which removes the texture, based on the dwell time above the beta transus. Sample DA(0) has the lowest velocity, closely followed by samples RA(0) and AR(0) with FBA(0) and CBA(0) having slightly higher velocity then the rest in the axial group. The data shows the influence of microstructure has a measurable, but modest, effect on the longitudinal acoustic wave velocity. What is evidently much more influential on the velocity is the effect of texture. The appears between (lowest velocity, sample AR(90) ~6,110 m/s), (highest velocity, a difference largest difference in velocity ~6,250 of approximately m/s) and 2.25%. sample We DA(0) see lesser velocity variations between orientations in the beta annealed samples, due to a reduction in the originally present texture. The directional velocity dependence in samples AR, DA, and RA is due to the presence of macroscopic texture. Even though samples DA and RA were heat treated to transform the microstructures, residual macroscopic texture is clearly still present in the structures. This is also evident from the metallographically prepared DA and RA samples which reveal a heavily banded post heat treated structure, similar to 39 that observed in the AR sample. Beta annealed samples FBA have trace amounts of texture still present even though and CBA the evidently also are microstructures dramatically altered from the as-received material condition. If we were to assume the material isotropic, then the measurements on the as- received sample demonstrate that the combination of Lame's constant are higher in the transverse direction. Hence, Young's Modulus is clearly higher in the transverse direction. This result comes from the relationship between acoustic velocity and stiffness, which can be expressed as: G.1) ¢) = Ae 2e p for an isotropic solid. This velocity dependence in the anisotropic as-received material is not too surprising since it has been known for many years that titanium often develops a preferred orientation due to deformation. The type of texture observed in the as-received bar stock is consistent with that of a wire texture with a uniform reduction in the crosssection of the bar. The Ti-6Al-4V bar material studied, has sustained the largest strain and deformation in the axial direction, therefore, we see the formation of a heavily banded macroscopic structure along the axis of the bar. What is not so obvious, is the orientation of the basal plane normals. The material is clearly highly textured 40 in terms of the alignment of features in the structure and the velocity data suggests the presence of a preferred crystallographic orientation in which the principal axis of the hexagonally symmetric alpha grains tend to preferentially lie in the transverse direction of the bar. This fact was corroborated by x-ray analysis, using Shultz back-reflection. The development of texture in the Ti-6Al-4V bar is due to a combination of intrinsic crystallographic and structural anisotropy [17]. The longitudinal velocity data indicates the basal normal directions are preferentially oriented to the transverse direction of the bar as depicted in Figure 3.7. This figure shows the principal directions tend to rotate to the x-y plane, but are not aligned uniformly in the radial direction of the bar. Similarly, in a rolled plate the principal crystallographic directions tend to align laterally to the high deformation directions used in the processing [1, 2, 18, 19]. We can see from the velocity data that the heat treatments have developed microstructures that are much alike in terms of the overall elastic properties, except in the beta annealed samples, which effect the longitudinal velocity significantly more normals tend to a preferred processing, but what is more orientation in the due to recrystallization. transverse direction as The basal a result influential to the velocity of propagating waves of is the presence of large colonies of similarly oriented crystals in the axial direction. These large oriented colonies tend to essentially behave as if they were single crystals that have very high longitudinal aspect ratio. Consequently, as the sound propagates in the axial direction of the textured material, the ultrasonic wavefront becomes distorted because of local velocity extremes. This phase distortion causes the coherent backwall echoes to vary in amplitude due to interference and can result in misleading measurements of the 4] Figure 3.7 The hexagonal principal directions preferentially rotate to the radial direction (x, y plane) in the as-received titanium bar stock, based on longitudinal velocity and x-ray analysis. ultrasonic properties if one is not careful to do significant averaging. Shear wave velocity and birefringence [20, 21] measurements allow more extensive understanding of the effects of texture and microstructure as these waves are generally polarized, Shear waves allowing greater flexibility for materials incident normal to the face of a sample will, characterization studies. thanks to birefringence, generally be resolved into two transverse waves vibrating at right angles to each other and propagating at different velocities. This effect is more dramatic in single crystals and anisotropic polycrystalline metals due to the directional dependence of velocity. The shear wave velocity data of Figure 3.8 shows significant velocity variations between the various samples, especially evident in the as-received samples (see Table 3.4 — Shear Wave Velocity, Appendix A). The most dramatic evidence of this influence is illustrated 42 sample, with the difference between the slow and fast pure mode in the AR(90) of approximately 4.8%. Similar to the longitudinal velocity data, we see in the as-received samples a steady decrease in the birefringence delta as the incidence angle goes from transverse to axial, with an axial birefringence delta of less than 1%. When evaluated in the axial orientation, all of the samples exhibit a modest (Ag < 1%) difference in velocity between the slow and fast pure shear modes. These low values for birefringence deltas are indication an that the macroscopic texture influencing the shear wave propagation in the axial direction is essentially axisymmetric. That is, the shear velocity is basically independent of the polarization direction for propagation in the axial direction. On the other hand, for wave propagation in the transverse direction, the shear polarization direction has a major influence on the wave velocity. It was interesting to note that for each sample evaluated in the transverse direction, with the exception of sample CBA(90), 3350 T ;0)wWww § ww wo 0° gg nq 31501NJ NAAN iS S slow mode A 72 47 3 ne in 310 +NZ \Z “Z «7GF A427 YZ sooo N40. WCQ shear velocity (m/s) 3300 + 7G 62,7, tast mode i . y y j jy G 4,4 y , wnrsgnunvyv 4NA NZ V9 NZ ZN GW [GAM YGNYGYVENMYVYWV TNGNWGYGNWVYNMYGV ,NA NA NZ NZ, NANA AR(90) AR(60) AR(45) AR(37) AR(30) AR(0) SANZ DA(0) DA(90) RA(0) RA(90) FBA(0) FBA(90)CBA(0) sample Figure 3.8 Shear wave velocity data taken at different polarization angles consistent with the slow and fast pure shear modes. 43 the direction consistently supporting was the fast mode polarized transversely with respect to the macroscopic inhomogeneity, as shown in Figure 3.9. Unfortunately, sample CBA(90) was too attenuative to take accurate velocity measurements. Wave propagation in the transverse direction gives rise to the fast mode only when the particle displacements are perpendicular to the macroscopic inhomogeneities. This finding makes sense in light of the longitudinal velocity data, which also exhibited faster velocities for waves traveling perpendicular to the aligned features, which are in the axial direction. Conversely, for transversely directed wave propagation, the slow mode occurs when the particle displacements are parallel to the inhomogeneity, the less stiff direction. The shear data is another indication that samples AR, DA, and RA demonstrate S more stiffness in the transverse direction. + -———-) “os ' me ~ ra Ft Pra ———— wv” fast mode of SK Figure 3.9 This diagram shows the fast and slow pure mode polarization direction in the 90° samples. Note the fast mode occurs when particle displacement is perpendicular to the alignment bands in the axial direction of the bar. 44 In terms of elastic properties, if we assume the polycrystalline titanium alloy is isotropic, then the shear modulus is clearly higher when the polarization is in the transverse direction as compared to when the polarization is in the axial direction. This result comes directly from Lame's elasticity relationship expressed as: c= (3.2) |= Pp for shear waves, where 11, is the same as the shear modulus, G. The RA and DA heat treatments appear to have only a minor influence on shear velocity. The two beta the same velocity sample has the and annealed samples have essentially The coarse beta annealed .5%. This is consistent with the idea that the birefringence characteristics. birefringence delta of approximately smallest relatively long time (6 hours) above the beta transus has effectively wiped out most of the initial texture, leaving behind a coarse grained, but randomly oriented microstructure. The other axially oriented measurements in AR(0), DA(0), and RA(O) all show nearly the same birefringence values, with minor variations in shear velocity between samples. The scatter in the shear velocity is slightly higher than in the longitudinal velocity data. This is primarily due to the fact that proper coupling is more difficult to achieve in shear measurements, requiring a viscous coupling agent, compared to water coupling for the longitudinal measurements. 45 3.3 Ultrasonic Attenuation Attenuation the loss of a signal's amplitude generally means with increasing propagation distance. The loss is defined as the ratio of two amplitudes and expressed in logarithmic units, Neper or decibel (dB) [22]: (3.3) L(Neper) = inh or L(dB) = 20 log AL, Ay Ay where A, and 4, denote the amplitude with and without attenuation, respectively. In some cases loss the JZ occurs locally as a result of interaction with a material discontinuity. These losses could be due, for example, to reflection and transmission at an interface or scattering losses at a rough surface. There are other losses which occur over a given distance as the wave propagates in the medium, but not necessarily proportional to the distance covered; such losses are usually associated with divergence of the beam. Attenuation of a medium is usually limited to phenomenon that cause loss proportional to the propagation distance, expressed as L =a d, where d is the propagation distance and a is the attenuation coefficient. Ultrasonic attenuation is the rate of decay of mechanical energy as a wave propagates through a material. There are two major classes of attenuation mechanisms considered important for ultrasonic materials characterization, which can be expressed as: & = & absorption (3.4) +a scattering’ 46 First, absorption conduction, acoustic converts energy to absorbed elastic hysteresis, etc. The heat viscosity, via energy relaxation, heat is irreversibly lost from the acoustic field and is dissipated in the medium. Second, scattering converts the energy of the coherent, collimated beam into incoherent, divergent waves as a result of interaction with inhomogeneities in the material. The scattered energy is not necessarily lost as at least part of it can be also picked up by the same transducer used to receive the coherent wave as backscatter. Hence, scattering does not only reduce the amplitude of the coherent signal, but also gives rise to an incoherent material noise which further limits the detectability of defects in the attenuated signal. On the other hand, the same signals that block our view to the development of low-level (fatigue) damage, also allow us to characterize materials, based on ultrasonic backscattering. Attenuation measurements were conducted in the same titanium sample set that velocity was measured, as described in the previous measurements were taken in accordance with the ASTM section. The attenuation standard practice, (E664) [23]. For these measurements a conventional immersion ultrasonic system was used to avoid difficulties in coupling breakdown suffered by contact ultrasonics. One of the main problems encountered was due to the fact the attenuation is spatially distributed in the titanium alloy samples due to the highly inhomogeneous nature of the material. Therefore, attenuation measurements were spatially averaged over approximately 1" x 1" on each sample. This was done by averaging signals while the transducer is being scanned over the sample. Approximately 2 k waveforms were averaged over a 1" x 1" area on each of the samples and repeated several times to ensure consistent results. The basic measurement involves digitally averaging waveforms while the transducer is moving, resulting in a spatial average. Each of the acquired spatially averaged signals correspond to multiple round trip responses inside the material, as illustrated in Figure 3.10. The transducer is pulsed via broadband excitation and then is a receiver for a relatively long period (about 1 ms) during which time the wave propagates back and forth multiple times through the 1.5" thick titanium samples. As the wave travels inside the material, the amplitude of the coherently reflected waves continuously diminish due to grain scattering, beam reflection, non slightly and divergence, parallel plane surfaces. beam The divergence problem is one of diffraction from an aperture and has been solved for longitudinal waves from a circular transducer of a given radius which radiates waves into different fluids [24, 25]. These solutions work well for wave propagation in focused transducer excitation pulse amplitude | specular reflection water path rn Ae ‘ ee “ ae backwall echoes re ' 7 15*round ' > ql t ' “‘sample' Figure 3.10 mclope ~ exp(-otx) 3" time trip echo Illustration of the ultrasonic wave propagating back and forth inside the ‘material and the associated waveform showing the multiple coherent echoes. 48 homogeneous, isotropic materials. The problem of beam divergence has also been solved for anisotropic media [26, 27] along axes of three, four, and six fold symmetry. The corrections required to compensate for diffraction can be experimentally measured for a given transducer or can be analytically calculated for models. The simplest and most often used model for ultrasonic transducers.is a circular piston radiator. The acoustic pressure field of an unfocused circular radiator is illustrated in Figure 3.11. There is a distinct difference in acoustic behavior between the near-field and far-field regions. The transition between these two regions is defined as: where a is the aperture and 4 is the wavelength. For these attenuation measurements all the sample thickness’ were the same and all measurements were taken in the far-field so beam corrections were not used. for the purpose of that the quantities we measure for While the attenuation is a valuable parameter to measure materials characterization, we should recognize attenuation are composed of many complex interacting phenomenon and not easily separated. Moreover, the attenuation is integrated over the entire sound path inside the material so the measurement is not well suited for locating anomalous indications or for resolving local structural variations. Attenuation measurements are also very dependent on the frequency and the coupling between the probe and the object. nfat far-field zone near-field Figure 3.11 The acoustic field of a circular piston source showing variations at discrete distances from the transducer. the pressure The data from the attenuation experiments demonstrate that the microstructural differences between the samples are quite modest, with the exception of the coarse beta annealed sample, which clearly has the highest attenuation. In Figure 3.12, a plot of the frequency dependent attenuation losses (see also Table 3.5 - Ultrasonic Attenuation, Appendix A). Its not surprising sample CBA is by far the highest attenuating sample in both axial and transverse orientations with its very large prior beta grain structure and the presence of large randomly oriented a+ colonies, which tend to strongly scatter the incident energy. However, one unfamiliar with the nature of attenuation in Ti-6Al-4V may be surprised to learn the AR(0) sample has the next highest apparent attenuation of the lot. This is especially interesting after realizing that in AR(0) the wave is traveling parallel to the aligned macroscopic colonies. Despite the fact that the waves are traveling parallel to the alignment direction, the attenuation is approximately double that of the transverse sample, AR(90). One of the reasons for this unusual attenuation phenomenon has to do with the phase sensitivity of the receiver and the fact that the phase of the wavefront is much more nonuniform in the axial direction. Part of this attenuation anomaly is thought to result from localized extremes in velocity that perturb the 25 4 ~N oO i l —* DA(90) —® RA(90) —4— FBA(90) —e CBA(90) —*- AR(90) j — — ~ nN N Oo ! Radial Attenuation faa as) \/ ~” N nN 2 8 10 12 14 frequency (MHz) a) aa) Mo) —o— DA(0) — RA(0) —2- FBA(0) 2 CBA(0) —*— AR(0) 1 — SN N Oo l Axial Attenuation iv) Ny — So \ ~~ nN WN 2 b) Figure 3.12 10 12 14 frequency (MHz) Frequency dependent attenuation losses for the titanium alloy samples in the (a) radial direction, and (b) axial direction. 51 wavefront. The phase distortion effect on attenuation, which is really not related to the loss of energy commonly associated with attenuation, will be discussed in Chapter 5 where a system to map microscopic phase variations has been developed. Samples DA(0) and RA(0) have about the same attenuation followed by the group consisting of FBA(0), AR(90), and FBA(90), respectively. Similar to the attenuation of AR(0), samples DA(0) and RA(O) have somewhat unusual behavior as the optically apparent microstructure in either sample is quite highly refined with equiaxed grains. The explanation for the relatively high apparent attenuation in DA(0) and RA(0) is due to the minor effect the annealing has on the attenuation. Remember, DA(0) was held in the a+B phase field for only one hour at 937 °C and RA(0) was held at only 900 °C for four hours, which, despite the fact that the microstructures clearly change, evidently does not perturb the elastic properties of the structures in a very meaningful way. It is also clear from the metallography results that the banded macrocolonies are leftover from the mill anneal in both DA(0) and RA(0) in spite of the highly refined microscopic appearance with equiaxed primary alpha grains averaging about 20 1m in diameter. The group consisting of samples FBA(0), AR(90), FBA(90) is closely separated in the attenuation chart, with FBA(0) being only slightly higher than the other two. Samples FBA(0) and FBA(90) are not distinguishable in terms of the optical microscopy between the two orientations and, being beta annealed, as some of the texture has been removed due to recrystallization. The ultrasonic properties of sample FBA are slightly directionally dependent. Sample AR has strong directional 52 dependencies and quite differently appearing microstructures between the two orientations of interest. Samples DA(90) and RA(90) also had very similar velocity and obviously are alike in the sense that the data is in both cases taken from the transverse orientation. In general, the transverse orientations tend to be less attenuative than the axial orientations. What is probably most intriguing about these lower attenuating samples, including AR(90), is the fact that the highest scattering direction coincides with lower attenuation. For example, sample AR(90) significantly has a higher degree of microstructural scatter (~12 dB) than AR(0). This is because in AR(90) the waves are propagating perpendicular to the aligned colonies of the mill annealed microstructure. Propagation of sound waved in this transverse direction is optimal for the generation of backscattered energy and yet, AR(90) is one of the lower attenuative samples. This is valuable and information underscores the importance of understanding peculiarities associated with titanium alloys. To further illustrate this point, Figure 3.13 shows two different oscilloscope traces taken at random from the transverse and axial direction of the mill annealed microstructure using identical data collection parameters (i.e., gain, damping, energy, focal distance, etc.). This figure demonstrates that the attenuation is not necessarily related to the microstructure scatter in Ti-6Al-4V, which is unusual considering the vast majority of materials have their attenuation primarily dictated by scattering. The amplitude of incoherent microstructural scattering is also very important from the standpoint of defect detection. Materials with higher scattering are generally more difficult in terms of detecting and characterizing tiny defects, as the flaws are masked by the scattering. The problem of evaluating the as-received titanium alloy from AR(90) microstructure scatter AR(0) frontwall echoes Figure 3.13. backwall echoes Oscilloscope signal traces showing the difference in amplitude microstructure scatter from the mill annealed samples. the transverse direction is closely related to the billet NDE of the qualification issue. Large titanium billets are also examined by probing from the transverse direction, hence the qualification of billets is hindered by the presence of very strong microstructure scatter, making it difficult to discriminate between benign scattering versus genuine defects. 3.4 Ultrasonic Grain Scattering Polycrystalline materials consist of individual grains of the constituent material, which compactly mold together forming a structure. The grains are generally a distribution of shapes and sizes filling the space within the boundaries of the medium with uniquely titanium oriented crystalline alloys, the grains structures. In inhomogeneous often tend to combine into large materials colony like some structures with preferential orientation existing at the macroscopic level. Hence it is possible to have a 54 material that is inhomogeneous on more than one dimensional scale due to a wide distribution of grain sizes or the presence of macroscopic grain colonies. The grains may state or by recrystallization during materials form by crystallization from the molten processing. For different grain morphologies, a single grain may be composed of a single crystal; it may have two or more phases breaking it up, or the structure may consist of both single grains and heterogeneous grains, as is typically the case for Ti-6Al-4V. Each individual grain can be assigned a set of axes corresponding to the principal crystal directions of the major constituent phase. The crystal orientation is generally different from grain to grain, except in samples containing a preferred orientation. The geometry of the grains may be flattened, elongated, or basically spherical, which are termed equiaxed. The simplest polycrystalline materials are equiaxed and homogeneous (narrow grain size distribution), with inclusions. Simple and no preferred microstructures are nicely single phase or orientation, colonies, voids, for ultrasonic materials suited characterization techniques to determine such parameters as grain size, stress state, and elastic moduli [31, 32, 33]. More complicated structures tend to be difficult to evaluate in terms of deriving quantitative material characteristics from the ultrasonic response. This difficulty is mainly due to multiple scattering and other interference effects. Scattering measurements, both forward and backward, have been conducted on a number of different titanium microstructures. This section is geared towards developing an understanding of ultrasonic scattering from a complicated polycrystalline material. The main goal for this section is to describe the scattering for the various titanium samples and discuss the mechanisms responsible for originating the scatter. One of the approaches 55 used to determine material characteristics is to collect data from different sample orientations. By measuring the scattering response from different directional orientations, we may determine if the signals are consistent with the presence of structural alignment, indicating mechanical anisotropy, based on the scattering coefficients. This section will also feature the importance of signal processing techniques which are crucial to the interpretation of the data. The frequency dependence of scatter is another area where useful information can be collected, but this is generally true for grain size measurement in simple microstructures only. For most materials, grain scattering primarily influences the attenuation. However for titanium alloys, this is not always the case due to factors related to materials processing and the resulting complex microstructure and development of texture. The anisotropy of titanium, especially in the highly textured mill annealed condition, is an important factor that influences not only the mechanical load response, but also the scattering and attenuation. Figure 3.14 shows the narrow-band and broadband experimental configurations for scattering measurements. For simple microstructures, the narrow-band experimental approach is typically used to acquire information regarding the nominal dimensions of scatterers in the material. This approach is taken to determine the frequency dependence of attenuation and scattering [34], which can be used to assess grain size in single phase materials with narrowly distributed equiaxed however, a broadband experimental grain sizes. In complex approach is usually taken as microstructures the frequency dependence does not provide much insight into characteristics of the nominal scatterer. The broadband approach allows the qualitative determination of different microstructure Zz digital oscilloscope computer controlled scanner X . pre-amplifier receiver output hd 2 9 t J 2 9 C=—) function generator ° rf power amplifier ° in © out & +55 dB trigger output sine wave burst ~ 15 cycles Narrow-Band Experiment computer Zz digital oscilloscope controlled transmitter / receiver Ur eee sync output Ad |! $F- Broadband Experiment Figure 3.14 different experimental of the two Schematic microstructure characterization using backscattering. used set-ups for types, but generally provides little in terms of quantitative material characteristics. Both the narrow-band and broadband experiments involve measurements of the backscattered signals. The main difference between narrow band versus configurations is the excitation source. For the narrow-band broad band experimental experiment, a sinusoidal burst of approximately 20 cycles of the desired harmonic frequency is amplified and used to excite the transducer. This results in the excitation of a narrow range of signal frequencies. For the broadband experiment, a spike (impulse) excitation is used, which results in the generation of a wide range of frequencies in the transmitted and received signals. 57 Many problems can arise in experimentally determining the frequency dependence of attenuation and scattering in polycrystalline metals. For example, every time the frequency is changed, the amount of energy imparted to the material also changes, hence gain-compensation should used in order to maintain consistent experimental conditions in terms of the incident energy. Generally, ultrasonic transducers are broadband enough that gain-compensation requirements are negligible so long as we use frequencies near the of the spectrum. middle arising problems due to translates into a problem gain compensation The frequency where is easily manageable beam dependent divergence. Beam compared to divergence frequency variations result in a different interaction volume every time the frequency is changed. In other words, lower frequencies tend to have higher beam divergence and increasing the frequency results in a more collimated beam. Hence, every increase in frequency results in a smaller and smaller interaction volume of material in which the waves interact with the medium. In practical terms what this means is that experimentally it is difficult to observe the predicted power of four frequency dependence in the Rayleigh scattering regime (power of two in stochastic regime) unless sophisticated diffraction corrections are implemented. This correction is not so difficult for simple microstructures, but is quite in the more in polycrystalline alloys challenging complicated microstructures seen in titanium alloys [24, 26, 27]. An important technique used to evaluate scattering involves spatial averaging in either the time or frequency domain [35, 36]. We know, for example, that from location to location on the samples the ultrasonic backscatter response changes. These changes are demonstrated in Figure 3.15 showing backscattering signals, 58 using a focused transducer, from five randomly chosen locations on the AR(90) sample. Each has position a different backscatter response because the structure within the interaction volume changes. Backscatter signals are highly influenced by slight changes in the experimental set-up. For example, by changing the incidence angle of the transducer \ amplitude fitter time b) vonateoonnny m ot mrelprartan bo d) > oN A >) iN . rectified rf signals a,b,c,d,e 10K rectified rf signals transducer focused on surface \——— transducer focused ~ 0.5" inside sample ‘\ Figure 3.15 Spatial averaging of backscattered longitudinal waves in the time domain, using broadband excitation with a transducer having a center frequency of approximately 8.5 MHz. 59 of a degree a fraction by only results in an entirely different backscattered signal. Likewise, a position change of approximately one grain diameter or one wavelength will suffice to drastically change the appearance of the signal. These signal disturbances in Ti6Al-4V could be due to many factors, such as grain-to-grain elastic property variations, In ultrasonic materials characterization, polycrystalline metals generate an incoherent and highly divergent field that appears as "grain noise" in the detected ultrasonic signals. It is important to realize that these signals do not necessarily correspond to any particular physical feature inside the material structure, but rather, are generated due to interference and multiple scatter effects of the sound waves interacting with a volume of material. Multiple scattering is especially important for titanium alloys as they tend to scatter ultrasonic energy on more than one dimensional scale consistent with the formation of microscopic features (i.e. grains) as well as macroscopic features (i.e. grain colonies). When produces a sound wave its own filled with tiny objects, each object traverses a volume array of scattered waves. These scattered waves reinforce in some directions and interfere in others, and the wave incident on each scatterer is affected by the presence of other scatterers. These interactions gives rise to coherent, incoherent, and multiple scattering. The coherently scattered signals contribute additively to the amplitude of the transmitted wave. By subtraction of the coherent response, the remaining energy is due to incoherent and multiple scatter. The incoherent scatter is apparent as low amplitude signals (40 to 60 dB below the coherent echoes) occurring between the two surface reflections from the front and back walls of the sample [37]. This incoherently scattered energy can be easily measured with a receiver which has the appropriate signal 60 amplification and filtering capabilities. For backscattering, the data collection scheme starts with the sample being placed in the immersion tank on a tilt stage and then leveled to approximately 4/10 or+5 pm flatness over about one square inch on the sample. The leveling is achieved by repeatedly scanning the transducer over the sample and adjusting the tilt stage so that as the transducer moves, there is no change in the water path length. If the water path varies as the transducer scans, the oscilloscope signal trace correspondingly varies in time. Once the sample is level, the time drift disappears. Sample leveling is extremely critical if accurate results are to be acquired. Next, the nominal coherent transducer response must be acquired by collecting a global spatially averaged signal. The transducer response is to every signal trace in the spatially averaged scan and generally the signal common requires performed 15 k to 20 k signal samples for adequate convergence. The averaging is as the transducer is scanned over approximately a 1" x 1" region on the sample. The transducer response must be re-acquired for each new material sample and the experimental conditions must not change from sample averaging measurement would be extremely time consuming to sample. The spatial and tedious without the advent of computer controlled scanning and an on-line digitization capability provided by a digitizing oscilloscope. With the transducer response acquired, the final data acquisition steps can be performed, resulting in the collection of signals representative of the average backscattered energy for each sample. A new spatial average is started, essentially repeating the process used to acquire the coherent response, except and for each collected rf signal, the coherent response is subtracted out on-line. What is remaining after the response coherent incoherent is subtracted at each signals are then squared and a point is the incoherent scattering. These of all of these signals spatial summation provides the input for the last step in the backscatter energy acquisition process. The last step is to take the summed average of the squared incoherent signals and then take the square root. The signal resulting from this last step is directly related to the backscatter energy response of the sample as it is an rms average of all the material responses. PART II: The New Forward Scattering Method 3.5 Ultrasonic Forward Scattering Forward scattering measurements provide information on the macroscopic structural composition of the alloy. Unfortunately, this research area is relatively new in terms of information documented in the literature. This is a little surprising, considering the advantages this type of measurement has over conventional backscattering. First, forward scattering requires a transmission experiment so the amplitude and divergence of the scatter are both viable sources of information. Secondly, the data evaluation is open to conventional C-scanning, rather than the more complicated spatial averaging used in backscattering. Finally, unlike backscattering, the time dependency of the scatter can be mapped in forward experiments. Figure 3.16 shows the schematic of the forward scattering experiment. A 0.25" contact transducer (transmitter) is mechanically mounted to one surface of the sample. On the opposite surface, a second transducer (receiver) is aligned at normal incidence to the transmitter. This position becomes the center of the C-scan to map the amplitude and divergence of the incoherent forward scatter. The transmitter launches a wave through the material receiver. the towards This coherent wave at the is detected receiver and immediately following its arrival begins the forward scattered energy, which is actually the backscatter from the internally reflected coherent wave. Peak-detecting electronic gates are used to sample the forward scattered data, which forms the image in the Cscans. By raster scanning the receiver the scattered field can be mapped to assess the influence of texture and microstructure on the amplitude and divergence of the scatter. Unfortunately, this mapping procedure can not be used in backscattering experiments, since the transmitter and receiver are one and the same. . . scanning receiver first coherent transmission sample / transmitter reflection th 7 incoherent forwar gate positions scatter <t | reflected —_ coherent waves ot Ci a FORWARD SCATTER mounted” transmitter Oscilloscope controlled scanner transmitter / receiver tt syne outpu water tank Figure 3.16 Schematic of the forward scattering experiment. 63 Spatial forward and directional and backward measurements averaging directions the Ti-6Al-4V been have samples. The conducted amplitude in the modulated incoherent scattering signals are specially suited to evaluate materials for microstructural variations. Monitoring of the orientation dependence and spatial variation of the amplitude modulation in scattered signals provides an indication of the shape of the nominal material inhomogeniety. For randomly oriented grain structures the scattering of sound waves from material inhomogeneities generally occurs in all directions. Randomly oriented materials with coarse grain structure generally have higher attenuation because the anisotropy of individual grains scatter away the energy of the propagating waves faster than in the equivalent fine grained material. In Ti-6Al-4V, the material inhomogeneity can take on the form of colonies or clusters of similarly oriented grains, which often appear as a result of wrought processing operations like forging. Figure 3.17 shows an example of a backscattered signal with and without high-gain amplification, including the front and back surface echoes. The normal incident spatially averaged backscatter energy data for the titanium samples in the transverse and axial orientations is shown in Figure 3.18 (see also Table 3.6 — Ultrasonic Backscattering, Appendix A). These signals were acquired using broadband excitation with a 10 MHz ultrasonic transducer focused at the surface of each sample. Surface focusing allows for the evaluation of grain scattering without the influence of focusing affects inside the material, since the propagating sound waves continuously diverge from the entry surface. The original cylindrically shaped samples were prepared such that data could be collected from both the axial and transverse directions through flat and parallel 64 I Broadband excitation 10 MHz longitudinal wave amplifier: + 19 dB rn | backwall echo frontwall echo Broadband excitation 10 MHz longitudinal wave amplifier: + 59 dB backscatter Figure 3.17 Backscattered signals from sample AR(90) showing the effect increasing the amplification by 40 dB to reveal the backscatter signals. of entry surfaces to allow direct comparison of scattering from the samples. The data was collected in both orientations using the same transducer, under the same scan parameters and hardware scan settings. At normal incidence, clearly the highest backscattering occurs in the transverse orientation of the mill annealed sample, AR(90). highly aligned nature of the inhomogeneity This is not too surprising given the in the axial direction of the sample. In contrast, the axial direction of the mill annealed sample (AR(0)) demonstrates the lowest backscattering of longitudinal waves, along with a group of other samples which all have 65 Figure 3.18 i ° ‘p amplitude (volts) o ro o a —™ AR(90) — DA(90) —*— RA(90) —®— FBA(90) —+- CBA(90) —— CBA(0) 1+ FBA(0) —t— AR(0) —— DA(0) Spatially averaged backscattered signals from the axial and transverse orientations, using flat entry surfaces with beam focused on front surface. nearly identical backscattering characteristics. With the exception of the beta annealed RA(90) also clearly show evidence of the highly aligned macroscopic features similar to sample AR(90). Sample DA(90) appears to have backscatter coefficients only slightly samples (i.e., FBA(0) and CBA(0)) all of the samples, in the axial direction, had virtually the same negligible backscatter response. Hence, the mill annealed sample has both the lowest and highest scattering conditions depending on which orientation the data is collected from. The scattering is highly anisotropic, with the optimal generation of scatter occurring the transverse orientation. Samples DA(90) and RA(90) also clearly show evidence of the highly aligned macroscopic features similar to sample AR(90). Sample DA(90) appears to have backscatter coefficients only slightly stronger than RA(90). The metallographic results show the DA and RA heat treatments have more equiaxed grains as compared to the heavily elongated grains and colonies of the mill annealed condition. However, both DA and RA samples clearly retain a significant portion of the original texture, and there clearly exists evidence of banding on a macroscopic scale. direction, the lowest backscatter energy For data collection in the transverse comes from sample CBA(90) with sample FBA(90) These beta annealed samples are more structures as indicated by the scattering only slightly stronger. closely representative and attenuation velocity of randomly results. oriented Moreover, the backscattering data shows that in sample CBA the response is not strongly influenced by orientation affects. The beta annealing heat treatments have significantly transformed the microstructures due to the dwell time above the beta transus resulting in a more thoroughly recrystallized structure where most of the original texture is overwritten by the more random orientation of the new grain structure. The reason for the stronger scatter in FBA(90) is due to the presence of retained texture. Hence, sample FBA is simple more anisotropic than sample CBA. In the axial orientation, the two beta annealed samples clearly backscatter the most, with sample CBA(0) the strongest, followed by sample FBA(0). This scattering is consistent with the idea that these samples generate more randomly directed scatter. Samples AR(0), DA(0), and RA(0) all backscatter energy the same low intensity manner, with virtually 100% overlap of the signals in the axial orientation. 67 Surprisingly, this minimal backscattering characteristic is true for the entire mill annealed sample set (i.e., samples AR(0), except AR(90). AR(30), AR(45), and AR(60)) Overall however, this should not be taken to imply that these samples scatter any less than AR(90). Rather, only the backscattering energy 180° (directed from the wave propagation) shows this minimum characteristic. At the other than orthogonal orientations, the scatter is simply directed away from the backward direction. Figure 3.19 shows results from the mill annealed samples in the axial and radial directions over a 2" x 2" area. For the axial direction the forward scatter is highly divergent and has a characteristically conic symmetric for all directions in the plane of the scan. Moreover, images. this divergence can be clearly observed In contrast, the mill annealed sample C-scan in the time-lapsed in the radial direction has a distinctly different pattern to the forward scatter. In this case, the initial divergence (i.e. gate 1) is much more highly concentrated and the image sequence shows a dramatically different scattering pattern as compared to the axial case. The scatter divergence emerges as a fan beam, rather than a cone beam as observed in the axial case. The for the reason contraction in the axial direction (indicated by the white arrow in gate 2) is that the elongated grains and grain colonies tend to act as a diffraction grating. Hence, for uniform and equiaxed grains, as viewed from the axial direction, the scatter diverges symmetrically as opposed to the elongated grain colonies, as viewed from the radial direction, which causes the divergence to be much more one-dimensional. In terms of amplitude, it is of great significance to note that the direction of wave propagation is of little or no consequence as the distribution of amplitude values is virtually the same 68 Sample AR(0) (axial direction is out of page) Sample AR(90) (axial direction indicated by white arrow) gate 1 Figure 3.19 gate 2 gate 3 Time-lapsed C-scan images mapping the amplitude and divergence of forward scattering in mill annealed Ti-6Al-4V, showing clearly distinguishable differences between the axial versus radial directions. between the axial and radial directions, only the shape of the divergence pattern changes. Figure 3.20 shows the amplitude and divergence of forward scatter for the coarse beta annealed (CBA) sample over a 2" x 2" area. This sample was held for approximately six hours above the beta transus (~1000 °C) providing ample time for the material to fully recrystallize, thereby removing virtually all of the initially present texture. As such, this sample has large randomly oriented grain colonies and upon viewing the metallographically prepared surfaces under magnification, it is not possible to distinguish the axial direction from the radial one. The disappearance of the macroscopic anisotropy and random grain orientation is also apparent in the forward scattering results. These results demonstrate the lack of distinguishing features between the axial and radial directions for a random structure, as opposed to the highly anisotropic mill annealed 69 Sample CBA(0) Sample CBA(90) gate 2 Figure 3.20 gate 3 Time-lapsed C-scan images mapping the amplitude and divergence of forward scattering in coarse beta annealed Ti-6Al-4V, showing indistinguishable differences between the axial versus radial directions. structure. Additionally, the amplitude distributions are virtually the same between axial and radial directions of the CBA samples, similar to that of the AR samples. Based on the backward and forward scattering results it is possible to more thoroughly describe the scattering cross-sections. This is particularly important in terms of understanding attenuation results the in attenuation results mill annealed the for the mill structure annealed are unusual microstructure. The the high due to backscattering in the radial direction, but low attenuation than that of the axial direction. By constructing the notional ultrasonic cross-sections, we can better understand why this unusual attenuation behavior exists. Figure 3.21 shows these notional scattering crosssections for the axial and radial directions. The total scattering for the axial direction in the mill annealed structure is larger than that of the radial direction. Therefore, despite the 70 stronger backscattering in the radial direction, the axial direction has significantly higher attenuation. WIA shear setetetetetete longitudinal back random structure Figure 3.21 Notional scattering cross-sections in the axial and radial directions for both anisotropic and random structures, which were generated based on the backward and forward scattering results. Chapter 3, References: 1. Materials Property Handbook on Titanium Alloys, eds., R. Boyer, G. Welsch, and E. W. Collings, (ASM International, 1994) pp. 483-500. 2. Titanium, A technical guide, ed., M. J. Donachie, Jr., (ASM International, 1988) p. 44. 3. S. L. Semiatin et. al., “Hot working of titanium alloys — An overview,” Advances in Titanium Alloy Processing, eds., 1. Weiss, et. al., Proceedings from the 125" TMS Meeting held in Anaheim, California, 5-8 February, 1996, (TMS, 1997) pp. 4-26. 4. Metals Handbook (ASM, 1973) pp. 45-47. 5. D. Eylon, "Faceted fracture in beta annealed titanium alloys," Met. Trans. A Vol. 10, pp. 311-317 (1979). 6. D. Eylon and J. A. Hall, “Fatigue behavior of beta processed titanium alloy IMI 685,” Met. Trans A, Vol. 8, pp. 981-990 (1977). 7. S. Hirsekorn, “The Dependence of Ultrasonic Propagation in Textured Polycrystals,” Journal of the Acoustical Society of America, T9(5), (1986). 8. C. M. Sayers, “Angular Dependent Ultrasonic Wave Velocities in Aggregates of Hexagonal Crystals,” Ultrasonics, Vol. 24, pp. 289-291 (1986). 9. C. M. Sayers, “Ultrasonic Velocities in Anisotropic Polycrystalline Aggregates,” Journal of Applied Physics D, 15, p. 2157, (1982). 10. E. S. Fischer and C. J. Renken, "Single-Crystal Elastic Moduli and HCP - BCC Transformation in Ti, Zr, and HF," Physical Review, 135(2A), pp. A482-A494, (1964). 11. A. Migliori, et. al., “Resonant Ultrasonic Spectroscopic Techniques for Measurement of the Elastic Moduli of Solids,” Physica B, Vol. 183, pp. 1-24, (1993). 12. L. Adler, “Ultrasonic Spectral Analysis for Nondestructive Evaluation,” (Plenum Press, 1981), p. 21-25. 13. A. F. Brown, “Materials Testing by Ultrasonic Spectroscopy,” Ultrasonics, September, pp. 202-210, (1973). 14. W. M. Visscher, et. al., “On the Normal Modes of Free Vibration on Inhomogeneous and Anisotropic Elastic Objects,” Journal of the Acoustical Society of America, 90(4), Part 1, pp..2154-2162, (1991). 15. ASTM Standard E 494-75, “Standard Practice for Measuring Ultrasonic Velocity in Materials,” ASTM (1985). 16. E. P. Papadakis, “ The Measurement of Ultrasonic Velocity,” Ultrasonic Measurement Methods, Physical Acoustics , Vol. XIX, Eds. R. N. Thurston and A. D. Pierce, (Academic Press, New York, 1990) p. 91. 17. F. Nye, "Physical Properties of Crystals - Their Representation by Tensors and Matrices,” J. (Clarendon Press, Oxford, 1985) p. 143. 18. In Proceedings on Advances in the Science and Technology of Titanium Alloy Processing, eds., I. Weiss et. al., (TMS Publication, Aneheim California, 1996) p. 16. 19. W. F. Hosford, "The Mechanics of Crystals and Textured Polycrystals," (Oxford University Press, New York, 1993), p. 16. 20. P. C. Waterman and L. J. Teuntonico, "Ultrasonic Double Refraction in Single Crystals," Journal of Applied Physics, 28(2), pp. 266-270 (1957). 21. R. Truell, L. J. Teuntonico, and P. W. Levy, "Detection of Directional Neutron Damage in Silicon by Means of Ultrasonic Double Refraction Measurements," Physical Review, 105(6), pp. 1723-1729 (1957). , 72 22. Ultrasonic Testing of Materials, eds., J. Krautkramer and H. Krautkramer, (Springer- Verlag, NY, 1990), pp. 108-115. 23. ASTM Standard E 664-93, “Measurement of the Apparent Attenuation of Longitudinal Ultrasonic Waves by Immersion Method,” ASTM (1989). 24. H. Seki, A. Granato, R. Truell, "Diffraction Effects in the Ultrasonic Field of a Piston Source and Their Importance in the Accurate Measurement of Attenuation," , Journal of the Acoustical Society of America, 28(2), pp. 230-238, (1956). 25. K. Tjadens, "Longitudinal Wave Ultrasonic Absorption in Aluminum Alloys," Acoustica, Vol. 11, pp. 127-336, (1961). 26. E. P. Papadakis, "Diffraction of Ultrasound Radiating into an Elastically Anisotropic Medium," Journal of the Acoustical Society of America, 36(3), pp. 414-422, (1964). 27. E. P. Papadakis, "Ultrasonic Diffraction Loss and Phase Change in Anisotropic Materials," Journal of the Acoustical Society of America, 40(7) pp. 836-876, (1966). 28. M. P. Blodgett, P. B. Nagy, “Anisotropic Grain Noise in Eddy Current Inspection of Noncubic Polycrystalline Metals,” Appl. Phys. Let., 72(9), pp. 1045-1047 (1998). 29. M. P. Blodgett, W. Hassan, P. B. Nagy, “Theoretical and Experimental Investigations of Lateral Resolution in Eddy Current Imaging,” Materials Evaluation, May, 2000. 30. Microscopy Handbooks 12, A. Briggs, “An introduction to scanning acoustic microscopy,” (Oxford University Press, 1985) pp. 1-15. 31. E. P. Papadakis, “Ultrasonic Attenuation due to Grain Scattering in Polycrystalline Metals,” Journal of the Acoustical Society of America, Vol. 37, p. 711 (1965). 32. J. Szilard, and G. Scruton, “Revealing the Grain Structure of Metals by Ultrasonics,” Ultrasonics, May, p. 114, (1973). 33. J. H. Rose, “Ultrasonic Backscatter from Microstructure,” in Review of Progress in ONDE, Vol. 11, eds D. O. Thompson, D. E. Chimenti (Plenum, New York 1992) p. 1677. 34. R. L. Smith, “The Effect of Grain Size Distribution on the Frequency Dependence of the Ultrasonic Attenuation in Polycrystalline Materials,” Ultrasonics, September, p. 211, (1982). 35. J. H. Rose, “Theory of Ultrasonic Backscatter From Multiphase Polycrystalline Solids,” in Review of Progress in ONDE, Vol. 12, eds D. O. Thompson, D. E. Chimenti (Plenum, New York 1993) p. 1719. 36. K. Goebels, “Structure Analysis by Scattered Ultrasonic Radiation,” Research Techniques in Nondestructive Testing, Vol. 4, edited by R. S. Sharpe, (Academic Press, New York, 1980), Vol. IV, pp. 87-157. 37. F. J. Margetan, R. B. Thompson, and I. Yalda-Mooshabad “Modeling Ultrasonic Microstructural Noise in Titanium Alloys,” Review of Progress in Quantitative Nondestructive Evaluation, Vol. 12B, eds. D. O. Thompson and D. E. Chimenti, (Plenum, New York, 1993), p. 1753. 73 CHAPTER IV SUBSURFACE MICROSTRUCTURE ANOMALY DETECTION IN BULK TITANIUM ALLOY USING ULTRASONIC BACKSCATTERING 4.1 Introduction In this chapter, the main objective is to demonstrate the capabilities of ultrasonic inspection for the purpose of identifying and characterizing anomalous conditions in 2.5" diameter cylindrical bar samples. Implanted microstructure anomalies were diffusion bonded in the axial and radial orientations to simulate processing remnant cast structure at the center of the forged bar. Primarily, conventional B-scanning [1, 2, 3] is used as a foundation for this chapter. B-scanning of the backscattered ultrasound a provides realistic representation of the statistically distributed ultrasonic data, unlike C-scanning, which generally provides only the peak-detected signals, thereby giving a distribution of data skewed towards the large set values. B-scans were collected from a number of different regions on each of the diffusion bonded samples to examine the extent to which the anomalous microstructure inserts could be detected and characterized. This approach was also taken in order to make an assessment of the optimal scan configuration to qualify the mechanical suitability forged Ti-6Al-4V bar stock for further processing. 74 Other backscattering, scan configurations including forward were investigated scattering and besides side conventional scattering, to ultrasonic determine the effectiveness of these nonconventional approaches to identify anomalous microstructure. Unfortunately, the results clearly demonstrate that there is no advantage to using forward or side scattering as an inspection tool over ultrasonic backscattering. In fact, forward and side scattering are complicated by the requirement for a pitch-catch (i.e. two transducer) scan configuration, which is difficult to keep aligned. Moreover, the main goal of the inspection is to determine the origin of anomalous indications inside of the alloy, and the side scattering approach is lacking in this regard due to the uncertainty in the source of the anomalous backscattering, signals. Forward but is complicated scattering provides by alignment essentially constraints the same data as and appears to offer no definitive advance in detection capabilities. Hence, no significant effort was made to develop new NDE techniques for this section, due to the need for more comprehensive data based on existing ultrasonic (backscattering) approaches. The forward scattering results revealed no clear advantage over backscattering, but they could be significantly improved through the use of more sophisticated articulation the transmitter and receiver to map the backscattering, forward which scatter as a function has been thoroughly of incidence researched angle. throughout Unlike ultrasonic the last couple of decades both experimentally and theoretically [4, 5, 6, 7, 8, 9, 10, 11], forward scattering remains an open for further research. Useful information could be further acquired through the collection of data at multiple frequencies to provide more quantitative detail into the physical aspects of material inhomogeneities. Unfortunately, these developments would require an investment of resources that goes beyond the scope of the current 75 research. However, the samples developed for this dissertation and the research area will remain an open opportunity for future efforts. 4.2 Sample Development sets of samples have been developed Two characterization of subsurface to investigate the detection and in the titanium bar stock. Additionally, anomalies a preliminary sample set was constructed of 1" diameter Ti-6A1-4V bar, to assess whether or not the diffusion bonding parameters (i.e., 950 °C, 15 ksi, 4 hours) were adequate to join the metal. These 1" diameter samples were axially bonded and were available for destructive metallographic examination. After the initial trial bond samples were generated, two additional sample sets were constructed so that anomalous microstructure detection could be investigated from both the axial and radial directions. For data collection in the axial direction, sample coupons of the 2.5" diameter bar stock were rough cut into disks of approximately 1" thick, as shown in Figure 4.1. Five different samples CBA/CBA, of this orientation were to be made, AR/FBA/AR, and AR/CBA/AR including AR/AR, AR/CBA, (see Table 4.1 — Axially Bonded Samples, Appendix A). Following the preparation of the roughly cut disks, some of the disks were heat treated to arrive at the fine and coarse beta annealed condition. Of course, heat treating in air results in a thick scale on the samples which is easily removed, but worse, the exterior 76 a >. —“— a bonded sample axial bond coupons Figure 4.1 Schematic for the axial diffusion bonded Ti-6A1-4V samples. material of the annealed samples becomes case hardened due to the chemical reaction that takes place between the room air and the Ti alloy. Hence, prior to joining, this alpha case must be removed if the coupons are to be appropriately joined. The case removal was achieved by a combination grinding and lathe machine work. The axially bonded samples were further prepared by generating flat mating surfaces on the lathe and then lapping the surfaces samples to approximately 1 1m rms roughness. The next step is to thoroughly clean the samples to remove any traces of contamination, like fingerprints, grease or grime. Finally, the samples were electron beam welded and placed into the hot isostatic press (HIP) where the combination of time, temperature, and pressure results in a diffusion bond. The actual samples are shown in Figure 4.2. For data collection in the radial direction, which is the important direction form an industrial billet qualification standpoint, an additional sample set was constructed. These samples have three different diameters of inserts with three different microstructures, 77 which are diffusion bonded to the interiors of the bored cylinders. Figure 4.3 shows the schematic for the radially bonded samples. Insert sizes of 0.2", 0.4", and 0.9" were AR/AR Figure 4.2 CBA/CBA AR/CBA The five axially diffusion bonded samples. Hp po ‘ : leckedesh —- Figure 4.3 AR/CBA/AR AR/FBA/AR : ptt L : : ——_— radial bonds ; 4 wd: : | radial inserts: 0.2", 0.4", Schematic for the radial diffusion bonded Ti-6A1-4V samples. 78 5 0.9" generated with the AR, FBA, and CBA microstructures. These inserts were first heat treated, then machined on the lathe to snugly fit into the appropriate bore of the 2.5" diameter cylinders. Following the final machining and cleaning steps, the inserts were press fitted into the cylinders and then electron beam welded on the ends prior to placement in the HIP chamber. The actual samples are shown in Figure 4.4 (see Table 4.2 — Radially Bonded Samples, Appendix A). Figure 4.5 shows a metallographic results from the destructive evaluation of the 1" diameter AR/AR Clearly, this sample has some defects bondline which diffusion bonded may have ‘been sample. due to contamination, but as a result of finding defects the bonding parameters were modified to: 950 °C, 30 ksi, 4 hours (i. e. the HIP pressure was doubled) to ensure better consolidation across the interface. The HIP process clearly softens the elongated grains and grain Figure 4.4 The nine radially diffusion bonded samples. 79 Cae original axial direction Figure 4.5 Metallographic results comparing the original as-received microstructure with that of the HIP'ed sample. Both images were taken at 200x from the same orientation of the bar. colonies present in the original mill annealed structure, resulting in more equiaxed grains with a distribution of grain sizes centered at approximately 30 zm and no evidence of macrocolonies. Figure 4.6 shows a comparison of the metallographic results from a alpha case contaminated bondline versus a clean bond interface. In the case of the contaminated bond, the alpha case hardened surface that results from heat treating in lab air was not thoroughly ground off prior to diffusion bonding. This sample was the motivation for extensive grinding of the beta heat treated sample surfaces that followed to assure there was no contamination at the interfaces. 4.3 Detection of Anomalous Microstructure The internal structure of the diffusion bonded samples was examined using a 10 MHz, 3 in. focal length ultrasonic transducer in the pulse-echo mode. For the five 80 samples bonded in the axial configuration, B-scans were taken from several locations through the flat entry surface on the ends of each of the samples, which were ground and polished for flat and parallel opposing surfaces. In contrast, for the radially bonded samples, the B-scans were taken through the curved entry surface on the sides of each of the samples. B-scanning is a means of visualizing the actual ultrasonic response of the structure in the form of an image. The B-scan is taken as a line of data with each successive interval representing a separate, but adjacent, data point. Figure 4.7 shows a B-scan image as it would look without grayscale coding. The scan is simply a series of closely separated ultrasonic signals, which are all stacked on top of one another to represent a cross-section of the material inspected. One of the disadvantages of the B-scanning approach is that it is possible to generate huge volumes of data that can quickly overwhelm the computational resource needed to process the data. To get around this problem, only a few randomly chosen data files will be shown from each of the samples. However, if one were interested in reconstructing the scattering response over a significant volume of material, extensive computer memory and processing power would be essential. Generally, B-scans are encoded according to a color scale as depicted in Figure 4.8. Also shown is a line plot taken directly out of the B-scan to demonstrate the relationship between the B-scan image, the ultrasonic waveform, and the color scale. In all of the B-scans collected, the gain on the receiver was turned up to + 52 dB in order to observe the grain scattering signals. The high gain saturates the front and back 81 Peers: alpha case contaminated bond axial direction Metallographic results from dissimilar microstructures samples (i.e. AR/CBA) comparing an alpha contaminated bond with a clean bond. Both images were taken at 100x from the same orientation of the bar. Figure 4.6 f rn AS ad oe length (inches) . Vy Y DLS ne 15) 4 \ fer sn as. Sven NY [ON Np ot. i "y 5 PPP avy VPN F PAL NINAi pnt . INI POOL * . a NVA Nees LIBS LANL LNGect “ny AVN, PAA Noe, i tN f Mi om LOLA o NL Figure 4.7 A vd. Sy YC backscattered Smt LALO ultrasonic B-scan coding. 82 n EPALIVS LLG, PRL / ny NES POD PS SIONON DLN NN. wa ae tinct a oN pl Ot LBL RL ONLEL VDL INS SAAS AN LN shown . NG vars SPOON SOI NON without OP IIL I \ et NLSO J et ff OLIN t LE Loren the r advent NS of color B-scan 85.61 . line plot backscattered grain noise 8917 1123 Pixels 0 color scale just beyond the center front wall echo ~10ps Figure 4.8 An illustration showing the relationship between the B-scan image, the backscattered waveform, and the associated color scale. wall echoes on the analog-to-digital converter used to collect the data, so there is really no useful information from either of these locations. As such, the B-scan gates from which the rf data is collected begin slightly after the saturated front wall echo and extend approximately 11 1s into the material, which corresponds to more than half way through the thickness of the 2.5" diameter samples, as shown in Figure 4.9. Based on this approach, the backscattered rf signals can be used as a means of differentiating between the different microstructures. The B-scans are essential to the process of discriminating between the different microstructures because the statistical distributions of the amplitude data between the AR, FBA, and CBA microstructures overlap extensively. The primary 83 feature that allows us to detect the anomalous microstructure is the correlation length [1, 2], which is a measure of the degree of association in a series of signals and changes between the three microstructures chosen to manufacture the bonded samples. A B-scan represents such a series. The autocorrelation of a two dimensional image (the B-scan) provides a measure of length over which the signals in a series are closely matched in terms of the amplitudes and phases of the scattering centers. In the axially bonded sample set, only one scan configuration was used, as shown in Figure 4.10.a. In this case, the samples were scanned through the flat entry surfaces on either end of the samples. In the radially bonded sample set, two different configurations were used to investigate the optimal scan plan for detecting anomalous microstructure, as a+ 1.5+ | interface signal from it -0.5+ -1+ (+ 52 dB gain) jil ost 0 {| 0.9" CBA implant ! ihWa HN Mh. bali A Ah at FOV RUN TORR | all 4"I 10 TTT UCTYYIE 12 ets time (us) ere 14 YTetene ie saidy 16 rediarerrdiid 18 Wz B -scan gate 20 h iN [b, a back wall echc TOT front wall echo 2 Figure 4.9 An ultrasonic waveform which was collected through the curved surface on the side of the radially bonded AR/CBA/0.9" sample. 84 = scan line = ~~ axial scan NTT To eee LA radial scan axially bonded sample radially bonded sample (a) (b) Figure 4.10 The B-scan samples. configurations used for the two sets of diffusion bonded shown in Figure 4.10.b. In this case B-scans were collected through the curved entry surface in both the axial and radial directions. For both sample sets (axially and radially bonded) data was collected at two different focal lengths such that the effect of focusing on the surface versus focusing below the surface could be assessed. 4.4 Results and Discussion Since the receiver was set to nearly full gain, it is important to remove from each B-scan the part of the signal associated with ringing of the transducer. To do this, the first step is to collect a spatial average in the time domain using approximately 10 k signals. 85 This coherent signal is then subtracted out of each B-scan such that all that remains is the backscatter response of the material. Figure 4.11 shows an example of a raw data B-scan compared with one that has the reference signal subtracted out. This reference subtraction is a simple processing routine which can be performed in any spreadsheet software, but is necessary in order that the material response data is not confused with the response of the ultrasonic transducer. The remaining images throughout this section will all have had this simple processing step performed on the data. Additionally, the B-scan image data histograms have been equalized to provide the optimal contrast. The remainder of this section is broken up into two different parts, which address the axially bonded samples and the radially bonded samples. 4.4.1 Axially Bonded Samples All of these samples were bonded with adjacently cut coupons in a manner that maintained the original alignment to minimize the evolution of a bondline interface and any associated ultrasonic reflection. Figures 4.12(a) shows the axially bonded AR/AR sample, Figure 4.12(b) and 4.12(c) shows the AR/CBA sample and Figure 4.12(d) shows the CBA/CBA sample. The AR/CBA and CBA/AR samples are the same, except flipped over to observe differences in scattering due to the initial microstructure encountered first. In the case of sample CBA/AR, there is no evident bondline indication from the Bscans. Moreover, there is no indication of a change in character of the scattering between the two different microstructures in CBA/AR, unlike the case when the sound enters the AR microstructure first in AR/CBA. Each of the surface focused, 10 MHz B-scans from Fig. 4.12 are 11.3 ps in length, which corresponds to a one way path length of 86 i ; P; Be it +& >é a ce go 3 477.59 spatially averaged reference 73.82 0 Pixels 509 raw - reference cuss Figure 4.11 =. An illustration showing how the transducer reference signal is removed from the B-scan, leaving primarily only the material response. approximately 1.35". In each case, the B-scans started at approximately 1.5 ps into the sample to avoid the meaningless saturated front wall echo signals and extended through the thickness of the sample such that the bond interface is approximately in the center of each of the scans. The bond interfaces in Fig. 4.12 are indicated by discontinuities in the B-scan data, which are consistent with the predicted arrival time from the reflections. From the backscattering results of Chapter 3, it became clear that the axial direction of the mill annealed structure yields the least in terms of backscattered ultrasonic energy. 87 The diffusion bonded AR/AR sample is true to form in this regard, even though the HIP processing changes the microstructure from one of elongated grains and grain colonies to Figure 4.12 ~ AR/CBA CBA/AR (b) (c) (d) Enhanced ultrasonic B-scans using equalized histograms from the axially bonded sample series. AR/AR shows a clearly evident bondline signal. AR/BA clearly demonstrates the typical change in the scattering characteristics between two different microstructures. No indications of any kind are present in BA/AR. BA/BA has a strong bondline signal and clearly did not properly form a good diffusion bond. 88 one with essentially equiaxed grains and no apparent grain colonies. While a bondline signal is clearly evident in the AR/AR sample, the structure noise on either side of the interface has essentially the same character, both spatially and temporally. In contrast, the axially bonded AR/CBA sample has a distinct change in the character of the signal noise before and after the bond interface. This change in the character of the signal noise is not represented by the statistical distribution of the amplitude data, but by the correlation length of the ultrasonic backscatter. A of the histograms comparison nearly shows completely overlapping signal distributions between the two distinctly different regions of the AR/CBA sample. However, a comparison of the correlation lengths of the same regions demonstrates a dramatic change, which is representative of the change in the degree of randomness (of the crystallographic grain structure) between the two scattering fields originating from the two different microstructures. Additional B-scan images were collected from the AR/AR subsurface, rather defocusing provides B-scan surface focusing. These significantly stronger ultrasonic than, and AR/CBA with showed that the signals, thereby images backscattering allowing for easier observation of signal variations from the joined samples. However, this increased scatter intensity is not necessarily helpful in terms of defect detection because both the scattering and the defect indications, from imperfections at the bond interfaces, are equally intensified, without a net improvement to the signal to noise ratio. Image enhancements, performed to these like histogram images to bring equalization and out more detail microstructures. 89 edge detection, in signal can be easily differences between The images from Figure show 4.13 B-scans from the AR/FBA/AR and Figure 4.13 Hci Ver. ica ae at AR/CBA/AR sandwich bonded samples. Both images cover approximately 19 us of Enhanced ultrasonic B-scans from the AR / X/ AR sample series. Both the FBA and CBA implant images show indistinguishable characteristics, with clear indications of scattering changes between the joined microstructures. 90 data in the interior of each sample, which corresponds to a one way path of approximately 2.4", allowing both bond interfaces of the sandwich samples to be captured. The initial microstructure transitions are clearly visible, based on the scattering distinction between the AR and FBA/CBA bond interface. However, the second transitions (back to the AR microstructure) are undetectable, similar to the problem observed with the CBA/AR sample. The scattering characteristics do not reverse back to those typified by the AR structure upon encountering the second bond interface. Hence, there is really no way of distinguishing the CBA structure from the AR or FBA structure after the second interfaces. While the signal differences are clearly distinct before and after the first bond interfaces, there is no distinction ‘at the second These interface. interface secondary reflections do not stand above the background level of the scattering, but are partially visible due of the coherency of the ultrasonic reflection from the interface. This signal coherency at the interface second the make bonds marginally but visible, without knowing in advance that the structure is bonded, the detection would be quite difficult if not impossible. It is also surprising that there is very little distinction between the scattering fields from the FBA sample versus the CBA sample. Without using any sophisticated signal processing, one would have great difficulty in differentiating between the FBA and examination, CBA there structures are subtle sandwiched FBA and CBA based the on characteristic B-scan differences data. scan between data. the Upon closer axially bonded inserts. However, these distinctions are only realized upon viewing the scattering correlation lengths between the two data sets. Basically, the CBA correlation length was shown to be slightly longer than that of the FBA sample, which is probably due to the larger prior beta grain size in the CBA sample insert. 91 4.4.2 Radially Bonded Samples In this section, 10 MHz B-scan results are presented from data collected through the curved surface on the sides of the 2.5" diameter radially bonded samples. Three different sets of samples were generated, which have different sizes of diffusion bonded inserts (i. e., 0.9", 0.4", and 0.2") and different microstructures (i.e., AR, FBA, and CBA). B-scans were taken from both the axial and radial configurations with surface focusing. Figure 4.14 shows axial B-scans from the three AR/AR/X samples, where X is 0.9", 0.5", or 0.2". In these images, the bond interfaces for the 0.9" and 0.4" inserts are marginally visible in the backscattered data as indicated by arrowheads. However, for the 0.2" insert, there is no evidence of any kind of indication. Moreover, beyond the indications from the interfaces, there is no evidence of a change in the character of the scattering results. Hence, we can conclude from this data that these samples are all well bonded and there exists no overwhelming indication of any bonding problems. In Figure 4.15 the same sample (i.e. sample AR/AR/X) was scanned, except this time in the radial direction, using surface focusing at 10 MHz. In this case, the bondlines are clearly detectable for the 0.9' and 0.5" inserts, but again there is no clear indication of a bondline reflection or the presence on an anomaly for the 0.2" insert. Somewhat surprisingly, defocusing fails to clarify the presence of bondline signals. The defocusing appears to brings out stronger signals, but these same signals mask over the bond reflections that were evident with surface focusing. Hence, radial scanning with surface focusing appears to highlight bond imperfections, but with internal focusing the detection of bondlines is sub-optimal. 92 Bey. ~AR/AR/0 Q/a Figure 4.14 Basst Pcesecis / AR/0.5/ a AR/AR/0.9/a B-scans in the axial direction through the curved surface on the side of the AR/AR/X bonded cylinders focused at surface at 10 MHz. Enhanced ultrasonic B-scans from the AR / AR / X / x sample series, where X designates the diameter of the implanted anomaly, and x indicates the scan orientation (a - axial). ee, AR/ 0.2 Figure 4.15 Enhanced B-scans from the AR / AR / X / x sample series, where X designates the diameter of the implanted anomaly, and x indicates the scan orientation (r - radial). No indications of diffusion bondlines or anomalies are evident for the 0.2" diameter implant, but bondlines are clearly evident in both of the samples with indicated with arrowheads. 94 larger diameter implanted anomalies, as is presented for the AR/FBA/X In Figure 4.16 the same data as in AR/AR/X sample set, where X is 0.9", 0.4", and 0.2". In this case, regardless of the focusing, the axial imaging approach demonstrates no clear indications of bondline reflections for either the 0.9" or 0.4" anomaly inserts. On the other hand, there are clearly obvious changes in the character of scattering data for both of the larger implanted anomalies. Unfortunately, AR/FBA/0.2" has a severe bondline defects, which makes it difficult to comment on the detectability due to uncertainty regarding how much acoustic energy is propagating through the imperfection. Moreover, the ultrasonic ringing from the unbonded interface causes difficulty in determining the origin (in time) of the reflection. Further inspection of the 0.2" sample showed lack of bonding over virtually the entire implant. Defocusing appears to have no advantage in detectability for bondline imperfections or microstructure variations. In Figure 4.17, the B-scans were collected in the radial direction. Again, it appears that scanning in this orientation tends to preferentially bring out the bondline signals, but really does not provide any greater insight into the scattering perturbation caused by the microstructure anomalies. Moreover, scanning in this orientation with defocusing appears to wash out all indications of the bondline, except for the 0.2" anomaly, which obviously did not fully join, as indicated by the strong echoes. This apparent inability of defocused radial scanning to highlight bond anomalies is unfortunate, considering a similar approach is a standard practice for large (~10" diameter) billet qualification examinations. The defocusing technique clearly suffers from beam disturbances, tending to mask over the grain scattering , which is the basis of the microstructure characterization. 95 ER RR x, AR/FBA/2/a Figure 4.16 Enhanced ultrasonic B-scans from the AR / FBA / X / x sample series, where X designates the diameter of the implanted anomaly, and x indicates the scan orientation (a - axial). A strong bondline appears from the 0.2" implant, which indicates the bond did not properly form. The nature of scattering clearly changes in both of the images from the bigger diameter implanted samples, but neither show evidence of bondline signals. . 96 OESAR/EBA/.9/ Figure 4.17 Enhanced ultrasonic B-scans from the AR / FBA / X / x sample series, where X designates the diameter of the implanted anomaly, and x indicates the scan orientation (r - radial). A strong bondline appears from the 0.2" implant, which indicates the bond did not properly form. A bondline signal is evident from the 0.4" implant, but only marginally so for the 0.9" implant. No characteristic scattering changes are evident in any of these images. In Figure 4.18 data is presented from the AR/CBA/X sample set, where X is 0.9", 0.4", and 0.2". The axially collected B-scans again demonstrate marginal detectability of the 0.9" and 0.4" microstructure anomalies, but fail to show any indications from the 0.2" anomalies. Clearly, all the samples from this set fused in such a way that the bondlines are virtually invisible. The coarse microstructure of the CBA inserts present a less obvious target in terms of detectability compared to the FBA targets, based on these backscattering results. In Figure 4.19, the radial scanning preferentially delineates the bondlines, at the expense of characteristic variations associated with microstructure scattering. The autocorrelation length, which is expressed as: (4.1) A(t) = [f(OF(t-aat provides a measure of the similarity between functions (or waveforms in this case). The correlation lengths of coarse and fine beta annealed inserts and the mill annealed inserts have been calculated from the radially bonded samples (surface focusing), as shown in Figure 4.20. This parameter appears to be the only reliable indicator that allows key differences in the scattering to be described. The capability to distinguish between the different imbedded microstructure statistical amplitude distributions microstructure anomalies tend overlap to has the longest correlation is important, especially considering the significantly. The length due to the aligned 98 mill annealed features in the structure, while the correlation lengths are much shorter for the beta annealed samples due to recrystallization. AR/CBA/2/a Figure 4.18 Meaapne AR/CBA/4/a.—AR/CBA/9/a Enhanced ultrasonic B-scans from the AR / CBA / X / x sample series, where X designates the diameter of the implanted anomaly, and x indicates the scan orientation (a - axial). No indications of diffusion bondlines or anomalies are evident for the 0.2" diameter implant, but the characteristics of the scatter appears to change in both of the samples with larger diameter implanted anomalies, as indicated with arrowheads. 99 ee AR/CBA/.4/r Enhanced ultrasonic B-scans from the AR / CBA / X / x sample series, where X designates the diameter of the implanted anomaly, and x indicates the scan orientation (r - radial). Bondlines are marginally detectable for the 0.2: implant. The bondline from the 0.4" implant is clearly detectable, as is that of the 0.9". No characteristic changes in the scattering are evident in any of these samples. 100 - + = coarse beta annealed i\ ———~ fine beta annealed 4 . 0.168" —— mill annealed 60 5 40 | J 0.064" pre 0.047" 207") 0 .10 .20 30 40 50 .60 length (inches) Figure 4.20 Correlation lengths of the radially bonded samples in the axial direction. 4.5 Summary Diffusion bonded samples have been constructed to simulate incomplete processing anomalies in the Ti-6Al-4V alloy. These samples were bonded in both the axial and radial orientations to allow an assessment to be made regarding the detection of anomalous microstructure. For the samples bonded in the radial orientation, an assessment was made of the optimal scan configuration for the detection of simulated processing anomalies, but for the axially bonded samples only one scan configuration was used. A 10 MHz, 3" focused, 0.5" aperture ultrasonic transducer was used to conduct these experiments. B-scans were taken at two different focal lengths (i.e. at the surface and defocused 1" from the surface) to ascertain whether or not it is advantageous to focus below the surface. The ultrasonic results from the axially bonded samples clearly show it is possible to detect anomalous microstructure via B-scanning in the axial direction. For the radially bonded samples, the results indicate the 0.9" and 0.4" diameter microstructure 101 inserts are marginally detectable when scanned in the axial direction and more difficult to detect when scanned in the radial direction. Radial scanning tends to preferentially resolve the diffusion bondlines at the expense of scattering information related to the microstructure anomalies. On the other hand, axial scanning tends to bring out the characteristic differences in scattering between the microstructure inserts and the base metal. Probably the most significant finding is that the 0.2" undetectable in all cases, except the AR/FBA/0.2" diameter inserts were sample, which clearly did not fully bond during the HIP process. The correlation length appears to be a reliable parameter to characterize the microstructure differences in the implanted anomalies, but the statistical amplitude distributions overlap significantly. 102 Chapter 4, References: 1. Fundamentals of Ultrasonic Nondestructive Evaluation - A Modeling Approach, L. W. Schmerr, Jr., (Plenum Press, NY, 1998). 2. Ultrasonic Testing of Materials, eds., J. Krautkramer and H. Krautkramer, (SpringerVerlag, NY, 1990). 3. Ultrasonic Waves in Solid Media, J. L. Rose, (Cambridge University Press, 1999). 4. E. P. Papadakis, “Ultrasonic Attenuation due to Grain Scattering in Polycrystalline Metals,” Journal of the Acoustical Society of America, Vol. 37, p. 711 (1965). 5. E. P. Papadakis, “Ultrasonic Attenuation Caused by Scattering in Polycrystalline Media,” Physical Acoustics Principles and Methods, Vol IV, Part B, Ed. Mason, W. P., (Academic Press, New York, 1968) p. 269. 6. F. J. Margetan, R. B. Thompson, and I. Yalda-Mooshabad “Modelling Ultrasonic Microstructural Noise in Titanium Alloys,” Review of Progress in Quantitative Nondestructive Evaluation, Vol. 12B, eds. D. O. Thompson and D. E. Chimenti, (Plenum, New York, 1993), p. 1753. 7. B. Fay, “Theoretical Considerations of Ultrasonic Backscatter,” Acoustica 28, p. 354 (1973). 8. C. B. Guo, P. Holler, and K. Goebbels, "Scattering of Ultrasonic Waves in Anisotropic Polycrystalline Metals," Acoustica 59, pp. 112-120 (1985). 9. S. Hirsekorn, “The Scattering of Ultrasonic Waves in Polycrystalline Materials with Texture,” Journal of the Acoustical Society of America, 77(3), (1985). 10. J. H. Rose, “Ultrasonic Backscattering from Polycrystalline Aggregates Using TimeDomain Linear Response Theory,” Review of Progress in QNDE, Vol 11, eds D. O. Thompson, D. E. Chimenti (Plenum, New York 1992) p. 1677. 11. J. H. Rose, “Ultrasonic Backscatter from Microstructure,” in Review of Progress in ONDE, Vol 11, eds D. O. Thompson, D. E. Chimenti (Plenum, New York 1992) p. 1677. 103 CHAPTER V THE INFLUENCE OF TEXTURE AND PHASE DISTORTION ON ULTRASONIC ATTENUATION IN Ti-6Al1-4V 5.1 Introduction In this chapter, a methodology to characterize microstructure and study the influence of texture and phase scattering on ultrasonic attenuation in Ti-6Al-4V was was also demonstrated. developed and developed to map the phase A high resolution experimental of ultrasonic and magnitude waves capability transmitted in the titanium solid. The advancement presented in this work is provided by laser detection of the ultrasonic energy over a microscopic aperture of approximately 50 um. The system is built around a computer controlled scanner and a confocal Fabry-Perot interferometer, which uses a diode pumped Nd:YAG laser as a light source. Wave propagation in the axial and radial directions of a 2.5" diameter bar of highly textured (mill annealed) Ti- 6Al-4V was investigated in this study. The work was motivated by the observation of unusually high apparent attenuation in the axial direction of the as-received bar, thought to be mainly associated with phase distortion rather than actual energy loss. The current phase mapping results, 104 using a focused show spot, laser high relatively wavefront distortion and more nonuniform distribution of the transmitted energy in the axial direction. The contribution to attenuation associated with phase cancellation loss was also investigated. These measurements show the laser detected attenuation to be substantially lower than the piezoelectrically measured attenuation. However, even the relative phase insensitivity of focused laser detection approach clearly indicates the attenuation to be strongest in the axial direction. This work demonstrates the orientation dependence of attenuation stems from scattering effects associated with texturing and the elongated macroscopic grain structure in the mill annealed Ti-6A1-4V bar generated during processing, which may also affect diffraction and divergence. beam This chapter provides a foundation for nondestructive evaluation to observe worked microstructure, indicated by structure bands, and may provide a process control method for alloy forging and ingot coarseness. For Ti-6AI1-4V, the manufacturing processes used to fabricate stock materials (e. g. bar, billet, plate) tend to impart a preferred crystallographic orientation due to the restricted nature of mechanical evidence of macroscopic slip [1], leaving the material with easily measurable anisotropy. For example, 5 MHz shear wave birefringence measurements conducted on 1.5"-thick sections of 2.5"-diameter mill annealed Ti-6A1-4V bar demonstrate only about 1% variation in velocity for wave propagation in the axial direction, while in the radial direction the difference in velocity between the slow and fast pure shear modes is nearly 5%. In addition, compared to the axial direction, the 10 MHz longitudinal wave velocity is approximately 2% faster in the radial direction. The observed velocity anisotropy is consistent with the development of a texture in which the oriented randomly basal normals of the alpha phase (hexagonal symmetry), which constitutes more than 90% of the alloy by volume, tend to preferentially rotate to the transverse direction of the bar during the course of processing. Ultrasonic measurements, using piezoelectric detection, show significantly higher attenuation in the axial direction as compared with the radial one. Similar results have been described for austenitic stainless steels in which the material develops a columnar grain structure due to solidification, rather than hot-working as is the case for titanium alloys [2, 3]. Typically, attenuation is considered to be based on energy decay of the propagating ultrasonic waves, which is proportional to length and mainly affected by absorption and scattering. However, for some polycrystalline titanium alloys and austenitic stainless steels comprised of primarily hexagonal symmetry, phase distortion of the propagating wavefront and the resulting signal diminution at the receiver has been speculated as another source of loss. This phase cancellation loss at the receiver, rather than actual energy-based losses, is thought to be partly responsible for the attenuation behavior observed in the axial direction of mill annealed Ti-6AI-4V. Recently, Panetta et al. demonstrated that fluctuations in the amplitude of backwall echoes observed as the ultrasonic beam is scanned over different samples of titanium alloys are not due to energy losses [4]. Their work supports the idea that phase distortion of the transmitted beam contributes to the higher attenuation values in the axial direction of the billet. Using a small diameter "point" receiver, they also revealed that the scattering and absorption as deduced from the transmitted energy are indeed anisotropic 106 in Ti-6Al-4V billets, but are smaller and isotropic in Ti-5Al-2Sn-2Zr-4Mo-4Cr (Ti-17) [5]. was finding This both fact that the despite observed showed alloys distinct characteristics of macroscopic anisotropy in the two orientations studied. Similar phase mapping studies have been undertaken using small aperture hydrophone receivers [6] and array transducers [7]. In this chapter, we use focused detection laser spot thereby minimizing the phase sensitivity, allowing high resolution imaging of structure-induced wavefront distortion. Spatial uniformity of the transmitted ultrasonic beam's pressure distribution has been shown to play an important role in the measurement of absorption and scattering [8]. For phase-sensitive receivers, impinging the aperture electrical signal. any fluctuations in the are instantaneously averaged, Conversely, the phase-insensitivity phase of the pressure field resulting in cancellation of the offered an by acoustoelectric receiver, constructed from a single crystal of the piezoelectric semiconductor cadmium [9], demonstrates sulfide more reliable estimates of quantitative intensity dependent parameters like attenuation, at the expense of reduced resolution and sensitivity. Medical ultrasonic characterization of tissue samples is another area where the development of phase distortion in the propagating wavefront has important technological ramifications. Diagnostic analysis of tissue inhomogeneities generally involves imaging provided by phased array ultrasonic systems. Tissue-induced wavefront distortion can severely degrade the quality of medical ultrasonic 107 imaging and often correction techniques are needed to account for phase aberrations to generate higher quality images (10, 11]. A transmitted focused laser incident beam on [12]. Besides acoustic radiation a suitable surface the obvious can advantages used be to detect offered by laser interferometry (e. g. noncontacting, elevated temperature usage, complex geometries) for detecting ultrasonic waves, diffraction limited apertures of sub-wavelength dimensions can be easily achieved via focusing of the incident laser beam. In contrast to other approaches used to restrict the aperture, such as applying a pinhole [13] or needlelike stylus tip [14] to achieve high resolution, laser interferometry offers the best sensitivity when the laser is concentrated to the smallest spot size [15]. However, in some highly backscattering materials like that of the transverse direction in mill annealed Ti-6Al-4V, the grain noise level significantly increases as the spot diameter of the laser decreases [16], which would obviously have adverse affects on the detection of small defects. For laser detection, as the incident beam is scanned over the optically reflective surface of the titanium sample undergoing surface displacements, the reflections are collected and passed to the interferometer. The surface displacements generated by the propagating ultrasonic waves cause the reflected laser beam to be frequency modulated, resulting in a Doppler shift of the carrier frequency interferometer demodulates corresponding to the the acoustic propagating in the sample. reflected light particle velocity [17]. The confocal to reconstruct (or pressure) Fabry-Perot a time-domain signal elastic waves of the The main goal of this chapter is to explore the use of laser interferometric detection to better understand the effects of macroscopic texture and phase perturbation on the attenuation of longitudinal waves in the mill annealed Ti-6A1-4V alloy. To begin, we will review the basic ultrasonic properties as a function of orientation in the mill annealed Ti-6Al-4V bar, measured with standard piezoelectric ultrasonic detection. We then discuss the operation of the laser interferometric detection experiment, detailing the differences in hardware components used to map the phase and magnitude of the transmitted ultrasonic waves. Next follows a presentation of the experimental results and discussion, demonstrating the structure-induced phase jitter in the axial direction to be significantly higher and more nonuniform than that of the radial direction. Finally, in order to assess the attenuation loss which develops due to phase cancellation, we compare attenuation measurements performed with a conventional wide-aperture piezoelectric receiver versus laser-based detection, using a microscopic spot size. 5.2 Ultrasonic Properties of Mill Annealed Ti-6A1-4V Samples of the mill annealed Ti-6A1-4V alloy were cut into 1.5"-thick sections with flat, plane parallel surfaces such that the angle between surface normals and the axis of the bar were at 0° (axial), 30°, 45°, 60°, and 90° (radial). Ultrasonic measurements of longitudinal and shear wave velocity, attenuation, and scattering (forward and backward) were conducted on the samples. The microstructures of the two key sample orientations are shown in Figure 5.1. The axial orientation of the Ti-6Al-4V bar shows a relatively equiaxed and random grain structure, while that of the transverse shows highly aligned colonies of elongated grains. The longitudinal ultrasonic velocity was measured using the standard pulse-echo- overlap technique at 10 MHz, consistent with guidelines specified in the literature for measuring ultrasonic velocity [18, 19]. The longitudinal wave velocity results are shown in Figure 5.2. The data clearly shows that the wave velocity is highest in the radial direction of the bar (see Table 5.1 - Orientation Dependence of Longitudinal Velocity, Appendix A). A difference of about 2% is observed in the average velocity between the axial and radial directions. The longitudinal data also shows a linear increase in velocity as the propagation direction changes from axial to radial. The velocity is associated with mechanical stiffness, therefore the material is stiffer in the radial direction. < ER hes as OWE Figure 5.1 ad i The as-received (mill annealed) condition of Ti-6AI-4V bar stock from surfaces of the (a) axial (0°) and (b) transverse (90°) samples. Both microstructures are shown at 200x. 110 longitudinal velocity (m/s) 30 45 60 orientation angle (degrees) Longitudinal wave velocity as a function of orientation for 1.5"-thick mill annealed Ti-6AI-4V samples. Figure 5.2 Measurements of the shear wave velocity provide more details on the macroscopic texture of the propagation structure direction. as these Shear waves waves can for the same to the surface of a sample will be polarized incident normal differently generally split into two transverse waves vibrating at right angles to each other and propagating at different velocities, due to birefringence [20, 21]. This effect is even more dramatic in single crystals than in anisotropic polycrystalline metals due to the stronger directional dependence of velocity. The shear wave velocity data of Figure 5.3 shows significant velocity variations between the differently oriented titanium alloy samples. For each sample, several measurements were taken at two different shear polarization angles consistent with the slow and fast pure shear modes. We can see from the average 3350 + Cislow mode 3300 + ca ES ba 60 | Pd 90 v4oS Loe) } ch 3200 aah oS Oo | 1 — 3150 + ww shear velocity (m/s) (fast mode 3050 + |x 3000 ba 45 Tite, 30 ° orientation angle (degrees) Figure 5.3 Shear wave velocity annealed Ti-6Al-4V. as a function of orientation for 1.5"-thick mill shear velocity data that the primary factor affecting the velocity is the presence of a preferred crystallographic orientation. The most dramatic evidence of this influence is illustrated in the 90° sample, with the difference between the slow and fast pure modes of 4.8%. We also see a steady increase in the birefringence delta (difference in velocity between the slow and fast pure modes) as the wave propagation direction goes from axial to radial, with an axial birefringence delta of only about 1%. This small degree of mechanical anisotropy is evident in the axial direction because the texture is slightly offaxis and varies a little for different locations along the bar. Hence, the shear velocity is minimally dependent on the polarization direction for propagation in the axial direction; however, for propagation in the radial direction, the shear polarization has a major influence on the wave velocity (see Table 112 5.2 - Orientation Dependence of Shear Velocity, Appendix A). Wave propagation in the radial direction always gives rise to the fast pure mode when the particle displacements are perpendicular to the elongated macroscopic inhomogeneities, while the slow pure mode always occurs when the particle displacements are parallel to the aligned inhomogeneities. Clearly, the particle oscillation of the fast pure shear mode is consistent with the preferred orientation of hexagonal basal normals, i. e. in the principal direction. This finding also appears to make sense in light of the longitudinal velocity results, showing faster speeds for waves traveling perpendicular to the axially aligned macroscopic structural features. Attenuation measurements were also conducted on the same set of 1.5"-thick samples. The ultrasonic attenuation coefficient is broadly defined as the rate of decay of the measured signal as a wave propagates through a material. The apparent attenuation can be significantly different depending on how the signal is measured (e. g., phasesensitive versus phase-insensitive measurements). Attenuation loss is defined as the ratio of two amplitudes and expressed in logarithmic units, neper or decibel. In some cases, the loss can occur locally as a result of interaction with a material discontinuity. Similarly, losses could originate from reflection and transmission at an interface, internal voids or inclusions, scattering at a rough surface, or from microstructural changes. There are other losses occurring over a given distance as the wave propagates in the medium, which are not necessarily proportional to the travel distance. Such losses are usually associated with beam divergence. The primary loss mechanisms investigated in this chapter are those due to the presence of texture and orientation. There are two major attenuation mechanisms considered important for ultrasonic materials characterization. The first is absorption, which converts acoustic energy to heat via viscosity, relaxation, heat conduction, elastic hysteresis, ete. The absorbed energy is irreversibly lost from the acoustic field and is dissipated in the medium. While absorption is difficult to separate from scattering in terms of a practical measurement, all the obvious indicators seem to point to the fact that the absorption is virtually the same, regardless of orientation in the mill annealed alloy. The second important attenuation mechanism is scattering, which converts the energy of the coherent beam into incoherent, divergent waves as a result of interaction with inhomogeneities in the material. The scattered energy is not necessarily lost as part of it can be recovered by the same transducer used to transmit, as backscatter. Scattering not only reduces the amplitude of the coherent signal, but also gives rise to an incoherent material noise which further limits the detectability of defects. Fortunately, the same incoherent noise that hinders our view to the development damage of low-level or defects also provides a source of information for ultrasonic characterization of grain structure [22, 23, 24]. The attenuation was measured underwater using a standard 10 MHz, 0.5"- diameter unfocused transducer, driven with narrow-band tone-burst excitation to allow easy adjustment of the frequency. The orientation dependence of the attenuation is shown in Figure 5.4 as a function of frequency (see Table 5.3 - Orientation Dependent Attenuation Loss, Appendix A). For these measurements, no effort was made to account for beam divergence losses since all the samples were of the same thickness and because we were mainly interested in making a simple comparison of relative attenuation 114 uO deg 16, x30 deg 14- .45 deg _ 124 960 deg 3~ 191 *90 deg 7 E : ZB > 6+ t 4t : ff ; i = & & § & | |5 i Hf 4 2+ 0 5 t t t t 6 7 8 9 10 t t 11 12 — 13 frequency (MHz) Figure 5.4 Orientation dependence of the average attenuation loss versus frequency for 1.5"-thick mill annealed Ti-6A1-4V samples. between the various sample orientations. All the immersion attenuation measurements were done with pulse-echo in the far-field of the transducer. The distance between the normally incident transducer and the front surface of each specimen was set to 3". A function generator and power rf amplifier were used to excite the transducer at a given frequency. For each frequency, the attenuation was measured by taking the Fast Fourier Transform (FFT) of the signal from the first backwall echo and subtracting the FFT of the second backwall echo. The difference spectrum was then evaluated at the excitation frequency (the role of the FFT processing was to average over the length of the signal, rather than to provide spectral information). Because the 115 ¢ signal amplitudes tend to have some modest spatial variations in the mill annealed alloy, ten different locations on each were repeated for approximately the measurements sample. The excitation frequency was then changed and the measurement repeated for a range of frequencies between 7 and calculated We 12 MHz. the impedance losses solid / water interface using a contributing to the attenuation based on the Ti-6A1-4V density of 4.4 grams/cc and a velocity of 6180 m/s, yielding a value for the reflection coefficient, R, of .9 or R* ~ —2 dB. This loss was accounted for in the results. The 0° sample (axial direction) clearly has the highest attenuation, followed by the 30°, and 45° samples, respectively. The 60° and 90° samples completely overlap in attenuation, considering the scatter in the data. The error bars of Figure 5.4 are based on + one standard deviation from the mean, representing the data scatter. The 45° sample showed very little scatter in the data (< + 0.3 dB), but the 0° and 30° samples showed significant scatter ( up to about + 2 dB) with increasing scatter at the higher frequencies. Moderate data scatter, on average of about + 1 dB, was observed in both the 60 and 90 The piezoelectrically measured degree samples. clearly orientation dependent. What be attributed to attenuation can we losses would coherent "plane wave" like to know from phase measurements were incurred attenuation is much of this perturbations of the conducted in the is how transmitted signals. Forward and backward scattering also titanium alloy samples for the axial and radial directions. Backscatter measurements were done at normal incidence, using a 10 MHz, .5"-diameter transducer with a 3" focal length. 116 The transducer was focused on the front surface and the backscatter was measured in based pulse-echo on a spatial average over approximately 1" a x 1" This area. measurement is relatively simple using a digital oscilloscope. The first step is to be sure the sample is parallel to the scan plane of the transducer such that the signal does not drift in time during scanning. Next, we remove the coherent component by taking a global spatial average in the rf and subtracting this averaged signal from each independent rf backscatter signal. What is left after subtracting the coherent part is squared and spatially averaged. Once the summed average runs its course, the final step is to take the square root of the spatially averaged backscatter response. in Figure returned as a function of time is shown Dependence Appendix of Backscattering, A). The average 5.5 (see Table Clearly, the significant backscatter is from the radial direction in which backscatter signals 5.4 - Orientation with any the incident waves are only sample traveling normal to the elongated grains and grain colonies. In each of the other samples, the bulk of the scattered energy is directed away from the backscatter path. The backscatter in samples cut at 0, 30, 45, and 60 degrees to the bar axis is negligible, but the overall scatter is not. For example, Figure 5.6 shows schematically what is happening for scattering in the 0 and 30 degree samples. Here, the forward and side scattered acoustic energy is equally, if not more important, to the backscatted energy for the purpose of measuring attenuation. Using a large-aperture phase-sensitive transducer, the effect of scattering is basically all the same insofar as the attenuation is concerned as part of the energy is removed from the coherent beam, scattering is detected or not. 117 regardless of whether the incoherent average backscatter (a. u.) sample (degrees) time (us) Average backscattering intensity as a function orientations of mill annealed Ti-6Al1-4V. Figure 5.5 of time in different Forward scattering measurements were performed in through-transmission using a 10 MHz, 0.25" diameter contact transducer, mechanically coupled to the backside of the sample to transmit the acoustic energy. A 10 MHz, 0.5"-diameter focused transducer was placed over the opposite surface to receive the transmitted scatter. Using this approach, the forward directed incoherent scatter can be observed between the first coherent transmission and the first subsequent coherent echo. By scanning the receiver over the forward scattered field, we can assess the magnitude and divergence of forward scattering. A simple way to view forward scatter is by taking a C-scan, gating on the amplitude modulated incoherent noise. Figure 5.7 shows the forward scattering C-scan results, using a 1 ps gate width, after averaging several different images with slightly different gate placements. Probably the most notable aspect of these results is the overall axial ieee transducer : axial transducer direction a direction / J 1 | | ; | : WW | f / iwi L\A Al. , wave direction primary scattering direction Y y 30 degree sample 0 degree sample higher / ge forward scatter . Figure 5.6 2 Schematic diagrams showing the effect of the macroscopic structure on scattering for the 0 and 30 degree samples. divergence of forward scatter in the axial direction. Divergence is inversely proportional to average grain size [22, 25], that is why wave propagation in the radial direction results in essentially the same divergence width in the horizontal direction (from Figure 5.7.b) as seen for the axial direction. The forward scattering is more divergent in the axial direction, but the amplitude of scatter signals in the radial direction is observed to be at least as strong to those of the axial direction. Intuitively, the combination of strong backscatter and relatively strong forward scatter in the radial direction may logically lead to the assumption that the radial attenuation should be higher, but this is clearly not the case. For radially directed wave propagation, due to the normal orientation of the elongated grain colonies, essentially all scattering is concentrated in the backward 119 Wie jaxial direction (b) . (a) C-scans of the forward scatter in the axial (a) and radial (b) directions, taken over a 1" x 1" area at 10 MHz with a 0.5"-diameter, 3"-focal length receiver, focused on the surface. Figure 5.7 and forward directions. Since the scattering is very weak in all but these two specific directions, the coherent wave attenuation is actually very low. Clearly, the total scattering cross-section must be considered to assess the orientation dependence of the coherent wave attenuation. This is an area in need of further investigation, especially the forward scattering, if we are to fully understand the higher attenuation losses in the axial direction of mill annealed Ti-6AI-4V. In the next section, we move to the laser-based ultrasonic detection experiment to discuss magnitude and phase disturbances on attenuation. 120 mapping and the influence of phase 5.3 Laser Detection of Ultrasonic Phase Distortion A system was developed to image the phase and magnitude of transmitted ultrasonic waves using a computer-controlled scanner in conjunction with a Fabry-Perot interferometer. Figure 5.8 shows the schematic diagrams of the experimental systems. The systems include a unique combination of analog and digital components, allowing the measurements to be made without extensive signal processing. The phase mapping system is configured to reconstruct the shape of the transmitted wavefront, including distortions which develop as the waves propagate. Several different sample thicknesses, including 0.25", 0.5", 1.0", and 1.5", were examined in the axial and radial directions. Unfortunately, phase mapping of the thinner samples (< 0.5") revealed some inherent near-field unevenness that we decided to exclude from this chapter. A 0.25"-diameter 10 MHz contact transducer was used to transmit the acoustic energy, which has a near-field transition length (NV = a’/ A) of about 16 mm (0.63"). This transducer was mechanically mounted to the backside each of the inspected samples to maintain constant acoustic coupling. The transmitter generates a tone-burst, derived from the combination of a function generator (HP 3314), capable of delivering a sine-wave burst at up to 20 MHz, and a power rf amplifier to boost the tone-burst signal by approximately 50 dB. The power amplifier output excites the transmitter which launches a burst of energy into the sample. The sample, with the attached transmitter, is fixed to a bracket on the x, z scanner and the laser is held stationary. The transmitted signal is detected on the front-side of the sample by a laser beam which passes the Doppler shifted carrier signal to a 100-MHz-bandwidth Fabry-Perot interferometer for demodulation and Nd:YAG Laser 10X microscope objective Transmitter X, Z Scanner on vertical stand ly, 4, feyeS O Fiber-optic cable Nw Computer : Collecting Fabry-Perot lens Interferometer Function Receiver generator (5072 PR) (HP 3314) RF amplifier (DG535) bd | — trigger in [ “| t : = = Analog oscilloscope (aN gate out signal in coerge Ane. Digital oscilloscope} 9310 AM) Signal averager (SR 280) Avg signal Delay generator eet Digital oscillosco Function generator (HP 3314) - ~_t {| (9310 AM) A rt j— Nd: YAG Laser 10X microscope objective Transmitter “ey, Fiber-optic cable Nw Fabry-Perot Interferometer (5072 PR) O Computer Collecting lens unction RF amplifier generator CHP 3318) RE Signal averager Del elay generator avg signal (SR 280) YE (DG535) out triggerin TR Digital oscilloscope X, Z Scanner on vertical stand A Ona Functi . Receiver Sg id - GO . be bf a RPL] |] signatin — sateout i ANB i Vertical signal out Digital oscillosco |e | fee] ES 9310 AM) | | Oscilloscope / Spectrum analyzer Gate signal marked trigger] mow it J in [———"_ | }} b) Figure 5.8 Schematic diagrams of the experimental set-ups used to map phase (a) and magnitide (b). reconstruction of the time-domain response. The laser beam is focused on the surface of the sample through a 10x microscope objective to make the spot size microscopic. The focused laser spot size on the surface was estimated, assuming a Gaussian beam, based on the following equation: (5.1) d(z) = dy)[1+ (442/ adj)? }” is the diameter of the focused laser spot and zis the distance from the waist of where d the laser beam [26]. The laser has a wavelength (1) of 532 nm and a divergence angle (0) of about 2.2 mrad. The output beam is assumed to be Gaussian with an initial beam diameter (d;) of 320 um and a waist diameter (dy) of approximately @=44/ nd) ~2.2 mrad). Using a value of 16.9 mm 307 um (from for the focal length of the 10x. microscope objective and assuming the objective was focused to within about 1 mm of the actual focal length, we estimate the spot size of the focused laser light to be from 25 to 50 um. The phase disturbances are measured by a precision time-to-voltage converter, assembled from general-purpose laboratory instruments. The titanium samples were finish polished with 9 pm diamond paste to yield nicely reflecting surfaces. The laser light reflects off of the shiny titanium alloy surface undergoing displacements and is gathered by a collecting lens which focuses the light onto a 400 .m-diameter fiber-optic 123 cable. This cable passes the light to the interferometer which is operated in reflection mode. The time-domain signal output from the interferometer is then routed to a receiver (Panametrics 5072 PR) to control the gain and then to an analog oscilloscope which has a trigger delayed output and pre-triggering capabilities. Pre-triggering allows local triggering on the waveform of interest. The delayed trigger output, which is dictated by the pre-trigger, is used to synchronize the boxcar averager with the phase variation of the interferometer output signal. A single cycle triangle wave is input to the boxcar averager to discriminate the phase variations. With this experimental set-up, any movement in arrival time (phase) of the interferometer signal simultaneously moves the gate on the boxcar averager and changes its corresponding voltage input according to the linear portion of the triangle wave. The boxcar gate is generally centered on the middle zeroof the triangle crossing wave. With this approach, variations in the phase of the interferometer signal, detected as the laser is scanned over the specimen, correspond to a linear change in voltage. To increase the signal to noise ratio in the images, we usually decrease the scanner speed to about 10 s per line (200 points per line) and increase the number of signal averages performed on the boxcar. The sensitivity of the phase imaging system is adjusted by manipulating the slope and amplitude of the triangle wave input to the boxcar averager. Obviously, higher frequency triangle waves yield steeper slopes and therefore provide higher sensitivity. Hence, a variation in phase as indicated by movement of the boxcar gate, can have dramatic or slight corresponding voltage change, determined by the characteristics of the triangle wave. 124 For magnitude imaging, the interferometer output signal is hardware gated with a stepless gate on the signal of interest and then introduced to a spectrum analyzer. The spectrum analyzer is tuned via a local oscillator to the appropriate frequency (generally the same frequency as the tone-burst signal) and the instrument's 3 MHz bandwidth determines the length and shape of the signal output. This output is a bell shaped curve whose amplitude corresponds that of the detected interferometer signal over the 3 MHz frequency range. Again, the boxcar averager is used to sample and hold the peak of the bell curve, except under normal triggering synchronization. The averaged output from the boxcar, which is proportional to the strength of the gated signal in the narrow frequency band centered around the chosen nominal measuring frequency, is fed into the analogdigital converter for sampling and digitization. The computer controlled scanner always generates 200 x 200 array images, regardless of the scan size (from 0.1" x 0.1" up to 2" x 2") that are digitized with 16 bit resolution. The C++ computer program we use to generate the images was written do some modest signal averaging internally, but we also generally over-sample extensively on the boxcar averager (100 or more measurements per pixel) and reduce the speed to increase the signal-to-noise ratio of the phase and magnitude pictures. While this scanner is modest in terms of the sophistication of the analog to digital conversion, it is versatile, allowing data to be collected from a number of different experimental configurations. 5.4 Results and Discussion The objective of phase imaging is to map the shape of the transmitted ultrasonic wavefront. The transmitter is mechanically fixed on the backside of the titanium alloy sample. The waves are launched from the transmitter and eventually begin to diverge after the near-field / far-field transition zone. As the waves diverge, the shape of the wavefront becomes curved due to diffraction and beam spreading. This wavefront curvature is especially apparent in points laterally displaced from the direct line-of-sight transmission axis. The phase lag increases for further off-axis locations on the sample, as shown schematically in Figure 5.9. The phase mapping system accommodates the sometimes very large phase variations over a given area simply by restarting the grayscale after every 2x phase shift. Hence, at every 2m phase transition a fringe is apparent. The development of fringes is consistent with the discontinuous presentation of phase, folding back after every 2n increment. It is possible to develop an algorithm to reconstruct the waveform in the absence of fringes, but this would not change the information held in the image, only reshape it. Moreover, in this paper we are mainly interested in high- resolution imaging of phase disturbances in the directly transmitted lowest order fringe. Figure 5.10 shows some examples of phase and magnitude imaging. These images were taken at 9.7 MHz through 0.5"-thick samples, cut to allow access to the axial (images a and b) and radial directions (images c and d) in the mill annealed material over a 1" x 1" area centered on the transmission axis. Both phase images were taken under identical scan parameter settings and reveal multiple 21 phase transitions due to the curvature of the propagating wavefront. Darker shades of gray in the phase images represent a phase lag. In this case (0.5"-thick samples) the first transmitted wave burst arrives at the detector prior to the near / far-field transition zone. As such, we see a highly 126 fundamental fringe (all data within a 27 phase field) Lo | subsequent NRA ff rw nna’ 4 WWMM 4 iA LALS- shifts J. A AA —~S = UV position aml : : Y MIATA ’ VA VW A AA MAA TNT NT NT MAA Av A A A A ae — LI fA > time LJ from oscilloscope mounted transmitter (b) (a) Figure 5.9 Schematic diagrams showing (a) three regions on the specimen over which the phase map is repeatedly folded back by 2x and (b) the variation in arrival time at different positions across the wavefront. collimated wavefront that dramatically lags in phase outside of the center fringe, thus the large number of fringes. Also evident is the near-field effect of phase lag in the centrally transmitted zone of the lowest order fringe. The corresponding magnitude images (Figure 5.10.b and 5.10.d) which were both taken with the same scan settings, are shown next to the phase images. Here, lighter shades indicate higher amplitudes. Figure 5.11 shows a similar effect to the wavefront distortion we are interested in observing in the mill annealed titanium samples. In this case, we mapped the phase through a 0.2"-thick Ti-6Al-4V beta annealed (~1040° sample, which was vacuum encapsulated in glass and C) for an exaggerated period, then slowly cooled to yield very 127 coarse lamellar «a + B grain colonies. In some places these colonies extended through the thickness of the sample. Here, we used a 0.5"-diameter transmitter to increase the interaction volume in hopes of encountering a grain colony with significantly different (b) (d) Figure 5.10 Phase (a) and corresponding magnitude (b) images taken at 9.7 MHz, through 0.5"-thick samples of the mill annealed Ti-6A1-4V from the axial direction, covering a 1" x 1" area. Same images are shown for the radial direction (c), (d). 128 (b) (a) Figure.5.11 Phase maps taken through a 0.2"-thick Ti-6Al-4V sample which had been heat treated to generate a coarse widmanstitten microstructure with very large lamellar o + B colonies. Phase images taken at (a) 7.5 MHz and the same region at (b) 9.7 MHz. wave propagation characteristics to the surrounding structure. The upper middle part of the center fringe shows the desired effect. Phase images were taken at two different frequencies (7.5 and 9.7 MHz) using the same settings as those from Figure 5.10. Clearly, in Figure 5.11 there is a region in the sample that lags significantly in phase from the surrounding material. At 9.7 MHz, the shorter wavelength causes the grayscale to reinitiate in the center of the anomalous colony, indicating the arrival time delays by more than a full cycle at that frequency. Similarly, Figure 5.12 shows the phase and magnitude images for a 1.5"-thick mill annealed Ti-6AI-4V sample in the axial direction (a, b) and radial direction (c, d) using the 0.25"-diameter transmitter at 9.7 MHz. In this case, the first transmitted waves are in the far field of the transmitter. These phase images are representative of the ultrasonic wavefront distortion in the axial and radial directions 129 through 1.5". In the axial direction, the phase perturbation was severe enough to cause the central fringe to be significantly distorted from the generally observed concentric ring pattern, as seen in the radial direction. Again, the corresponding magnitude images are shown next to the phase images. Figure 5.12 Phase (a) and corresponding magnitude (b) images over a 1" x 1" area through a 1.5"-thick sample of mill annealed Ti-6A]-4V in the axial direction at 9.7 MHz. Same are shown for the radial direction (c), (d). 130 So far, we have shown representative data from 0.5"-thick and 1.5"-thick mill annealed samples, with neither case being ideal for the generation of phase images due to the presence of near-field artifacts or too much phase variation. Figure 5.13, a and b shows the raw phase images from a 1.0"-thick sample in the axial and radial orientations, respectively, using the 0.25" transmitter at 9.7 MHz. In this case, we optimized the sensitivity to see the best contrast in the center fringe of the image. The scan settings were identical for both the axial and radial samples. Each of these scans were taken over a 0.4" x 0.4" area with a 0.002" stepping increment. To allow easier manipulation of the data, we stripped the first 27 phase transition from the phase images. We also filtered the data to remove low spatial frequency phase variations associated with beam divergence. The resulting high-pass filtered images are shown in Figure 5.13, c and d. From the statistical distribution of the high-pass filtered data, we measured approximately 30% higher phase variation in the axial direction, as compared with the radial direction, for a 130 x 150 pixel array covering the center portion of data in each the magnitude of these high-frequency image. While modest, only up to about 20 degrees, nonuniform in the axial direction. the phase wavefront distortions is rather scatter is clearly higher and more The patterns generated by the high-frequency phase variations are also quite different between the two orientations. In the axial direction the phase contrast appears as more or less random speckles, while in the radial direction the contrast reveals long vertical striations. These patterns are both consistent with the corresponding macroscopic grain structures. Figure 5.13 Phase images taken through 1"-thick samples of the mill annealed Ti-6Al4V covering a 0.4" x 0.4" area using the 0.25"-diameter transmitter at 9.7 MHz. Raw data from the axial direction (a) after removing the first 2x foldback, same for the radial direction (b). High-pass filtered data from axial (c) and radial (d) directions. Finally in addition to texture, we wanted to better understand the influence of phase distortion on attenuation. Laser-based attenuation results were collected similar to the approach used in the pulse-echo immersion experiment, where we simply measured 132 the loss in dB between the first and second echoes. backwall The only significant difference for the laser detection approach is that we needed to account for the impedance mismatch loss due to mechanical mounting of the transmitter to the sample. This loss is due to the relatively low reflection coefficient at the interface between the specimen and the coupled contact transmitter, compared to the specimen / water interface. Here, part of the energy is lost due to wave propagation back into the transmitter. This loss was measured to be approximately 6 dB. Otherwise, except for the fact that the measurement was performed in through-transmission, the data collection routine was the same as the immersion attenuation measurements. Figure 5.14 shows the outcome of these attenuation a O deg, 0.5" aperture 0 90 deg, 0.5" aperture average loss (dB) e 0 deg, laser spot o 90 deg, laser spot 9 10 11 frequency (MHz) Figure 5.14 Comparison of attenuation measurements from a 0.5"-diameter unfocused immersion transducer in pulse-echo, and from laser interferometric detection using a spot size of approximately 50 pm in throughtransmission. measurements, directly comparing to data from the immersion experiment (0.5"-aperture transmitter / receiver) with the laser experiment (approximately 50 jm- spot size). The error bars represent the scatter in the data points collected, derived from + one standard deviation from the mean. The laser measurements consistently show a reduction in attenuation across the range of frequencies examined for the axial direction, with only slight overlap in the scatter. In contrast, the laser detected attenuation results from the radial direction are virtually indistinguishable from the immersion measured attenuation results, with extensive data overlapping. Clearly, phase perturbation accounts for a significant part of the attenuation in the axial direction and has no measurable effect in the radial direction (see also Table 5.5 - Phase-Sensitive Versus Phase-Insensitive Attenuation, Appendix A). 5.5 Summary The influence of crystallographic texture and the resulting phase perturbation on attenuation was experimentally investigated for the axial and transverse directions of a mill annealed Ti-6AI-4V bar. The orientation dependence of longitudinal and shear wave velocities indicate the material is highly textured and this was also verified with analytical x-ray results. While the structural features including grains, grain boundaries, and macroscopic grain colonies all tend to be aligned in the axial direction of the bar, the principal direction of the hexagonal crystallographic lattice (i.e. the basal normal) for the grains preferentially lie in the transverse plane of the bar. The phase mapping results revealed approximately 30% higher scatter, measured over the center fringe in the axial direction, relative to the transverse direction. The higher degree of phase scattering in the axial direction of the bar is due to the presence of elongated macroscopic inhomogeneities aligned in the same direction. These extended colonies of similarly oriented grains in the axial direction are thought to essentially behave as single crystals, resulting in local disturbances in the arrival time of the propagating wavefront that extend beyond the finite aperture of the transmitter. While there also exists phase scatter in the radial direction, it is significantly smaller and more uniform than that of the axial direction. Measurements also revealed substantially lower attenuation values for the laser experiment due its relative phase-insensitivity, provided by the microscopic footprint of the focused laser detector. While laser detection measurements demonstrate immersion results, the attenuation is still clearly dominant lower attenuation than the in the axial direction, as compared with the transverse direction. These measurements demonstrate that the wider apertures generally used by conventional transducers clearly suffer phase cancellation losses, resulting in significantly higher attenuation results. The attenuation in this material is unusual considering that the backscatter is 2 to 3 times stronger in the radial direction than in the axial one; and the forward scatter is at least as strong, although less divergent, than in the axial direction. This is because elongated grains and grain colonies present larger total scattering cross-section (larger effective size) along their axis compared to normal to them, but backscattering and forward scattering coefficients are actually higher in the normal direction. Therefore, despite the stronger backscatter in the radial direction, the axial direction has higher attenuation. Finally, this method development made possible to image the microstructure by using the laser detection technology. 135 it Chapter 5, References: 1. J. C. Williams and E. A. Starke, Jr., "The Role of Thermomechanical Processing in Tailoring the Properties of Aluminum and Titanium Alloys," Deformation, Processing, and Structure, G. Krauss, ed. (ASM International, 1984) pp. 306-314. 2. A. Juva and M. Haarvisto, "On the Effects of Microstructure on the Attenuation of Ultrasonic Waves in Austenitic Stainless Steels," The British Journal of Nondestructive Testing, 19(6), pp. 293-297 (1977). 3. B. L. Baikie, A. R. Wagg, M. J. Whittle, and D. Yapp, "Ultrasonic Inspection of Austenitic Welds,” Journal of the British Nuclear Energy Society, 15(3), pp. 257-261 (1976). 4. P. D. Panetta, F. J. Margetan, I. Yalda, and R. B. Thompson, "Ultrasonic Attenuation Measurements in Jet-engine Titanium Alloys," Review of Progress in Quantitative Nondestructive Evaluation, Vol. 15B, Thompson and Chimenti, eds., (Plenum Press, NY, 1996) pp. 1525-1532. 5. P. D. Panetta, F. J. Margetan, I. Yalda, and R. B. Thompson, "Observation and Interpretation of Microstructurally Induced Fluctuations of Back-surface Signals and Ultrasonic Attenuation in Titanium Alloys," Review of Progress in Quantitative Nondestructive Evaluation, Vol. 16B, Eds., D. O. Thompson and D. E. Chimenti, (Plenum Press, NY, 1997) pp. 1547-1554. 6. T. Seldis and C. Pecorari, "Scattering-induced Attenuation of an Ultrasonic Beam in Austenitic Steel," Journal of the Acoustical Society of America (submitted for publication). 7.M. R. Hollard and J. G. Miller, "Phase-insensitive and Phase-sensitive Quantitative Imaging of Scattered Ultrasound Using a Two-dimensional Pseudo-array," Ultrasonics Symp. Proc. IEEE Cat. No. 88CH2578-3, pp. 815-819 (1988). 8. L. J. Busse and J. G. Miller, "Response Characteristics of a Finite Aperture, Phase Insensitive Ultrasonic Receiver Based Upon the Acoustoelectric Effect," Journal of the Acoustical Society of America, 70(5), pp. 1370-1376 (1981). 9. L. J. Busse and J. G. Miller, "Detection of Spatially Nonuniform Ultrasonic Radiation with Phase-sensitive (piezoelectric) and Phase-insensitive (acoustoelectric) Receivers," Journal of the Acoustical Society of America, 70(5), pp. 1377-1386 (1981). 10. D. Liu and R. C. Waag, "Estimation and Correction of Ultrasonic Wavefront Distortion Using Pulse-echo Data," Ultrasonics Symp. Proc. EEE Cat. No. 0-7803-36151/96, pp. 1391-1394 (1996). 11. J. Lwin and W. D. O'Brien, Jr., "Tissue-induced Ultrasonic Wavefront Distortion,” Ultrasonics Symp. Proc. YEEE Cat. No. 0-7803-3615-1/96, pp. 1415-1418 (1996). 12. J. P. Monchalin, "Optical Detection of Ultrasound," JEEE Transactions on Ultrasonics, Ferroelectrics, and Frequency Control, 33, pp. 485-499 (1986). 13. S. Hirsekorn and W. Amold, "High Resolution Materials Characterization by Conventional Near-field Acoustic Microscopy," Ultrasonics, 36, pp. 491-498, 1998. 14. B. T. Khuri-Yakub, S. Akamine, B. Hadimoglu, H. Yamada, and C. F. Quate, "Nearfield Acoustic Microscopy," Scanning Microscopy Instrumentation (SPIE, Bellingham, 1992) No. 1556, pp. 30-39. 15. J. P. Monchalin, R. Heon, P. Bouchard, and C. Padioleau, "Broadband Optical Detection of Ultrasound by Sideband Stripping with a Confocal Fabry-Perot," Applied Physics Letters, 55(16) pp. 1612-1614 (1989). 16. W. Hassan and P. B. Nagy, "Experimental Investigation of the Grain Noise in Interferometric Detection of Ultrasonic Waves," Journal of NDE (submitted for publication). 17. M. Hercher, "The Spherical Mirror Fabry- -Perot Interferometer," Applied Optics, 7(5), pp. 951-966 (1968). 18. ASTM Standard E 494-75, “Standard Practice for Measuring Ultrasonic Velocity in Materials,” ASTM (1985). 19. E. P. Papadakis, “ The Measurement of Ultrasonic Velocity,” Ultrasonic Measurement Methods, Physical Acoustics , Vol XIX, Eds. R. N. Thurston and A. D. Pierce, (Academic Press, New York, 1990) p. 91. 20. P. C. Waterman and L. J. Teuntonico, "Ultrasonic Double Refraction in Single Crystals," Journal of Applied Physics, 28(2), pp. 266-270 (1957). 21. R. Truell, L. J. Teuntonico, and P. W. Levy, "Detection of Directional Neutron Damage in Silicon by Means of Ultrasonic Double Refraction Measurements," Physical Review, 105(6), pp. 1723-1729 (1957). 22. K. Goebbels, "Structure Analysis by Scattered Ultrasonic Radiation," Research Techniques in Nondestructive Testing, Sharpe, ed. (Academic, New York, 1980) Vol. IV, pp. 87-157. 23. H. Willems and K. Goebbels, "Characterization of Microstructure by Backscattered Ultrasonic Waves," Metal Science, 15, pp. 549-553 (1981). 24. C. B. Guo, P. Hiller, and K. Goebbels, "Scattering of Ultrasonic Waves in Anisotropic Polycrystalline Metals," Acoustica, 59(2), pp. 112- 120 (1985). 25. K. Goebbels, "Evaluation of the Structure of Steels by Ultrasonic Scattering," Materials Testing. 77(7), p. 231-233 (1975). 26. H. W. Koglenik and T. Li, "Laser Beams and Resonators," Applied Optics, 5, pp. 1550-1555 (1966). CHAPTER VI DEVELOPMENT OF AN EDDY CURRENT MATERIALS CHARACTERIZATION METHOD FOR TITANIUM ALLOYS 6.1 Introduction This chapter discusses the role electrical anisotropy plays in the structural integrity of polycrystalline assessment titanium alloys from the standpoint of fatigue crack detection and the related issue of microstructural noise. In eddy current inspection of noncubic crystallographic classes of polycrystalline metals the electric anisotropy of individual grains produces an inherent microstructural variation or noise that is very similar to the well-known acoustic noise produced by the elastic anisotropy of both cubic and noncubic materials in ultrasonic characterization. The presented results demonstrate the electrical that although grain noise is detrimental in eddy current nondestructive testing for small flaws, it can be also exploited for characterization of the microstructure in noncubic polycrystalline materials such as titanium alloys in the same way acoustic grain noise characterization of the microstructure in different materials. 138 is used for ultrasonic Elastic anisotropy of single crystals plays an important role in ultrasonic materials characterization of polycrystalline materials. Microscopically homogeneous but randomly oriented individual medium which grains produces make up incoherent isotropic but inhomogeneous scattering commonly called "grain noise." a macroscopically wave While acoustic grain noise has an obvious adverse, often prohibitive, effect on ultrasonic flaw detection, [1, 2] it can be also exploited for ultrasonic characterization of the grain structure [3, 4, 5, 6]. Electric anisotropy exhibited by specific types of crystallographic classes can play a very similar role in electromagnetic testing of polycrystalline metals. All physical properties relating two first-order tensor quantities are characterized by second-order tensors, the directivity of which can be represented by a symmetric ellipsoid (7, 8]. Such properties include electrical and thermal conductivity, thermoelectricity, diaand paramagnetism, and dielectricity. In the most common degenerates into a sphere and these properties become cubic system, the ellipsoid fully isotropic. However, in noncubic materials the same physical properties are inherently anisotropic. In contrast, elastic material properties relate two second-order tensor quantities they are characterized by fourth-order tensors [9]. As a result, from an elastic point of view, cubic crystals are also anisotropic just like other crystallographic classes. In nondestructive materials characterization electrical conductivity is usually measured by the non-contacting eddy current method. Neighbor was the first to extend the eddy current method to electrically anisotropic materials and showed theoretically that one can obtain the full conductance tensor from such measurements. Special eddy current coil configurations that allow the simultaneous measurement of electrical conductivity in two principal directions have been developed for texture assessment in plates [10, 11]. Just like in the case of elastic anisotropy, the source of electrical anisotropy can be either (i) intrinsic crystallographic anisotropy in single crystals and textured polycrystals or (ii) structural anisotropy caused by oriented reinforcement in composite materials. The latter can be exploited for eddy current assessment of constituent volume fractions in metal matrix composites [12, 13]. Grain boundary contributions to the electrical resistivity [14] can cause additional electrical anisotropy materials with elongated in polycrystalline grains aligned in preferred orientation due to thermal or mechanical treatment. It should be emphasized that, in contrast with elastic properties, the electric conductivity is completely isotropic in cubic crystals which constitute the overwhelming majority of polycrystalline therefore, metals; the role of intrinsic crystallographic anisotropy in eddy current testing has not been investigated in detail. However, less common materials of hexagonal symmetry can exhibit strong electrical anisotropy with significant difference in conductivity between the basal plane and normal to it. Titanium is one of the few structural metals applications, which preferentially of practical crystallizes importance, in hexagonal especially in aerospace symmetry and therefore exhibits strong electrical anisotropy. There are two areas where this electrical anisotropy becomes very relevant from the point of view of mechanical fatigue in titanium alloys. First, eddy current inspection is probably the most commonly used nondestructive testing technique for fatigue crack detection in airframe structures and engine components and electrical grain noise presents the same problem in eddy current crack detection as acoustic grain noise does in ultrasonic flaw detection. Second, proper microstructure is 140 absolutely essential for assuring good grain electrical noise exploited be can fatigue tolerance in the material [15] and the of the characterization for nondestructive microstructure by eddy current inspection in the same way as acoustic grain noise is used in ultrasonic characterization of the microstructure. Another goal of this chapter is to investigate the feasibility of exploiting the eddy unique current grain noise the for alloys in titanium observed purposes of nondestructive materials characterization. The main achievements of this effort are (i) the strong electrical grain demonstrated, (ii) the physical very noise in titanium alloys for responsible mechanism been has this experimentally contrast been has theoretically explained, (iii) analytical, finite element, and experimental methods were used to investigate lateral that assure characterization were the best developed eddy of resolution procedures optimization materials the imaging and current microscopy, resolution verified. It is and (iv) for microstructural shown that electric anisotropy exhibited by noncubic crystallographic classes of materials can play a very similar role in electromagnetic materials characterization of polycrystalline metals to that of elastic anisotropy in ultrasonic materials characterization. Titanium is one of the few ~ structural metals of practical importance, especially in aerospace applications, which preferentially crystallizes in hexagonal symmetry and therefore exhibits strong electrical anisotropy. At the same time, the titanium alloy microstructures of interest tend to form a rather coarse, locally textured microstructure featuring large colonies of hexagonal alpha grains of similar orientation. The fracture and fatigue resistance of this material is strongly affected by the type of microstructure, hence there is a continued need for new 141 nondestructive evaluation techniques that are capable of both imaging and quantitatively characterizing microstructure. It was found that the lateral resolution of eddy current imaging is ultimately limited by the probe-coil geometry and dimensions, but both the inspection frequency and the phase angle can be used to optimize the resolution, to some degree, at the expense of sensitivity. Although eddy current imaging is still in its infancy, a direct comparison of 5-MHz eddy current and 40-MHz acoustic microscopic images of the same coarse-grained Ti-6Al-4V indicated that the same sample features can be observed by both methods at approximately the same resolution level. This work also shows experimentally that eddy microscopy current can be enhanced via a high- resolution, small diameter probe-coil which delivers a unique materials characterization tool well suited for the evaluation of Ti alloys. Eddy current imaging has been used for years in flaw detection, corrosion mapping, and other nondestructive testing and materials characterization applications to increase the amount of information obtained by conventional point-by-point inspection and enhance its clarity [16, 17, 18, 19, 20, 21, 22, 23]. Another emerging technology capable of mapping electrical conductivity distributions at considerably better resolution though with inherently less sensitivity is near-field microwave imaging. Ash and Nichols were probably the first to use near-field scanning for electromagnetic imaging in 1972 [24]. Using 3-cm microwave radiation, they achieved a resolution of direction and 2/20 for two-dimensional 4/60 in one objects. Since the early use of microwave radiation, several techniques have been developed for near-field imaging. Among the various versions developed during the last few years, the transmission-line resonator technique appears to be the most promising for high-resolution near-field microwave inspection of surface conductivity [25, 26, 27]. Eddy current inspection can be readily adapted for imaging via automated scanning since it is noncontacting in nature and less sensitive to surface topography than near-field microwave imaging. One of the most important advantages of eddy current imaging over ultrasonic imaging is that there is no grain noise in cubic materials which constitute the overwhelming majority of structural However, metals. electrical that noncubic it was recently found that presents a serious anisotropy limitation materials for flaw exhibit substantial detection [28]. In particular, in many titanium alloys microstructure scatter generates noise that can be very severe since often large colonies of similarly oriented alpha grains form a coarse, textured structure that effectively hides small defects. As an example, Figure 6.1 shows the eddy current images of small fatigue cracks of approximately 0.025 inch in length in 2024 aluminum and Ti-6Al-4V specimens at 2 MHz. A 0.060"-diameter coil was used to scan a 0.5" x 0.5" area over the surface of the specimens. The crack is somewhat smaller than the inside diameter of the coil, therefore it produces a characteristic double-image. Two dark spots appear at the two opposite ends of the crack when it intersects the path of the eddy currents. The diameter of the dark spots is approximately 0.040" and is determined by the size of the probe as discussed later. It is quite obvious from these results that even this less than fully resolved flaw is readily detected in aluminum since the background is quite uniform, i.e., there is no microstructural noise. In comparison, a similar crack is barely detectable in titanium because of the presence of strong grain noise caused by the unique electrical anisotropy of Ti-6Al-4V, 0.026-in.-crack Al 2024, 0.025-in. crack Eddy current images of small fatigue cracks in 2024 aluminum and Ti6AI-4V specimens (0.5" x 0.5", 2 MHz, 0.060"-diameter coil). Figure 6.1 hexagonal crystallites. The physical origin of this unique crystallographic noise and its possible role in quantitative analysis of the grain structure in titanium alloys was recently shown using single crystals of cubic Al and Cu, and hexagonal Cd [28]. This letter experimentally demonstrates the dependence of electrical conductivity on the relationship between the surface normal direction and the principle direction in noncubic crystallographic classes of materials. Based on our current understanding, we know (i) that in coarse-grained Ti-6Al4V, the detectability of small fatigue cracks by eddy current inspection is ultimately limited by grain noise and (ii) that this grain noise can be exploited for the purpose of characterizing the microstructure by high-resolution eddy current imaging. Because of the small penetration depth of eddy current inspection, only the grain structure at the surface can be inspected. The electrical anisotropy of noncubic materials like titanium and its 144 alloys can be also exploited to assess texture at the surface and, by direct contact resistivity measurements, even in the interior of the material. The grain noise observed in eddy current images is due to the variation of the average conductivity between different crystallographic planes. In spite of the difference in the origin of contrast, namely electrical versus mechanical properties of the material, the eddy current contrast mechanism is quite similar to the contrast produced in spherical acoustic microscopy and therefore [29]. Both techniques are sensitive to crystallographic orientation produce images on which large colonies orientation show up as essentially homogeneous of smaller grains of similar domains. This occurs in spite of the partitioned appearance of individual grains of a macroscopic colony as featured in an optical micrograph of metallurgically prepared samples which, due to chemical etching (as in Fig. 6.2.a), produces a contrast regardless whether there is a difference in orientation between neighboring features or not. Figure 6.2 illustrates that essentially the same macroscopic inhomogeneity of the microstructure can be observed in coarse-grained polycrystalline Ti-6Al-4V via eddy current imaging and acoustic microscopy. These 1"x 1" images were scanned from a specially heat treated sample to bring about a high degree of grain consolidation to the structure. The large lamellar colonies are composed of alternating plates of the alpha and beta phases. The principal direction in each of these large colonies is essentially uniform, thereby forcing the colony to behave as if it were a single crystal. The main reason for generating a sample with such exaggerated structure is to determine if like features could be observed with both eddy current and acoustic scanning techniques. Clearly, these like features do simultaneously appear in the images 145 | @ f% Figure oie fs c) 40 MH zZ acoustic m. icrog raph Comparison of (a) optical, (b) eddy current, and (c) acoustic microscopic images of a coarse-grained Ti-6Al-4V sample (1" x 1") from nearly the same area on the sample. Figure 6.2 of b) 5 MHz eddy current ) / a) optical image ese due 6.2, to the relative crystallographic orientation the of various microstructural features encountered during scanning. 6.2 Eddy Current Experiments Single crystal materials generally behave anisotropically in response to a given stimulus such as heat, electricity, or force due the symmetry conditions of the atomic lattice structure. In contrast, polycrystalline materials tend to behave essentially effect associated with the presence of a random crystallographic orientation distribution of the individually homogeneous grains which isotropically due to the averaging constitute the solid. This paper demonstrates that eddy current evaluation is capable of resolving the crystallographically related directional dependence of electrical conductivity in noncubic polycrystalline metals. In coarse-grained structural alloys of polycrystalline titanium the spatial variation in electrical conductivity significantly reduces flaw detectability. On the other hand, the physical mechanism responsible for reduced flaw 146 delectability provides a source of information about the microstructural make-up of the material, and therefore can be used for materials characterization. Eddy current testing is the most common electromagnetic nondestructive evaluation method and is widely used in the aerospace industry. Small diameter coils combined with a computer controlled scanning mechanism can be readily used for eddy current imaging. The coil impedance is determined by the resistivity of the specimen as measured by the eddy current, which runs parallel to the surface in a concentric circle with the coil. In this way, an eddy current probe measures the average resistivity in a given plane rather than in a given direction. As the probe is moved along the surface, it measures the local average resistivity along the path of the eddy current in the plane of the surface. The resistivity is integrated over the entire probe circumference in the eddy current path, resulting in grain contrast that is proportional to the average resistivity between the different crystallographic planes. This contrast is similar to the mechanical contrast produced by spherical acoustic microscopy, which is determined by the variation of the average surface wave velocity between different crystallographic planes [29]. For a hexagonal crystal like pure titanium and its most common alloys the axial symmetry around the principal direction (the hexagonal axis) allows the directional dependence of the electrical resistivity to be described over the entire space by two orthogonal axes and the directivity can be represented as an ellipsoid: (6.1) pg) = p,cos’¢ +p sin’¢, 147 in cubic materials the electrical resistivity is fully isotropic due to the balanced symmetry of the lattice structure, i.e., the resistivity becomes a single scalar value and the ellipsoid describing measurements eddy current to a sphere. For of electrical resistivity in a hexagonally symmetric single crystal, the its directional dependence degenerates average surface resistivity can be expressed from Eq. (6.1) as: pO) = 1/2[p, sin’ 0+ p (it cos’), where 0 denotes the inclination angle between the basal plane and the surface of the specimen. For example, in pure titanium the resistivity is approximately 6% P 1 = 48 pQcem and p= 45.35 wOcm, lower in the basal plane than normal 1.., to it (30). Because of the above described averaging effect of eddy current inspection, the actual grain contrast is expected to be 50% lower in eddy current inspection. In titanium, the average resistivity is approximately 3% lower when the basal plane is parallel to the surface than when it is normal to it. In order to assess the feasibility of eddy current materials characterization and flaw detection in structural alloys of noncubic symmetry, we carried out two sets of experiments. First, we used an eddy current probe to measure the directional variation of the electrical conductivity in pure single crystals of aluminum, copper, and cadmium; the former two materials consist of a cubically symmetric crystallographic lattice, the latter one consists of a hexagonally symmetric lattice (unfortunately, titanium single crystals be cannot grown to sizes enough large eddy accurate for current conductivity measurements). Second, we used an eddy current scanner to map the electrical grain noise in Ti-6Al-4V titanium alloy specimens of different microstructures. The Al, Cu, and Cd single crystals used in this study were of random orientation. Each specimen was a solid cylinder of approximately 2" length and 0.5" diameter, large enough to section into multiple test samples of varying surface orientation. Eddy current resistivity measurements were taken on the various single crystal sample sets using a Nortec 19e eddy current instrument and a 0.060"-diameter probe at 2 MHz. If the eddy current probe is energized in air, away from the conducting sample, the instrument will register the baseline impedance of the coil. As the probe is moved closer to the conductive sample, the impedance indication will begin to change and becomes strongest when the probe is directly on the sample. The variation in the eddy current indication related to the spacing between the coil and the sample is known as "lift-off." When the probe is in contact with the sample, the coil impedance indicated on the instrument represents the electrical resistivity. Moreover, the fact that the lift-off curve approaches the resistivity curve at an angle allows the separation of lift-off from resistivity by proper adjustment of the phase angle on the instrument [31]. For each set of samples, the phase angle was set to isolate lift-off to the horizontal direction, the sensitivity was adjusted, and the instrument was nulled. The vertical output from the eddy current instrument, corresponding to the average electrical resistivity, was captured on a digital oscilloscope. With this automated approach, it was possible to statistically analyze the population of average surface resistivity values corresponding 149 to approximately 500 individual measurements (see sample for each Electrical - Normalized 6.1 Table Resistivity, Appendix A). The measured data are shown in Figures 6.3.a through ¢ as histograms of the probability distributions of the surface resistivity for various surface crystallographic orientations in the three single crystals. For each set, only three surfaces showing the most extreme differences in average resistivity are displayed. It should be noted that these values in average electrical resistivity are subject to a variety of small experimental an errors, including thermal drift from the instrument or sample, probe alignment and associated probe rocking effect, inevitable thickness and edge effects, etc., hence the variability in the data. These factors were considered during the data collection and efforts were taken to minimize their The affects. data from 6.3.c Figure clearly of the electrical resistivity in cadmium demonstrates the crystallographic dependence representing noncubic materials, as opposed to the lack of separation demonstrated by cubic copper and aluminum in Figures 6.3.a and 6.3.b. In the cadmium crystal the values of electrical resistivity are py = 83 pQcm and p= 6.8 wQcm, a relatively large difference of approximately 22% between the basal plane and the normal to it. Due to the averaging effect, the most extreme resistivity separation which could be expected in Cd by eddy current measurement present in the randomly is approximately cut Cd samples was 11%. The average resistivity variation clearly measurable with a maximum variation of approximately 3% in resistivity. Considering that we did not necessarily find the principal planes of maximum separation, the measured variation is reasonable. Probability Density a) Aluminum Probability Density b) Copper 2 n Swo Qa 2 6 Fe 2 ° a Probability Densi = Normalized Surface Resistivity [%] d) Figure 6.3 Electrical resistivity probability distributions for three single crystal surface orientations in a) aluminum, b) copper, and c) cadmium and d) on the surface of polycrystalline Ti-6V-4V (solid lines are best fitting Gaussian distributions). Because of our particular interest in nondestructive testing of high-strength titanium alloys by eddy current methods, a special attempt was made to obtain the same type of data from a pure alpha (hexagonal) phase Ti single crystal. However, due to the inherently small size of the available Ti single crystals, it was not possible to collect data actually representative of the material's electrical resistivity due to edge affects, which tend to diminish the accuracy of the measurements. Moreover, in titanium, the maximum difference in average resistivity is expected to be only about 3%, i.e., only one fourth of the corresponding variation in Cd. Nevertheless, based on the results from the Cd crystal sample, the evidence of electrical anisotropy in noncubic crystalline materials is clearly further supported. To distribution of the demonstrate surface this point, Figure resistivity for a Ti-6Al-4V 6.3.d shows polycrystalline the probability specimen. As expected, there is a significantly wider variation in the resistivity from point to point than on single crystals, which will be shown later to be caused by the relatively coarse grain structure. Structural alloys of titanium are comprised of microscopically anisotropic grains of random order, which macroscopically behave isotropically. However, often the materials fabrication process results in both small- and large-scale structures which lack the degree of randomness in the crystallites’ orientation required to allow the behavior to be fully isotropic. These materials are said to contain texture, which is generally imparted to the material via plastic deformation, like forging. Texture results, for example, in the alignment of like crystallographic slip planes parallel to the rolling plane, while certain slip directions tend to align in the direction of rolling or wire drawing. The development of this preferred orientation also tends to align microstructural features like inclusions, second phase particles, or grain boundaries and the texture affects are often observed on a large scale relative to the individually homogeneous grains, often spanning several inches or more. In some polycrystalline titanium alloys, certain microstructural conditions give rise to a highly localized form of crystallographic microtexture causing fractures to preferentially occur along certain weak erystallographic directions [15]. The macroscopic inhomogeneity of the microstructure in polycrystalline Ti-6AlIAV can be observed via eddy current imaging as shown in Figure 6.4, which correspond 1" area on the samples. The specimen shown in Figure 6.4.a contains a gross to 1" x microstructural anomaly clearly visible on the right side of the eddy current image. Figure 6.4.b shows a typical billet microstructure with texture related features in the horizontal direction while Figure 6.4.c shows a large grained sample and Figure 6.4.d shows an equiaxed annealed beta microstructure. In short, some interesting parallels can be observed between the reported electromagnetic approach and conventional ultrasonic evaluation methods. can be used to exploit the fact that in Ultrasonic techniques polycrystalline materials, grain to grain differences in crystallographic orientation and the presence of grain boundaries provide a source for scattering of ultrasonic energy. The presence of texture and additional phases of material also play an important role in the ultrasonic response of the material and the scatter provides a source of data which can be used to characterize the microstructural features. In ultrasonic flaw detection, the acoustic grain noise is clearly detrimental due to reduced detection threshold. Likewise, electromagnetic inspection techniques benefit from the fact that noncubic systems exhibit d) Scanned eddy current images of different Ti-6Al-4V microstructures; a) sample containing a severe microstructure anomaly (right side, middle); b) the billet microstructure showing texture related features in the horizontal direction; c) a large grained sample; and d) equiaxed beta annealed microstructure (dimension 1" x 1"). electrically anisotropic properties, characterization, for microstructural allowing and suffer from the fact that the electrical scatter originating from varying local resistivity raises the noise floor, thereby reducing The detectability. flaw electrical anisotropy observed with eddy currents in noncubic metals is therefore analogous to the elastic anisotropy observed nondestructive with ultrasonic of polycrystalline evaluation anisotropy of noncubic knowledge, the techniques crystals significant role titanium known is a well played and has by the strong alloys. physical microscopic implications for the the electrical Although fact, to the best of our electrical anisotropy of individual grains in the macroscopic eddy current response of the polycrystalline material has never been pointed out or investigated in any depth. 6.3 Resolution of Eddy Current Imaging In contrast to other more conventional applications of eddy current imaging, microstructure characterization crucially depends on achieving truly microscopic imaging resolutions. The lack of published theoretical and experimental investigations on the factors affecting the lateral resolution of eddy current inspection prompted us to initiate this research effort aimed at laying down the groundwork for developing a high- resolution eddy current microscope capable of resolving the fine details of the textured microstructure in titanium alloys. Numerous efforts have been made in the past to study the eddy current distribution of probe-coils by analytical, numerical, and experimental means. In spite of the general nature of these methods, most of the published results were focused mainly on the axial penetration depth of the probe, which is undoubtedly the primary consideration in the overwhelming majority of NDT 155 applications. Much less is known about the radial penetration depth or lateral resolution of eddy current inspection except that it is essentially governed by the geometry and dimensions of the probe-coil. It is usually assumed that the eddy currents induced in the material are essentially mirroring the excitation in the probe-coil, current running therefore the lateral extent of the inspected region under the coil is more or less independent of frequency. However, there is clearly a radial spreading of the eddy current distribution as the frequency is lowered, which will adversely affect the lateral resolution of the eddy current inspection by small coils used in eddy current microscopy. This is because in eddy current microscopy, such as microstructural characterization of polycrystalline titanium alloys, we are trying to optimize the inspection parameters for maximum lateral resolution in the near-surface region of the specimen, therefore very small probes are used at very high frequencies. By comparison, in conventional eddy current imaging, such as flaw detection or corrosion assessment, a relatively large penetration depth is desirable which inevitably requires a large probe diameter [32, 33]. All the necessary analytical [34, 35] and numerical [36, 37] tools for investigating the radial penetration depth or lateral resolution of eddy current microscopy are readily available in the literature. It has been previously demonstrated that on simple axisymmetric configurations these techniques provide essentially identical solutions for the axial penetration depth of the eddy current [38], as such they are also expected to work equally well for radial penetration depth estimations. In our finite element (FE) calculations we used a commercially available software capable of simulating the magnetic field and eddy current density distributions for axisymmetric configurations 156 Figures 6.5 and 6.6 show the magnetic [39]. As an example, field and eddy current distributions, respectively, produced by a small pancake coil in titanium at four different frequencies (air-core, inside diameter 1 mm, outside diameter 1.5 mm, height 2 mm). most The convenient, albeit somewhat over-simplified measure of frequency relative to the conductivity of the material and the size of the probe is the so-called standard penetration depth [40] OHO @ where (6.3) : 5 = | 2 is the angular frequency, and specimen, of the w= 4n10-7Vs/Am conducting half-space respectively and 6 (in and our o finite are the permeability and conductivity element simulations we took A/ Vm). For a plane wave incident on a o = 214x 10° actually gives the 1/e skin depth of the exponentially decaying eddy current density, but for any finite-sized probe-coil it is only a useful parameter which happens to be the upper limit for the axial penetration depth. Whenever the standard penetration depth is very large with respect to the dimensions of the coil the magnetic field is essentially unaffected by the flow of eddy currents in the specimen and can be approximated by the magnetic field produced by the coil far away from the conducting half-space. In this frequency range (up to about 10 kHz or 6 ~ 3.4mm in our case) the eddy current distribution is also independent of frequency while its absolute density proportionally increases with frequency. Whenever the standard penetration depth 157 1mm <——_>| Figure 6.5 Magnetic field distribution produced by a small pancake coil in titanium at four different frequencies. 1 mm <—_>| Figure 6.6 ~° Eddy current distribution produced by a small pancake coil in titanium at four different frequencies. is very small with respect to the dimensions of the coil the magnetic field is essentially eliminated by the flow of eddy currents in the specimen below a certain skin depth which approaches the standard penetration depth. In this frequency range (above approximately 1 MHz or & ~ 034mm _ in our case) the eddy current distribution is also limited to this shallow layer determined by the standard penetration depth. Figure 6.7 shows the axial penetration depth versus frequency curve for a 1-mm-inside diameter pancake coil in titanium. The symbols represent the numerical results calculated by finite element (FE) simulation, the solid line represents the general trend of the FE data, and the dashed line is the plane wave asymptote calculated from the standard penetration depth according to Eq. (6.3). The axial penetration depth was calculated from the eddy current intensity directly under the coil at its middle dpiddle = (douter + Ginner)/2. This figure well demonstrates the substantial difference between the true axial penetration depth of a finite-diameter coil and the standard penetration depth at low frequencies, which was first pointed out by Mott [32] and later further investigated by Stucky and Lord [38]. Our main interest in this study is a similar investigation of the radial penetration depth of eddy currents generated by a finite-diameter probe-coil, which can be readily done by analyzing the same set of FE data*. Figure 6.8 shows the radial penetration versus frequency for a 1-mm-diameter pancake coil in titanium. The solid circles represent the numerical results calculated by finite element (FE) simulation and the solid line illustrates the general trend of the FE data. In addition, the empty circles represent‘the analytical results calculated by Dodd and Deed's method [34] and the dashed line illustrates the general trend of the analytical data. The radial penetration depth was 160 measured from the axis of the coil to the point on the surface of the conducting half-space where the eddy current density dropped to 1/e relative to the maximum, which is directly under the middle of the coil. In spite of a minor numerical discrepancy between the FE simulation and the analytical results the agreement is quite acceptable. Both methods indicate that the penetration radial approaches low-frequency a asymptote of approximately 1.8 mm. This value is clearly sensitive to the shape of the coil (both insideto-outside diameter and diameter-to-height ratios) but it is roughly equal to the outer coil 10000 standard penetration depth —®— = finite element =. [a G~~ QO. oO A c io) eer iss} 5 oO q oO [a is 10 0.00001 v Lf LU q 0.0001. 0.001 0.01 0.1 Frequency [MHz] Figure 6.7 Axial penetration depth versus frequency for a 1-mm-diameter pancake coil in titanium. The symbols represent the numerical results calculated by finite element (FE) simulation, the solid line represents the general trend of the FE data, and the dashed line is the plane wave asymptote calculated from the standard penetration depth according to Eq. (6.3). 161 e veeee $ wen eeeee finite element ---Q-- analytical ip Radial Penetration [mm] 18 ——@®—_ 1.2 4 0.8 0.00001 t T LJ T J v 0.0001 0.001 0.01 0.1 1 10 100 Frequency [MHz] Figure 6.8 Radial penetration p versus frequency for a 1-mm-diameter pancake coil in titanium. The solid circles represent the numerical results calculated by finite element (FE) simulation, the solid line illustrates the general trend of the FE data, the empty circles represent the analytical results calculated by Dodd and Deed's method, and the dashed line illustrates the general trend of the analytical data. diameter. At high frequencies, the radial effective outer diameter of the coil is slightly larger than the outer diameter of the coil, ie., dog > doyter. In order to better illustrate this relationship, Figure 6.9 shows the corrected radial penetration Peg, = P — deg 12 versus frequency curve for the same 1-mm-diameter pancake coil in titanium compared to the standard penetration depth, which is of course independent of the size and shape of the coil. The effective diameter was found simply by best fitting the high-frequency 162 asymptotic behavior as predicted by the FE results to the standard penetration depth. In case considered here, turned diameter effective the out to be the particular dey ~ 1.79mm, i.e., approximately 19 % higher than the outer diameter of the coil. This value is clearly sensitive to the shape of the coil but we can still conclude that over a wide frequency range the radial penetration can be more or less accurately approximated = 10000 Stes ‘ ) . 3 q = 1000 > 100 - —@®— standard penetration depth finite element e _@ s © 5 Ay “s 3fa a 2 3) oO q fo) O 10 0.00001 T + T J LU v 0.0001 0.001 0.01 0.1 ] 10 100 Frequency [MHz] Figure 6.9 Corrected radial penetration Pcorr = P — Feg /2 (dog = 119 douter) versus frequency for a 1-mm-diameter pancake coil in titanium. The solid circles represent the numerical results calculated by FE simulation, the solid line illustrates the general trend of the FE data, and the dashed line is the standard penetration depth calculated from Eq. (6.3). 163 as the sum of the standard penetration depth, which depends on the material properties and frequency, and the effective coil diameter, which depends on the size and, to some degree, on the shape of the coil. 6.4 Results and Discussion The now until results indicate in eddy that current the microscopy lateral resolution is essentially determined by the probe diameter although it slightly improves with frequency. increasing commercially These predictions Nortec 3-mm-diameter available were experimentally pencil-probe of tested by using 500 kHz-1 a MHz frequency range. The nominal diameter actually indicates the case only and the inner and 2.5 outer diameters of the small coil were measured to be approximately 1.25 mm and the mm, respectively. The height of the coil could not be established without damaging probe. It was also verified that the coil had a ferrite core, but according to our finite element simulations such a core has only a relatively weak effect on the lateral distribution of the eddy current in the specimen. In order to measure the lateral resolution of this probe we first used a polycrystalline Ti-6Al-4V specimen of very coarse grain structure, zoomed in on the interface between the two largest neighboring colonies, and the determined the 10% to 90% width of the transition range by analyzing the contrast of eddy current micrograph. However, it was found that since the grain boundary is not necessarily normal to the surface of the specimen the measured lateral resolution is also affected by the axial penetration depth. In order to eliminate this artifact, we prepared a special specimen made of two flat and well polished Ti-6A1-4V halves of very fine grain structure, which were tightly compressed and held together by a bolt. The 164 reduced electrical conductivity of the resulting imperfect interface between the halves produced a dark or bright stripe on the eddy current micrograph depending on the rotation angle used. The width of this stripe was used as a measure of the lateral resolution of the eddy current probe at the given inspection frequency. Since the contrast profile was found to be approximately Gaussian in most cases, the accuracy and repeatability of the measurement was further increased by first approximating the measured profile with a normal distribution and then defining the lateral resolution as the standard deviation (1/e halfwidth) of the best fitting Gaussian curve. The measurement was carried out over a very wide frequency range from 15 kHz to 10 MHz far in excess of the 500 kHz-1 MHz nominal frequency range of the probe. In order to facilitate reliable measurements at both very low and very high frequencies where the sensitivity of the probe was marginal at did not use the Nortec best we investigation. 19e eddy current scope Instead, we used a low-impedance as we did in our previous (5 (2) bridge and an SR 530 low- frequency analog lock-in amplifier up to 100 kHz and a high-impedance (50 Q) bridge and an SR 844 high-frequency digital lock-in amplifier above 100 kHz. As we will demonstrate later, the lateral resolution is also affected by the phase angle of the detection (rotation angle of the impedance plane) therefore we had to assure that this magnitude. angle remains consistent as the frequency changes over three orders of One possibility is to use a changing reference angle, for example, always adjust first the lift-off to be horizontal and then measure the vertical component, as we did during imaging experiments. Figure 6.10 shows the experimental impedance diagram at 2 MHz in titanium. The rotation angle (-156° reference phase angle on the lock-in amplifier) was chosen so that the lift-off curve is horizontal and the instrument was nulled when the probe was above the titanium specimen but far away from the interface. As the probe approached the interface (without changing the lift-off distance from the surface of the specimen) the impedance point moved downward and to the right, which corresponds to an apparent decrease of conductivity and increase of lift-off distance. The of measuring only the vertical change advantage at horizontal lift-off is that surface alignment, curvature, and topography will not affect the contrast. However, this mode of operation would make the change of frequency rather cumbersome therefore we used a similar, but much simpler approach which is not available on commercial eddy current scopes but can be done very simply when a lock-in amplifier is used. We applied phaseinsensitive absolute value measurement to quantify the imbalance of the bridge. Since the differential output signal of the bridge is mainly due to the presence of eddy currents in the conductive half-space (lift-off curve) and is only weakly perturbed by the imperfect interface (interface curve), measuring the magnitude of the differential output signal of the bridge is essentially the same as measuring the component of the interface curve that is parallel to the lift-off curve without repeated re-adjustment of the reference phase. inspection experimentally 6.11 shows the frequency curve for the previously Figure determined described lateral commercial resolution versus pencil-probe in titanium (see Table 6.2 — Frequency Dependent Lateral Resolution, Appendix A). The solid line represents only the trend of the theoretical predictions based on our previously shown finite element simulations. Direct comparison of these predictions and the experimental data is not feasible partly because of the lack of accuracy in the geometrical 166 50 n fan) Vertical Amplitude [mV] Oo j * lift-off interface “100 4eeeee-eeee feceeeeeeeeee leveeeeeeeeees a ee -150 100 50 0 -50 150 Horizontal Amplitude [mV] Figure 6.10 | Experimental impedance diagram in titanium at 2 MHz. The rotation angle was chosen so that the lift-off curve is horizontal. dimensions and material properties (core) of the probe-coil used in our calculations but mainly because of the conceptual differences in how the radial penetration was calculated and the lateral resolution was measured. The calculations require an axisymmetric geometry therefore the flat boundary or interface cannot be incorporated into the model therefore the calculated radial penetration is not expected to be quantitatively identical to the experimentally measured lateral resolution. However, both parameters are related to the lateral spread of the eddy current distribution in the conducting specimen therefore 167 1800 - 1600 _ = 1400 —— adjusted FE prediction @ low-frequency lock-in O high-frequency lock-in Ei 1200 4 i= -= = 9 % 1000 ‘S5 600 4 2% - 800 4 3 400 - 200 - — 0 10 1 0.1 0.01 Frequency [MHz] Figure 6.11 | Experimentally determined lateral resolution versus inspection frequency by a commercial pencil-probe in titanium. The adjusted FE prediction for the radial penetration is plotted to indicate the trend of the data. they are expected to exhibit very similar frequency dependence. In order to bring out this similarity, the calculated radial penetration was shifted upward by an arbitrary amount of 180 pm. With this adjustment, the qualitative agreement between the theoretical predictions and the experimental data is quite good except maybe at very high frequencies where the experimentally determined lateral resolution does not follow the still dropping radial penetration. Over all, the effect of inspection frequency on the lateral resolution is rather modest; over the studied three orders of magnitudes the resolution changes barely by a factor of two, which confirms our previous analytical conclusion that the principal 168 parameter is the diameter of the probe-coil with only minor improvement produced by increasing the frequency. The effect of frequency on lateral resolution is demonstrated in Figure 6.12. In a this case we used a commercially available Uniwest 1180-b eddy current probe with coil diameter of approximately 0.020" and an extended, sharpened ferrite core. This to probe, while having a slightly reduced frequency range and sensitivity in comparison the 0.060" Nortec probe, has the obvious advantage of a smaller coil diameter, thereby providing better lateral scanning resolution. The rather small improvement produced by increasing the inspection frequency prompted us to look at the other parameter that can be most easily adjusted to optimize the lateral resolution of eddy current microscopy, namely the rotation angle. Assuming a) 300 kHz Figure 6.12 b) 1 MHz c) 3 MHz Eddy current images (0.5" x 0.5") taken at three different frequencies to demonstrate the effect on lateral resolution. These images were scanned from an extremely large grained polycrystalline titanium alloy. 169 that the specimen is flat therefore lift-off effects can be eliminated by careful parallel alignment of the specimen with respect to the scanning plane, the rotation angle can be freely chosen to achieve the best combination of resolution and sensitivity. Figure 6.13 shows the impedance diagrams and resolution profiles at three different rotational angles relative to the horizontal lift-off angle at 2 MHz in titanium. Solid lines show the measured data and dotted lines are best fitting Gaussian distributions used to approximate the measured contrast profiles. The thereby obtained standard deviation represents the experimentally determined lateral resolution. The rotation angle is measured from the reference phase where the lift-off direction is horizontal. In this case (Figure 6.13.a), which was used in all of our earlier shown eddy current images because of its complete cancellation of lift-off effects and was also shown in Figure 6.13, the lateral resolution is about 790 pm. The apparent curvature of the interface line on the impedance diagram allows us to further increase the resolution by simply increasing the rotation angle. For example at 120° (Figure 6.13.b), the rotated impedance locus first moves upwards as the probe approaches the interface then turns downwards as the probe gets closer to it. As a result, the lateral resolution as calculated from the Gaussian fit reduces to 510 pm at the expense of slightly reduced and somewhat distorted contrast. The contrast distortion produces weak bright bands on both sides of the otherwise dark interface region. We will show later that even better lateral resolution can be achieved by further rotating the impedance diagram up to about 140°, but at that point the reduction and distortion in contrast both become unacceptable for imaging purposes. At even higher rotational angles the contrast becomes positive as the two bright bands gradually overtake the dark stripe at the center. In this range the contrast is both low and badly distorted. At 160° (Figure a) 0° from lift-off, 790 pm standard deviation Ss ; ob : § eso :: 3,o of iB ; , oO 3 NN > > s : 3 2 ' 2 } Oo . oO > wl 7 = ' = 4 6 > Distance [1 mm/div] Horizontal Voltage [a. u.] Vertical Voltage [a. u.] Vertical Voltage [a. u.] b) 120° from lift-off, 510 pm standard deviation Distance [1 mm/div] Horizontal Voltage [a. u.] c) 160° from lift-off, 1,020 um standard deviation 3 2, 8 i) > : 3 . rossrsrs sss : Sp 8 i) > — : an oO 3 3 2 = 2 = > > o o Distance [1 mm/div] Horizontal Voltage [a. u.] Figure 6.13 1| Impedance diagrams and resolution profiles at three different rotational angles relative to the horizontal lift-off angle in titanium at 2 MHz. 171 6.13.c), the dark center line completely disappears and the contrast profile becomes Gaussian again, however the lateral resolution is very poor at about 1,020 pm. Similar measurements were carried out at different frequencies to establish the relationship between the phase dependence of the lateral resolution and the previously considered frequency dependence. Figure 6.14 shows the experimentally measured lateral resolution versus rotation angle curves for three different frequencies (see Table 6.3 — Phase Angle Dependent Lateral Resolution, Appendix A). At frequency, each the rotational angle was measured from the reference angle necessary to assure that the liftoff curve is horizontal. Clearly, all three curves reveal essentially the same dependence on the rotational increasing angle frequency. as well as a slight improvement It should be remembered in the lateral resolution that the previously presented with lateral resolution versus frequency curve (see Figure 6.11) was measured in a phase-insensitive way and represents a medium value corresponding to a rotational angle around, but not exactly, 90°. As mentioned before, measuring the absolute value of the relatively large output signal of the bridge yields only that component of the small interface signal which is parallel to the main signal. Assuming that the only imbalance in the bridge is due to the eddy current load of the measuring coil with respect to the unloaded reference coil a positioned in the shielded case, the absolute value measurement would correspond to rotation angle of 90°. However, the inevitable lack of symmetry between the measuring and reference coils and possible differences on the order of a few tenth of a percent between the driver resistances cause the bridge to be slightly out of balance even without magnetic loads on the measuring coil, therefore the lift-off signal is not exactly in phase 172 with the output signal of the bridge. In addition, the lift-off curve is slightly curved and the rotation angle is measured not from its average direction but from the slope at its end as shown in the blown-up picture of Figure 6.10. It should be emphasized again that although Figure 6.14 seems to indicate that the lateral resolution can be as low as 300-350 jum, but the accompanying contrast distortion renders this optimum impractical for imaging purposes. However, considering the rather oo = &. c Sgvar) a ° N oO fa Sspo 2 is] — 60 Rotation Angle [deg] Figure 6.14 | Experimentally measured lateral resolution versus rotation angle curves for three different frequencies in titanium. of modest improvements produced by increasing the inspection frequency, the benefits the this much simpler approach should not be underestimated. Figure 6.14 indicates that the lateral resolution can be improved by approximately a factor of two via optimizing rotational angle with respect to the routinely used horizontal lift-off. As an example Figure shows 6.15 the eddy current c-scan images of a coarse-grained Ti-6A1-4V evident specimen at 5 MHz and two different rotational angles (0.5" x 0.5"). It is quite eddy that the higher rotation angle resulted in a significant resolution improvement. The Figure current and acoustic micrographs of the same specimen were previously shown on 6.15 6.2 at a two-times lower magnification. The large bright grain at the center of Figure can be clearly recognized inverted). The optimized by its characteristic 5-MHz eddy shape in Figure current micrograph 6.2 (the contrast is of Figure 6.15.b exhibits essentially the same fine details as the 40 MHz acoustic micrograph shown in Figure b) 125° from horizontal lift-off a) at horizontal lift-off Figure 6.15 | Eddy current c-scan images of a coarse-grained Ti-6A1-4V specimen at 5 MHz and two different rotational angles (0.5" x 0.5"). 174 6.2.c. Even be could better resolution at higher achieved although frequencies the decreasing sensitivity makes it increasingly difficult to visualize the rather weak (1-2%) conductivity variations grains. between Further in resolution improvements without sacrificing the sensitivity must come from further reducing the coil size. 6.5 Summary The main goal of this effort was to investigate the feasibility of exploiting the unique eddy current noise grain observed in titanium alloys the for purposes of nondestructive materials characterization. It was shown that electric anisotropy exhibited by noncubic crystallographic electromagnetic materials classes of materials can play a very similar role in characterization of polycrystalline metals to that of elastic anisotropy in ultrasonic materials characterization. Titanium is one of the few structural metals of practical importance, especially in aerospace applications, which preferentially crystallizes in hexagonal symmetry and therefore exhibits strong electrical anisotropy. At the same time, titanium tends to form a rather coarse, locally textured microstructure characterized by the presence of large colonies of hexagonal alpha grains featuring similar orientation. The fracture resistance of the material is strongly affected by this microstructure therefore there is a continued need for new nondestructive evaluation techniques that are capable of both imaging and quantitatively characterizing this microstructure. During the previous year, we already demonstrated that the strong grain noise in titanium alloys is caused by the unusual electrical anisotropy of hexagonal alpha grains and theoretically explained the physical mechanism responsible for this contrast. In this year, we further developed our eddy current imaging technique with special emphasis 175 on improving the imaging Higher resolution. resolution current eddy micrographs allowed us to determine the types of microstructure that most seriously hinder fatigue crack detection and to demonstrate that eddy current imaging can be used to detect locally textured microstructures that adversely affect fatigue resistance. In pursuit of even better resolution we used analytical, finite element simulation, and experimental methods to investigate the lateral resolution of eddy current microscopy and developed optimization procedures that assure the best imaging resolution for microstructural materials characterization. It was found that the lateral resolution of eddy current imaging is ultimately limited by the probe-coil dimension, but both the inspection frequency and the phase angle can be used to optimize the resolution, to some degree, at the expense of sensitivity. Generally, the higher the inspection frequency, the higher the imaging resolution, but the improvement was found to be rather modest. Although eddy current imaging is still in its infancy, a direct comparison of 5-MHz eddy current and 40MHz acoustic microscopic images of the same coarse-grained Ti-6Al-4V sample indicated that the same features can be observed by both methods at approximately the same resolution level. It is expected that in the future eddy current microscopy can be further enhanced by the introduction of special high-resolution, microscopic probe-coils which will make it a truly unique materials characterization tool especially well suited for microstructural evaluation of titanium alloys. * TI would like to express my sincere thanks to Dr. Waled Hassan (United Technologies Research Center) for his expert support with the analytical calculations and finite element results. Chapter 6, References: 1. J. H. Rose, "Ultrasonic backscatter from microstructure,” in Rev. Progr. Quant. Nondestr. Eval. Vol. 11B (Plenum, New York, 1992) pp. 1677-1684. 2. F. J. Margetan, R. B. Thompson, and I. Yalda-Mooshabad, "Backscattered microstructural noise in ultrasonic toneburst inspections," J. Nondestr. Eval. 13, 111-136 (1994). 3. K. Goebbels, "Structure analysis by ultrasonic radiation," in Research Techniques in Nondestructive Testing Vol. TV (Academic, New York, 1980) pp. 87-157. 4. A. Hecht, R. Thiel, E. Neumann, and E. Mundry, "Nondestructive determination of grain size in austenitic sheet by ultrasonic backscattering,” Mat. Eval. 39, 934-938 (1981). 5. H. Willems and K. Goebbels, "Characterization of microstructure by backscattered ultrasonic waves," Met. Sci. 15, 549-553 (1981). 6. C. B. Guo, P. Holler, and K. Goebbels, "Scattering of ultrasonic waves in anisotropic polycrystalline metals," Acustica 59, 112-120 (1985). 7. W. A. Wooster, A Textbook on Crystal Physics (University Press, Cambridge, 193 8). 8. J. F. Nye, Physical Properties of Crystals, Their Representation by Tensors and Matrices (Clarendon Press, Oxford, 1985). 9. J. E. Neighbor, "Eddy current method for measuring anisotropic resistivity," J. Appl. Phys. 40, 3078-3080 (1969). 10. V. M. Tatarnikov, "Use of probes for measuring the electrical conductance of anisotropic plates," Meas. Tech. 13, 877-881 (1970). 11. V. IL. Gordienko and V. G. Rybachuk, "Device for measuring the anisotropy of electrical conductivity," Meas. Tech. 32, 371-374 (1989). 12. R. Pitchumani, P. K. Liaw, D. C. Yao, D. K. Hsue, and H. Jeong, "Eddy current technique for the measurement of constituent volume fractions in a three-phase metalmatrix composite," J. Comp. Mat. 28, 1742-1769 (1994). 13. R. Pitchumani, P. K. Liaw, D. C. Yao, D. K. Hsue, and H. Jeong, "Theoretical models for the anisotropic conductivities of two-phase and three-phase metal-matrix composites," Acta. Metal. Mat. 43, 3045-3059 (1995). 14. P. L. Rossitier, The Electrical Resistivity of Metals and Alloys (Cambridge University Press, Cambridge, 1987) pp. 185-195. 15. D. Eylon, "Faceted fracture in beta annealed titanium alloys," Met. Trans. A 10, 311317 (1979). 16. “Device Eases Skin Crack Detection,” Aviat. Week Space Tech. Vol. 131, p. 137 (1989). 17. “Eddy Current Imaging Device Offers Prospect of Simple Crack Detection,” Aviat. Week Space Tech. Vol. 132, p. 96 (1990). 18. N. C. Haywood and J. R. Bowler, “Eddy Current Imaging of Buried Cracks by Inverting Field Data,” JEEE Trans. Magn. Vol. 28, p. 1336 (1992). 19. T. W. Guettinger, K. Grotz, and H. Wezel, “Eddy Current Imaging,” Mat. Eval. Vol. 51, p. 444 (1993). 20. G. L. Fitzpatrick, D. K. Thome, and R. L. Skaugset, “Magneto-optic/Eddy Current Imaging of Aging Aircraft; a new NDI technique,” Mat. Eval. Vol. 51, p. 1402 (1993). 177 21. J. H. Hippler, H. Ermert, and L. von Bernus, “Broadband Holography Applied to Eddy Current Imaging by Using Signals with Multiplied Phases,” J. Nondestr. Eval. Vol. 12, p. 153 (1993). 22. M. Gramz, “Eddy Current Imaging - A Theoretical Approach,” Compel. Vol. 13, p. 19 (1994). 23. M. W. Kirby, “Experience with Eddy Current Imaging Detection for the Detection of Cracks within and under the Clad in a Pressure Vessel,” Mat. Eval. Vol. 54, p. 30 (1996). 24. E. A. Ash and G. Nichols, “Super-resolution Aperture Scanning Microscope,” Nature 237, pp. 510-512 (1972). 25. C. P. Viahacos, R. C. Black, S. M. Anlage, A. Amar, and F. C. Wellstood, “Near-field Scanning Microwave Microscope with 100 pm Resolution,” Appl. Phys. Lett. Vol. 69, pp. 3272-3274 (1996). 26. D. E. Steinhauer, C. P. Vlahacos, S. K. Dutta, F. C. Wellstood, and S. M. Anlage, “Surface Resistance Imaging with Scanning Near-field Microwave Microscope,” Appl. Phys. Lett. Vol. 71, pp. 1736-1738 (1997). 27. D. E. Steinhauer, C. P. Vlahacos, S. K. Dutta, B. J. Feenstra, and F. C. Wellstood, “Quantitative Imaging of Sheet Resistance with a Scanning Near-field Microwave Microscope,” Appl. Phys. Lett. Vol. 72, pp. 861-863 (1998). 28. M. P. Blodgett, P. B. Nagy, “Anisotropic Grain Noise in Eddy Current Inspection of Noncubic Polycrystalline Metals,” Appl. Phys. Let., 72(9), pp. 1045-1047 (1998). 29. A. Briggs, An Introduction to Scanning Acoustic Microscopy (Oxford University Press, Oxford, 1985). pp. 35-41. 30. G. T. Meaden, Electrical Resistance of Metals (Plenum, New York, 1965) p. 31. 31. H. L. Libby, Introduction to Electromagnetic Nondestructive Test Methods (WileyInterscience, New York, 1971). 32. Z. Mottl, “The Quantitative Relations Between True and Standard Depth of Penetration for Air-cored Probe Coils in Eddy Current Testing,” NDT Intern. Vol. 23, pp. 11-18 (1990). 33. F. Thollon, B. Lebrun, N. Burais, and Y. Jayet, “Numerical and Experimental Study of Eddy Current Probes in NDT of Structures with Deep Flaws” NDT&E Intern. Vol. 28, pp. 97-102 (1995). 34. C. V. Dodd and W. E. Deeds, “Analytical Solutions to Eddy Current Probe-coil Problems,” J. Appl. Phys. Vol. 39, pp. 2829-2838 (1968). 35. C. V. Dodd, “The Use of Computer Modelling for Eddy Current Testing,” in Research Techniques in NDT Vol. 3, Ch. 13 (Academic Press, London, 1977). 36. R. Palanisamy and W. Lord, “Prediction of Eddy Current Probe Signal Trajectories,” IEEE. Trans. Magn. Vol. 16, pp. 1083-1085 (1980). 37. R. Becker, K. Betzold, K. D. Boness, R. Collins, C. C. Holt, and J. Simkin, “The Modeling of Electrical Current NDT Methods: Its Application to Weld Testing (Part 1),” British. J. NDT Vol. 28, pp. 286-294 (1986). 38. P. Stucky and W. Lord, “Skin Depth Considerations in Eddy Current NDT,” in Rev. Progr. Quant. Nondestr. Eval. Vol. 11A (Plenum, New York, 1992) pp. 299-306. 39. PC-OPERA for Windows, Vector Fields, Inc. (Aurora, Illinois, 1998). 40. D. J. Hagemaier, “Eddy Current Standard Depth of Penetration,” Mat. Eval. Vol. 43, pp. 1438-1442 (1985). CHAPTER VII CONCLUDING REMARKS 7.1 Summary effort was A comprehensive undertaken to develop evaluation nondestructive methods to characterize microstructure and material inhomogeneities in Ti-6Al-4V. The influence of texture, grain structure, and crystallographic orientation on ultrasonic wave propagation, electrical conductance, and microstructure anomaly detection in heat treated investigated. samples was received material types microstructure establish a and baseline Samples were generated additional samples were (generated via heat treatments for this study. these studies in the as- generated to make Joined in five of the different common as-received material) samples of microstructure the to same composition, but dissimilar microstructures were also generated by diffusion bonding to simulate processing anomalies in the axial and radial orientations. The metallographic results indicate the presence of strong macroscopic texture in the as-received alloy, exhibiting alignment of the grains, grain boundaries, and grain colonies, which is a key factor in assessing flaw (anomaly) detectability due to the generation of microstructural background noise. The heat treated samples showed the 179 these diminution of the macroscopically aligned grain features in the microstructure, but samples maintain most of the macroscopic texture, except in the coarse beta annealed the two samples. All of the samples from this study were axisymmetric and evaluated in a principal directions (i.e. axial and radial). Only the beta annealed samples demonstrated also reduction in the texture, due to recrystallization. Ultrasonic velocity measurements the basal indicated the presence of a strong preferred crystallographic orientation in which by normals preferentially rotate to the transverse direction. This finding is supported analytical x-ray results, using Shultz back-reflection. The orientation dependence of ultrasonic scattering and attenuation in the textured studied bar material was also investigated. Conventional ultrasonic backscattering was tering and a forward scattering method was developed to assess texture. The backscat is not results indicate defect detection in the radial direction of the as-received bar stock in the favorable due to the strongest presence of ultrasonic grain noise. Backscattering is axial direction is negligible; however, the ultrasonic attenuation in the radial direction current significantly lower than in the axial direction, a behavior not predicted from t theoretical scattering models. Typically, the direction of highest attenuation is consisten broken with the direction of highest backscattering energy. This common rule is clearly reasoning for the case of the Ti-6Al-4V forged bar material studied in this work. The of behind the unusual properties was demonstrated to be related to the forward scatter be ultrasonic energy in the elongated microstructure. This work shows that scatter must considered from multiple directions, due to multiple scattering, in order to relate scatter with attenuation. as measured significantly misleading due to the phase-sensitivity receivers. This result was verified based on the shown was means, conventional using attenuation, The of the wide-aperture to be ultrasonic of a laser ultrasonic development detection method, which removes the phase-sensitivity of the receiver by focusing the light down to a 50 pm spot, a subwavelength aperture. The laser detection system allows mapping of the phase and amplitude of waves transmitted in a solid. This high resolution mapping capability revealed for the first time, images of microstructural features based on phase distortion of the propagating waves. An eddy current materials characterization technique was also developed to map electrical conductance variations from resulting the grain anisotropy and hexagonal symmetryof the Ti-6AIl-4V alloy. This research demonstrated for the first time, an eddy current materials characterization capability geared specifically towards titanium alloys to map electrical properties of the macroscopic grain structure. An optimization of the electrical conductivity mapping technique demonstrated a factor of two increase in spatial resolution by increasing the AC frequency. Further improvements were achieved by appropriately rotating the phase angle; however, this research clearly shows the primary factor influencing the resolution is the diameter of the probe. This research has lead to new research in development of nearfield scanning microwave microscopy, which is promising for further improving the spatial Similarly, while the electrical mapping microstructure characterization, it has resolution technique was lead to new magnetoresistance sensors for deeper inspection depths. of eddy current microscopy. geared towards near surface interests in the development of Finally, an investigation was conducted to detect and characterize anomalous microstructure, using diffusion bonded samples containing microstructure inserts of three different sizes (i.e., 0.9", 0.4" and 0.2"). The results demonstrate that the anomalous microstructure inserts can be detected via ultrasonic backscattering in all but the smallest dimension. The correlation parameter in making length of the grain scattering was an assessment shown of microstructural uniformity. to be the key Unfortunately, the scattering signals lack uniqueness, are frequency dependent, and are difficult to interpret in terms of actual physical features within the material, due to signal interference and multiple scatter. Hence, these ultrasonic backscattering results are qualitative, and lack the specifics to allow a comprehensive description of the microstructure characteristics (e.g., grain size, shape, anisotropy, etc.). 7.2. Conclusions General 1) Background ultrasonic and eddy current measurements have revealed unusual elastic and electrical properties in the Ti-6Al-4V forged bar stock. What is unusual is the simultaneous high attenuation and low backscattering in the axial direction of the forged bar stock, due to forward scattering; and the fluxuating local electrical conductivity, due to the hexagonal symmetry of the alpha phase These measurements also showed the material contains significant deformation and crystallographic textue, and this fact was verified using x-ray diffraction. 182 Ultrasonic 2) backscattering results measurement show 0.9” the and 0.4” subsurface anomalies are detectable, but the 0.2” anomaly inserts are undetectable. These measurements demonstrated the difficulty involved with finding small localized interior microstructure anomalies in forged titanium alloys, especially in the noisy (i.e radial) direction. 3) Three new NDE methods have been developed and demonstrated during this dissertation to characterize titanium alloys. These methods are: i) forward scattering for qualitative assessment assessment and evaluation significantly influences ii) ultrasonic of texture, of distortions attenuation; phase mapping the propagating in for grain and ili) electrical anisotropy, alignment wavefront, which using eddy current evaluation, for mapping of macroscopic grain structure. 4) The new NDE methods represent a sound experimental foundation for anomaly detection and characterization in titanium alloy products. 5) This work should be continued to further enhance the NDE capabilities for titanium alloy billet, bar, and plate. Ultrasonic Phase Mapping 1) Conventional ultrasonic measurements anisotropy. 183 show evidence of strong elastic 2) Phase mapping allows texture-induced wafefront disturbances to be visualized. 3) The sample orientation (i.e. described by the angle between the surface normal and the bar axis) has a primary effect on ultrasonic properties (e.g. attenuation, scattering, and velocity). 4) Phase cancellation is mainly evident in the axial direction of the Ti-6Al-4V forged bar stock; the transverse direction has much smaller degree of distortion. 5) Phase cancellation at the receiver (in conventional wide-aperture transducers) significantly influences attenuation and can generate misleading results for Ti alloys. Eddy Current Microscopy 1) Eddy current measurements revealed strong electrical anisotropy in the Ti-6Al4V alloy due to the hexagonal nature of the alpha phase. This study revealed the difficulty in detecting small flaws in titanium alloys due to strong background noise. 2) Good agreement was demonstrated between the analytical, finite element, and experimental results on the eddy current lateral resolution investigation. 3) The lateral resolution is primarily driven by the eddy current probe diameter. An increase in the AC frequency can also improve the resolution by about a factor of two and phase angle can improve the resolution, but these are both secondary effects. 7.3 Comparison of Results for Microstructure Anomaly Characterization Multiple NDE conventional ultrasonic NDE techniques and backscatter) and some newly developed including forward scattering, laser interferometric phase mapping, and eddy These dissertation. (measurements methods, approaches have been investigated throughout the course of this of velocity, current microscopy Conventional approaches attenuation, (see Table ultrasonic include 7.1 — Comparison approaches provide of NDE a baseline Results, capability to Appendix A). evaluate the influence of texture and microstructure on the elastic properties. Despite the fact that these techniques are not new, they are invaluable and provide unique insight into the structural make-up of polycrystalline Specifically, metals. ultrasonic velocity measurements (longitudinal and shear wave velocity) provide a means to evaluate texture in a quantitative manner. For example, the forged bar investigated in this dissertation clearly contained a significant degree of mechanical texture and the ultrasonic velocity measurements supported this fact, which was later confirmed via X-ray measurements (Shultz back reflection) showing the basal normals of the forged bar tend to preferentially rotate to the transverse direction during fabrication. Conventional ultrasonic attenuation and backscattering measurements were also taken to provide information on the micro- and macroscopic structural makeup of the alloy. The attenuation measurements provide an indication of the impedance to propagating waves as a function of length, but do not specify what material features are to responsible for the decay in signal amplitude. A number of factors have been shown affect attenuation in this dissertation, including phase sensitivity of the receiver, surface 185 texture interactions, the grain orientation, and grain size, and shape. Unfortunately, attenuation is a bulk measurement and all of these interactions combine unseperably in what we call attenuation. Ultrasonic backscattering measurements are generally not more specific in terms of their behavior (compared to attenuation) and are also difficult to interpret due to competing and effects related to elastic property crystallographic variations, grain and colony boundaries, frequency dependence, and multiple scatter. In titanium alloys, this problem is even more complicated by second phase particles and density their associated differences constituent, relative to the primary phase which generates another source of scattering. Moreover, conventional scattering theory (based on the Rayleigh, stochastic, and geometric scattering regimes) is not geared for the complexity of most titanium alloys. However, recent research demonstrates that by taking the two-point elastic correlation, as measured by crystallographic mapping of the grain a much improved prediction structure, provides complex microstructures, backscatter measurements characteristic, similar to attenuation. While these for backscattering. provide Generally, for only an overall response measurements (attenuation and backscatter) are not specific, they are useful for comparison purposes to observe relative differences between effects of microstructure, texture, and manufacturing process. The forward scattering results described in this dissertation add another useful bulk measurement tool for microstructure characterization. This tool is particularly useful for characterization of macroscopic texture, based on the amplitude and divergence of the forward scatter. The results indicate this tool is useful for separating textured versus random structures. 186 Phase mapping, using laser interferometric ultrasonic detection, was demonstrated to be useful to characterize the influence of macroscopic orientation and the resulting phase perturbation on attenuation in the axial and transverse directions of a mill annealed Ti-6Al-4V bar. These phase mapping results revealed approximately 30% higher phase scatter, measured over the center fringe in the axial direction, relative to the transverse direction. The higher degree of phase scattering in the axial direction of the bar is due to the presence of elongated macroscopic inhomogeneities aligned in the same direction. These extended colonies of similarly oriented grains in the axial direction result in local disturbances in the arrival time of the propagating wavefront that extend beyond the finite aperture of the transmitter. While there also exists phase scatter in the radial direction, it is significantly smaller and more uniform than that of the axial direction. Measurements also revealed substantially lower attenuation values for the laser experiment due its relative phase-insensitivity, provided by the microscopic footprint of the focused laser detector. While the laser interferometric measurements demonstrate lower attenuation than the immersion results, the attenuation is still clearly dominant in the axial direction, as compared with the transverse direction. These measurements demonstrate that the wider apertures generally used by conventional transducers clearly suffer phase cancellation losses, resulting in significantly higher attenuation results. The attenuation in this material is unusual considering that the backscatter is 2 to 3 times stronger in the radial direction than in the axial one; and the forward scatter is at least as strong, although less divergent, than in the axial direction. This is because elongated grains and grain colonies present 187 larger total scattering cross-section (larger effective size) along their axis compared to normal to them, but backscattering and forward scattering coefficients are actually higher in the normal direction. Therefore, despite the stronger backscatter in the radial direction, the axial direction has higher attenuation. This method made it possible to image the and microstructure by using the laser detection technology. The macroscopic orientation directionality can be clearly identified and possibly quantified. Work is now in progress to apply the same techniques to other Ti-6Al-4V microstructures with varying degrees of structural directionality, to confirm this finding. Finally, was microscopy current eddy demonstrated as the result of an investigation on the feasibility of exploiting the unique eddy current grain noise observed in titanium alloys for the purposes of nondestructive materials characterization. It was shown that electric anisotropy exhibited by noncubic crystallographic classes of materials can play a very similar role in electromagnetic materials characterization of polycrystalline metals to that of elastic anisotropy in ultrasonic materials characterization. Titanium is one of the few structural metals of practical importance, especially in aerospace applications, which preferentially crystallizes in hexagonal symmetry and a therefore exhibits strong electrical anisotropy. At the same time, titanium tends to form rather coarse, locally textured microstructure characterized by the presence of large colonies of hexagonal alpha grains featuring similar orientation. The fracture resistance of the material is strongly affected by this microstructure therefore there is a continued need for new nondestructive evaluation techniques that are capable of both imaging and quantitatively characterizing this microstructure. 188 In pursuit of better resolution analytical, finite element simulation, and experimental methods were used to investigate the lateral resolution of eddy current best imaging resolution for microstructural that assure the characterization. It was found that the lateral resolution of eddy microscopy materials current imaging is ultimately limited by the probe-coil dimension, but both the inspection frequency and the phase angle can be used to optimize the resolution, to some degree, at the expense of sensitivity. Generally, the higher the inspection frequency, the higher the imaging resolution, but the improvement was found to be rather modest. Although eddy current imaging is still in its infancy, a direct comparison of 5-MHz eddy current and 40-MHz acoustic microscopic images of the same coarse-grained Ti-6Al-4V sample indicated that the same features can be observed by both methods at approximately the same resolution level. It is expected that in the future eddy current microscopy can be further enhanced by the introduction of special high-resolution, microscopic probe-coils which will make it a truly unique materials characterization tool especially well suited for microstructural evaluation of titanium alloys. 7.4 7.4.1 Future Nondestructive Materials Characterization Research Application and Development of Array Transducers One area that will be beneficial for future study is the development and use of phased array transducers. The main problem of detecting anomalous scatterers in bulk material is that only one incidence angle is generally investigated (i.e. normal incidence). 189 Phased arrays enable the capability to probe the material for obscure (nonspecular) scatterers from a multitude of incidence angles, thereby increasing the probability of detection. Unfortunately, these array systems have a high capital equipment investment barrier that makes their application difficult for research. However, with the increased need to find and characterize smaller and smaller internal flaws, especially in fracturecritical engine equipment. components, Phased this approach array systems is viable, could provide despite the large investment a tremendous in increase in inspection throughput for critical aircraft airframe and engine components. Moreover, the electronic beam steering and dynamic focusing offered by phased arrays makes their use a potential for improved billet inspection in the titanium industry. 7.4.2 Geometry Insensitive Techniques One of the key difficulties in the nondestructive inspection of aircraft components is due to their geometric complexity. Some recent developments in thermoelectric materials evaluation have demonstrated the capability to discriminate between varying degrees of damage (simulated by varying degrees of plastic deformation) without the strong sensitivity to shape or surface character that some practical methods, like eddy current inspection, demonstrate. The technique is based on the measurement of magnetic field variations resulting from thermal currents in a conductor imposed by an external thermal gradient (i.e., magnetic sensing of the Seebeck effect). While little research has been done to describe the details of what may be possible in this area, the preliminary research suggests this could be an extremely valuable NDE method. One application that would represent a breakthrough is the demonstration of a capability to nondestructively 190 would a play role strong in assessment capability A fatigue damage provide local assessments of fatigue damage. better developing prediction life cycling mechanical of thermal and tools and general on fracture critical understanding of the effects components. While this capability would likely be geared towards more macroscopic measurements, on the order of 0.25" spatial resolution, the benefits of observing the progress of damage without destructive microscopic initiation and techniques would premature to judge as the fundamental experimental foundation still needs to be built. Clearly, it will take some unparalleled. be at this point The implications of this capability and examination are a little theoretical time to lay down the groundwork, make some investments in instrumentation, and develop practical methods. 7.4.3 Residual Stress Gradient Measurement This research is directed towards the capability to nondestructively measure residual stress depth gradients, which is a goal for improved life prediction practices. NDE research is needed to enable the capability to measure approximately 400 ym stress depth profiles to in shot peened materials and deeper in laser shock peened materials. Currently residual stress gradients can be measured, but only destructively by surface material and measuring the stress by x-ray diffraction. successively removing Scanned x-ray diffraction measurement measurements over a small area, but capabilities also exist, which provide stress very slow. Ultrasonic stress measurement capabilities also exist, based on the acoustoelastic effect, but these techniques generally only provide an evaluation of the bulk material, similar to an attenuation measurement. Local stress measurements using scanning acoustic microscopy (i.e., V(z) measurements 191 of the Rayleigh wave or longitudinal surface skimming wave) have been demonstrated to be stress sensitive, but further research is necessary to describe the limits and prove the approach. Similar to velocity, electrical measurements are marginally sensitive to stress and cold working. Researchers are currently investigating the potential of frequency dependent local acoustic velocity measurements and eddy current stress measurements. Unfortunately, the surface roughness occurring as a result of shot / laser shock peening is clearly a troublesome hurdle to these approaches. As such, other approaches should be simultaneously investigated to overcome the obvious limitations of the ultrasonics and Alternate approaches may electrical approaches. include: alternating current potential difference, thermal mapping, thermoelectric, and energy dependent x-ray. It's not obvious that we will ever be able to achieve a nondestructive stress depth profiling capability, so this is clearly a high risk research area, but the payoff for improved life prediction / life extension requires us to do this research. 7.4.4 Forward Scattering Using Dual-Transducer Articulation Finally, preliminary research results from this thesis project have demonstrated the some capabilities of ultrasonic forward scattering to characterize bulk material inhomogeneities. However, further research is needed to detail some of the fundamental aspects of forward scattering, such as experimentally describing the scattering crosssections in Ti-6Al-4V. The majority of past efforts have concentrated on ultrasonic backscattering, which is a well investigated area, due to the vast amounts of literature currently available. More details in forward research in ultrasonic materials characterization. scattering is an opportunity to further APPENDIX A - Data Tables Sample Description Nomenclature As-received (0 degrees relative to the axial AR(0) direction of the forged bar) As-received (30 degrees relative to the axial direction of the forged bar) AR(30) As-received (37 degrees relative to the axial direction of the forged bar) AR(37) As-received (45 degrees relative to the axial direction of the forged bar) AR(45) As-received (60 degrees relative to the axial direction of the forged bar) AR(60) As-received (90 degrees relative to the axial direction of the forged bar) AR(90) Duplex anneal (axial direction) DA(0) Duplex anneal (transverse direction) DA(90) Recrystallization anneal (axial direction) RA(0) Recrystallization anneal (transverse direction) RA(90) Fine beta annealed (axial direction) FBA(0) Fine beta annealed (transverse direction) FBA(90) Coarse beta annealed (axial direction) CBA(0) Coarse beta annealed (transverse direction) CBA(90) Table 3.2 Ti-6AI-4V sample nomenclature for ultrasonic measurements. Sample descriptions are based on the microstructure and the orientation of wave propagation. Longitudinal Wave Velocity Table 3.3 Sample Velocity (m/s) Difference (%) AR(90) 6248.5 AR(0) 6131.3 DA(90) DA(0) 6230.8 6110.1 1.93 RA(90) 6234.7 1.70 RA(0) 6128.3 FBA(90 (90) FBA(0) CBA(90) 6200.8 6150.8 6197.4 CBA(0) 6173.9 187 0.80 038 Microstructural effects on longitudinal wave velocity, based on data from Figure 3.9. All samples are approximately 1.5” thick. Shear Wave Velocity Sample Table 3.4 | fast mode (m/s)| slow mode (m/s)|__ fast / slow AR(90) AR(0) 3319.57 3163.79 3164.39 3140.27 1.04903 1.00749 DA(90) DA(0) RA(90) RA(0) 3281.54 3158.39 1.03899 3140.01 3273.65 3153.03 3119.29 3143.27 3133.84 FBA(90) 3204.37 3165.33 1.00664 1.04148 1.00612 1.01235 FBA(0) CBA(0) CBA(90) 3181.85 3178.00 * 3158.34 3168.00 * 1.00744 1.00315 * Microstructural effects on shear wave velocity, based on data from Figure 3.12. All samples are approximately 1.5" thick. 194 to get reproducible results) attenuative too * Data not available (sample was [je Ul rojoRZ & Ape] st UoNeSedoid saeM oY} JO UOTRIUSO "Wao }daoxa ‘soydures “sATENusye 4svo] oY} SEM (06) VC pue satjenusye ysour oy) ApIeaTD sem (06 PUe O HIG) VAD sduIeg “LTE aMBrJ Jo sy[nsor UO paseq SI BEC] “UOHRIOXS JsINq -quo} Suisn ‘yIBusT YEdioyem ,,L"] & YIM “AaJoweIP ,,¢°Q pesnoogun “ZHY| OT & Sew Joonpsuel] “gP ul sa0yoo L’81 S61 L1é TOr II col vil 6 Tl vl v8 Lg $6 6 Lgl LI ss Ls cL L’6 yuasyoo JUanbasqns u99M49q SSO] [BUSIS UO paseg ‘sesso] UOTENUS}e yuopuodap Aouanbey pue TemyonysoIoly| ysl (o)va (06) Vda (ova (06) UV (06)VdO (o)VdoO (qp) uonenuapy stuoseni S'€ BQRL ‘uoTOOIIp (sa1d0p 0) jeIxe oy} UeY}) Joneosyoeg JosUONs ATUBOIJTUSIS SayeIsUOWSp UOTAIIP (sa1Zap 06) aSISAsUeI} 9} ‘AT[eIBUAyH ‘Iayyeosyorq [RINJONsOIOIUN OU AjyenyA syussaidos YOTYM “AW €TE0'O 1% SUIZ9q P[OYSEAY} SSIoU [BOLO9[9 aU} OT & Sem Joonpsuell JY} SION ‘YISUaNs Zuployeosyorq JO Japio Burpusosep Ut sojdures YIM ‘ZE"¢ SINSIJ UO paseq SI eye] “UOTEWOX asindur Zuisn ‘ysucy yyedioyem ,,¢ & YIM sd"JINS oY} UO posndoy ‘JooWeIp ,.S°0 ‘ZH ‘SULIOJOARM (NOT ATayeuTxosdde Burdesoae Alyereds soye (sv) auI Jo UOTOUNY ev se BUTIOWeOSyoeQ OTUOSeNI/) sTeteo0 STETEO0 STETEO0 STETEO0 STETEOO STETEO0 STETEO'O SIETEOO STETEO0 809700 STETEO0 LOEvr0'0 STEIEO0 £tplv00 cIETEO0 CZIE90'O vL8S0'°0 Lv9S7l'0 8L6101°0 90L790°0 6S1$90'°0 9PPTLO'0 889200 605680°0 E1E8S0°0 9TT6170 CCPOPT 0: STELTI'0 eTclolo (06) Va STI6b7 0 ILbLee0 COPETH 0 S8STI810 9758070 86EP0E0 6r76~L0°0 vS€1800 6608L0°0 767880°0 vE8680°0 6LSS60°0 TEzeOr oO vIPror'o ILestloO 9096110 LOEVET 0 COPOST'O 9989910 ST6IL1°0 COL6P70 906LST°0 CSPI9T'O 91SL07'0 9°€ BIQRL STETEOO STETEO 0 STETEO0 eL6l110 STSSETO 8786910 (06) Va pySL90T'0 STETEO0 STETEO0 €S6£800 CLZOLT'0 (06)VdO 618870°0 chIELo'O Lysvso0 8086200 €67090'°0 8L7760'0 107890°0 (O6)UV pS9OSET 0 SsssIl oO STETEO'0 886010 (o)vao T7€8S0'0 7EPI9N'O L0S690°0 L6SS€0'0 679£0'0 69LS€0°0 v06rE0°0 6S9£0°0 elespo0'o (o)Vas (A Ul) SuLsyeOsWoe_ [BINjonsOIOI] posesoay Ayjeneds copa ope] mM] 869990'0 819€0'0 pL8ev00 (UV +l 9¢8LL0°0 L6EC80°0 v6Svr0'0 (o)va an] (06) Va (ova NI Axially Bonded Samples Table 4.1 Sample Thickness (in.) AR/ AR 1.79 AR/CBA 1.82 CBA/ CBA 2.33 AR/FBA/ AR 2.94 AR/CBA/ AR 2.65 Diffusion bonded samples oriented such that the ultrasound propagates in the axial direction. Radially Bonded Samples Table 4.2 Sample Insert Thickness (in.) AR/ AR AR/ AR AR/ AR AR/FBA AR/FBA AR/FBA 0.2 0.5 0.9 0.2 0.4 0.9 AR/ CBA AR/ CBA AR/ CBA 0.2 0.4 0.9 Diffusion bonded samples oriented such that the ultrasound propagates in the transverse direction. Samples were all 2.5" in diameter, with varying sizes of implants and implant microstructures. Orientation Dependence of Longitudinal Velocity Table 5.1 Sample Velocity (m/s) AR(0) 6131.32 AR(30) 6159.37 AR@G7) 6175.64 AR(45) 6193.10 AR(60) 6215.44 AR(90) 6248.52 Difference (%) 0.46 0.28 0.53 Orientation dependence of longitudinal wave velocity, based on data from Figure 5.2. All samples are approximately 1.5" thick. Note, the difference between the two extremes in orientation is approximately 2%. Scatter in the data is + 0.5%. Orientation Dependence of Shear Velocity Sample | fast mode (m/s)} slow mode (m/s)} __ fast / slow 3163.79 3217.53 3231.50 3243.28 3293.06 3319.57 AR(0) AR(30) AR(37) AR(45) AR(60) AR(90) Table 5.2 3140.27 3166.77 3173.00 3171.62 3161.31 3164.39 1.007 1.016 1.018 1.022 1.041 1.049 Orientation dependence of shear wave velocity, based on data from Figure 5.3. All samples are approximately 1.5" thick. Note, the difference between the fast versus slow modes is less than 1% in the axial (0 degree) direction and approximately 5% in the radial (90 degree) direction. This data supports the conclusion that the material is highly textured, with basal normals preferentially oriented in the transverse direction. Scatter in the data is + 0.5%. Orientation Dependent Attenuation Loss (dB) sample freq ty Table 5.3 | AR(O) | ARGO) | ARGS) | AR(60) | AR(90) 8.10 5.58 4.25 3.38 3.28 8.96 6.33 4.65 3.52 3.44 10.19 11.87 6.80 7.52 4.82 3.76 3.79 5.35 3.98 4.18 13.38 15.28 8.36 | 9.56 5.80 6.60 4.42 4.94 5.25 6.07 Orientation dependent attenuation losses, based on signal loss between subsequent coherent echoes in dB. Transducer was a 10 MHz, unfocused 0.5" diameter, with a 1.7" waterpath length. Data is based on results of Figure 5.4. Orientation Dependence of Backscattering (mV) AR(O) |AR(30) | AR(45) | AR(60) | AR(90) | 0.0448 | 0.0387 0.4162 0.0358 | 0.0357 | 0.0360 | 0.0372 0.2988 0.0353 | 0.0358 | 0.0348 | 0.0333 0.2508 | 0.0313 | 0.0313 | 0.0331 0.2047 0.0313 | 0.0313 | 0.0313 | 0.0320 0.1611 0.0313 | 0.0313 | 0.0313 | 0.0318 0.1561 0.0313 | 0.0313 | 0.0313 {0.0319 0.1358 0.0313 | 0.0313 | 0.0313 | 0.0308 0.1191 {| 0.0313 | 0.0310 0.1067 Oo [& [dy npn lL Lolr 0.0445 | 0.0442 Table 5.4 0.0313 0.0313 | 0.0313 Orientation dependence of ultrasonic backscattering as a function of time after spatially averaging approximately 2K waveforms at each position. Transducer was a 10 MHz, 0.5" diameter, focused on the surface with a 3" waterpath length. Data is based on Figure 5.5. Note that the electrical noise threshold begins at 0.0313 mV. 199 Phase-Sensitive (PS) -vs- Phase-Insensitive (PI) Attenuation (dB) sample freq (MHz Table 5.5 | AR(90)PS | AR(90)PI 4.65 4.02 4.61 4.05 4.46 4.10 AR(O)PS 8.86 9.68 10.56 | AR(O)PI 7.4 8.48 9.04 12.3 10.6 4.47 4.98 13.7 11.48 5.26 5.7 15.4 12.72 6.37 5.91 Phase-sensitive (0.5" aperture, piezoelectric) versus phase-insensitive (50 uum, laser interferometric) measurements of attenuation in dB, based on Figure 5.14. Note, a significant contribution of phase sensitivity to attenuation is only a factor in the axial (0 degree) direction; however, sample orientation is clearly the primary factor affecting attenuation in this material. Normalized Electrical Resistivity (%) Al Cu Cd Ti-6AI1-4V 0.1 -0.05 | -1.7. | data ranges Surface 2 | 0.15 -0.1 0.025 from 15 to 15 Surface 3 | 0.2 0.055 1.6 Surface 1] Table 6.1 Average electrical resistivity for three single crystal surface orientations, which were arbitrarily determined, and from the surface of polycrystalline Data is based on Figure 6.3. All data was calibrated using different aluminum alloys as the Ti-6Al-4V. measured and Note, the standard. differences in resistivity for the cubic crystals (Al and Cu) are well within the scatter of the measurement. However, the hexagonally symmetric Cd crystal has orientation dependent resistivity variations that extend well beyond the data scatter. 201 Frequency Dependent Lateral Resolution Frequency (MHz) Resolution (um) 0.01 1300 0.1 1175 900 725 1.0 10.0 Table 6.2 Frequency dependent lateral resolution, demonstrating approximately a factor of two improvement in resolution over three orders of magnitude in frequency. Data is based on Figure 6.11. Phase Angle Dependent Lateral Resolution Table 6.3 Phase Angle Resolution (degrees) 0 30 60 1 MHz (um) |2MHz | 3 MHz 876 736 660 |743 | 685 1655 _ | 615 |607 | 582 90 599 1558.5 | 541 496 1419 884 _|472.6 | 471 |1348 [1183 |746 | 687 Phase dependence of lateral resolution for three different frequencies over 180 degrees. Note, the cycle repeats for every 1 phase angle increment. This chart demonstrates that, similar to frequency, phase can lead to approximately a factor of slightly better than two in resolution, although at the expense of lift-off insensitivity. Data is based on Figure 6.14. Comparison of NDE Results Usage NDE Approach . Velocity Advantages Disadvantages Provides information on Accuracy of +,- 0.5% can be This ia a bulk measurement elastic property variations routinely achieved and velocity and does not allow for and texture resulting from is directly related to stiffness. . Attenuation Backscatter Forward Scatter” Can potentially provide Measurement is simple and information on grain size not time consuming. Phase Mapping” source of the loss and not to be compare materials. used for locating anomalies. Mainly used to characterize A single transducer is used as Measurement is complicated grain structure in simple, cubic transmitter and receiver. materials, can be used as a materials discriminator. different materials. Anomaly detection possible. and not unique. Used to qualitatively assess Data is in the form of an image the structure of metals. Allows insight into the No real advance for anomaly detection and experimental A good way to discriminate effective grain size and orientation effects Not suited to production environ. Phase-insensitive due to the narrow 50 micron aperture. Allows imaging of the wavefront, microstructure and phase distortion. Experimental system is complex and instrumentation intensive. Approach is for bulk wave evaluation of structures. Not for anomaly detection. Laser detection allows high resolution imaging of the transmitted acoustic field. Experimental system is complex and instrumentation intensive. Approach is for bulk wave Offers a unique way to describe the shape of a propagating ultrasonic wavefront and map m'structures phase distortion. Laser Based : : Amplitude Mapping A bulk measurement tool not specific in terms of the Allows for an easy way to and grain morphology. Allows for an basis to compare between textured-vs-random mat. Laser Based localization of anomalies except via surface waves. processing. Used to map the strength of ultrasonic waves propagating .* | through a solid. e ' ” and time consuming. | Data is frequency dependent set-up is complicated. evaluation of structures. Not for anomaly detection. ready Curr ent icroscopy Used to image local electrical conductivity variations in noncubic materials. Near surface detection of microstructure anomalies. Noncontacting, computer Surface and near surface only. controlled scanning of materials. | Resolution is primarily driven Resolution is at the physical limit. | by the coil diameter. Eddy current materials evaluation | Further probe developments are required to yield higher resolution is in its infancy. * New method, developed under this dissertation project. Table 7.1 Comparison of NDE results. 203