LASER ULTRASONIC THERMOELASTIC/ABLATION GENERATION WITH LASER INTERFEROMETRIC DETECTION IN GRAPHITE/POLYMER COMPOSITES

Similar documents
LASER GENERATION AND DETECTION OF SURFACE ACOUSTIC WAVES

NEW LASER ULTRASONIC INTERFEROMETER FOR INDUSTRIAL APPLICATIONS B.Pouet and S.Breugnot Bossa Nova Technologies; Venice, CA, USA

C. Edwards, A. AI-Kassim* and S.B. Palmer Department of Physics University of Warwick, UK

B.R. Tittmann, R.S. Linebarger and R.C. Addison, Jr.

AN ACTIVELY-STABILIZED FIBER-OPTIC INTERFEROMETER FOR

EFFECT OF SURFACE COATINGS ON GENERATION OF LASER BASED ULTRASOUND

STUDY ON SAW ATTENUATION OF PMMA USING LASER ULTRASONIC

RAPID INSPECTION OF COMPOSITES USING LASER-BASED ULTRASOUND

(1) LASER GENERATION OF "DIRECTED" ULTRASOUND IN SOLIDS USING SPATIAL AND TEMPORAL BEAM MODULATION

INSPECTION OF COMPONENTS HA VING COMPLEX GEOMETRIES. Andrew D. W. McKie and Robert C. Addison, Jr.

MEASUREMENT OF RAYLEIGH WAVE ATTENUATION IN GRANITE USING


CHAPTER 5 FINE-TUNING OF AN ECDL WITH AN INTRACAVITY LIQUID CRYSTAL ELEMENT

Detection of a Surface-Breaking Crack Depth by Using the Surface Waves of Multiple Laser Beams

SENSITIVITY OF AN EMBEDDED FIBER OPTIC ULTRASOUND SENSOR

MASSACHUSETTS INSTITUTE OF TECHNOLOGY Department of Electrical Engineering and Computer Science

STABILIZED FIBER OPTIC SENSOR FOR ULTRASOUND DETECI10N

LASER ULTRASONIC INSPECTION OF GRAPHITE EPOXY LAMINATES

la. Smith and C.P. Burger Department of Mechanical Engineering Texas A&M University College Station Tx

DETECTION OF LEAKY-RAYLEIGH WA YES AT AIR-SOLID INTERFACES BY

NUMERICAL MODELING OF AIR-COUPLED ULTRASOUND WITH EFIT. D. E. Chimenti Center of Nondestructive Evaluation Iowa State University Ames, Iowa, USA

A NON-CONTACT LASER-EMAT SYSTEM FOR CRACK AND HOLE

ULTRASONIC TECHNIQUES TO QUANTIFY MATERIAL DEGRADATION IN

ULTRASONIC GUIDED WAVES FOR AGING WIRE INSULATION ASSESSMENT

LAMB WA VB TOMOGRAPHY USING LASER-BASED ULTRASONICS

Evaluation of Scientific Solutions Liquid Crystal Fabry-Perot Etalon

Photorefractive Ultrasonic Sensor for Weld Quality Monitoring

Periodic Error Correction in Heterodyne Interferometry

DETERMINATION OF TIlE ABSOLUTE SENSITIVITY LIMIT OF A PIEZOELECfRIC

OPTICAL FIBER-BASED SENSING OF STRAIN AND TEMPERATURE

Theory and Applications of Frequency Domain Laser Ultrasonics

A NEW APPROACH FOR THE ANALYSIS OF IMPACT-ECHO DATA

FIDELITY OF MICHELSON INTERFEROMETRIC AND CONICAL PIEZOELECTRIC

IMPROVED LASER INTERFEROMETRY FOR ULTRASONIC NDE

ASSESSMENT OF WALL-THINNING IN CARBON STEEL PIPE BY USING LASER-GENERATED GUIDED WAVE

Non-contact Photoacoustic Tomography using holographic full field detection

FATIGUE CRACK CHARACTERIZATION IN CONDUCTING SHEETS BY NON

DAMAGE DETECTION IN PLATE STRUCTURES USING SPARSE ULTRASONIC TRANSDUCER ARRAYS AND ACOUSTIC WAVEFIELD IMAGING

A miniature all-optical photoacoustic imaging probe

Absolute distance interferometer in LaserTracer geometry

Spring 2004 M2.1. Lab M2. Ultrasound: Interference, Wavelength, and Velocity

William R. Scott, Stephen Huber*, and Martin Ryan

Laser Lock-in Thermal Wave Imaging for Nondestructive Evaluation

SA210-Series Scanning Fabry Perot Interferometer

SINGLE-FREQUENCY PULSED LASER OSCILLATOR AND SYSTEM FOR LASER-ULTRASONICS

R. J. Jones Optical Sciences OPTI 511L Fall 2017

Doppler-Free Spetroscopy of Rubidium

Experimental Physics. Experiment C & D: Pulsed Laser & Dye Laser. Course: FY12. Project: The Pulsed Laser. Done by: Wael Al-Assadi & Irvin Mangwiza

Generation Laser Scanning Method for Visualizing Ultrasonic Waves Propagating on a 3-D Object

Stabilizing an Interferometric Delay with PI Control

D.C. Emmony, M.W. Godfrey and R.G. White

Measurement of phase velocity dispersion curves and group velocities in a plate using leaky Lamb waves

Sonic Distance Sensors

Development of a Low Cost 3x3 Coupler. Mach-Zehnder Interferometric Optical Fibre Vibration. Sensor

CONTACT LASER ULTRASONIC EVALUATION OF CONSTRUCTION MATERIALS

DEVELOPMENT OF HEAT-RESISTANT OPTICAL FIBER AE SENSOR

Lamb Wave Ultrasonic Stylus

SECOND HARMONIC GENERATION AND Q-SWITCHING

Setup of the four-wavelength Doppler lidar system with feedback controlled pulse shaping

R. D. Huber and G. H. Thomas

Laser Locking with Doppler-free Saturated Absorption Spectroscopy

PCS-150 / PCI-200 High Speed Boxcar Modules

Influence of the anisotropy on zero-group velocity Lamb modes

ULTRASONIC DETECTION OF FATIGUE CRACKS BY THERMO-OPTICAL

Laser Beam Analysis Using Image Processing

Laser Telemetric System (Metrology)

Basics of INTERFEROMETRY

AIR-COUPLED ULTRASONIC TESTING OF MATERIALS

improved stability (compared with

First and second order systems. Part 1: First order systems: RC low pass filter and Thermopile. Goals: Department of Physics

3 General Principles of Operation of the S7500 Laser

Experiment 1: Fraunhofer Diffraction of Light by a Single Slit

Capabilities of Flip Chip Defects Inspection Method by Using Laser Techniques

Synchronization in Chaotic Vertical-Cavity Surface-Emitting Semiconductor Lasers

High-Power, Passively Q-switched Microlaser - Power Amplifier System

Dispersion measurement in optical fibres over the entire spectral range from 1.1 mm to 1.7 mm

Wavelength Control and Locking with Sub-MHz Precision

Ultrasonic Testing using a unipolar pulse

ULTRASONIC GUIDED WAVE ANNULAR ARRAY TRANSDUCERS FOR STRUCTURAL HEALTH MONITORING

OPTI510R: Photonics. Khanh Kieu College of Optical Sciences, University of Arizona Meinel building R.626

Photon Count. for Brainies.

Lab 12 Microwave Optics.

Silicon Photodiodes - SXUV Series with Platinum Silicide Front Entrance Windows

Unit-23 Michelson Interferometer I

LOPUT Laser: A novel concept to realize single longitudinal mode laser

DEFECT CHARACTERIZATION IN THICK COMPOSITES BY ULTRASOUND. David K. Hsu and Ali Minachi Center for NDE Iowa State University Ames, IA 50011

NON-AMPLIFIED PHOTODETECTOR USER S GUIDE

Basics of INTERFEROMETRY

DIODE LASER SPECTROSCOPY (160309)

Multiply Resonant EOM for the LIGO 40-meter Interferometer

visibility values: 1) V1=0.5 2) V2=0.9 3) V3=0.99 b) In the three cases considered, what are the values of FSR (Free Spectral Range) and

Ultrasonic Infrared Thermal Wave Technology and Its Applications in. Nondestructive Evaluation

Lecture 9 External Modulators and Detectors

USE OF GUIDED WAVES FOR DETECTION OF INTERIOR FLAWS IN LAYERED

FRAUNHOFER AND FRESNEL DIFFRACTION IN ONE DIMENSION

Physical Optics. Diffraction.

EXPERIMENTAL TRANSFER FUNCTIONS OF PRACTICAL ACOUSTIC EMISSION SENSORS

Exp 3 COLCULATE THE RESPONSE TIME FOR THE SILICON DETECTOR

An acousto-electromagnetic sensor for locating land mines

Research Article An Investigation of Structural Damage Location Based on Ultrasonic Excitation-Fiber Bragg Grating Detection

Transcription:

LASER ULTRASONIC THERMOELASTIC/ABLATION GENERATION WITH LASER INTERFEROMETRIC DETECTION IN GRAPHITE/POLYMER COMPOSITES INTRODUCTION James N. Caron and James B. Mehl Department of Physics University of Delaware Newark, DE 19716-2570 Karl V. Steiner Center for Composite Materials University of Delaware Newark DE 19716-3144 Ultrasonic signals have been generated and detected in graphite/polymer composites by optical methods. A Doppler interferometric technique was used for detection. The output voltage of this type of interferometer is proportional to the surface velocity of a sample area which is illuminated by cw laser light. Ultrasonic signals were generated by thermoelastic and ablation processes which occur as a consequence of laser pulses incident on the opposite surface of the sample. The evolution of the magnitude and shape of the detected signals was measured as a function of the pulse energy of the generating laser. Low-energy laser pulses generated ultrasound without causing obvious surface damage. At higher energies surface damage was observable in post inspection but could also be detected by observing (through protective goggles) bright flashes near the illuminated area. The energy at which these processes first occur is qualitatively referred to as the ablation threshold. Changes in the observed waveform were evident at energies above the ablation threshold. The higher-energy waveforms were found to consist of a superposition of a thermoelastic component and an ablatic component, whose relative magnitudes changed with laser power. A delay in the initiation of the ablatic wave relative to the thermoelastic wave was observed to be of the order of 0.3 Its, consistent with observations in pure polymer. [1] Photoelectric detection measurements of the ablation plume also showed a clear threshold and a time scale for growth of the ablation products with a characteristic time scale on the order of 0.3 Its. Thermoelastic generation occurs as a consequence of local heating of the irradiated sample. Laser energy is absorbed and thermalized causing a local Review of Progress in Quantitative Nondestructive Evaluation. Vol. 15 Edited by D.O. Thompson and D.E. Chimenti, Plenum Press, New York, 1996 577

expansion, which in turn acts as a source of elastic waves. Thermoelastic generation is non-destructive and generally desirable for NDE applications. As the laser pulse energy is increased, eventually local temperatures reach levels at which material is vaporized and ablated from the surface. An elastic wave is generated by the surface stresses which accompany the ablation. Laser generation of elastic waves in metals has been the subject of considerable experimental and theoretical work. [2] The relatively sparse literature on laser generation in other materials includes a paper by Taylor et ai., [3] describing experimental work on generation by a CO2 laser in pure polymer, and a paper by McKie and Addison, [4] comparing the effectiveness of Nd:YAG and CO 2 lasers for generation in graphite/polymer composites. Taylor et al. demonstrated that ultrasound waveforms at intermediate laser power levels can be interpreted as a superposition of waves generated by thermoelastic and ablation processes, with an additional component due to plasma breakdown at higher powers. McKie and Addison showed that Nd:YAG lasers were more effective than CO2 lasers for generation of ultrasound in graphite/polymer composites; however, surfaces could be more easily damaged by the Nd:YAG lasers. APPARATUS AND EXPERIMENTAL PROCEDURES Ultrasonic waves were generated by high-energy laser pulses incident upon the sample surface, as shown in Fig. 1. The laser generated 5 ns pulses of 532 nm light at a rate of 20 Hz, with pulse energies up to a nominal maximum of 200 mj. The illuminated area on the sample had a diameter of approximately 5 mm. The energy could be varied by the laser flash-pump voltage setting. Unfortunately, a suitable power meter was not available until the later stages of the research reported here, and the relation between laser energy and flash-pump voltage setting drifted with time, so Nd:YAG cw laser VBS PBS4 Nd:YAG pulsed laser Dl PBSl CFP Feedback PBS2 PBS3 Sample Figure 1: Schematic diagram of the experimental system. Ultrasound is generated by 532-nm pulses of energies up to 200 mj generated at a rate of 20 Hz by the Nd:YAG pulsed laser. Ultrasound is generated on the right side of the sample and detected on the left. Light from the 532-nm, 200-m W, single-mode cw laser is directed towards the sample through a series of variable and polarizing beam splitters (VBS, PBS). Dopplershifted scattered light is collected and directed into the confocal Fabry-Perot (CFP) interferometer, which is stabilized by a feedback signal provided by detector D2. Light transmitted through the CFP to detector Dl provides the ultrasound signal. 578

that the energy scales reported here should be regarded as nominal, with no significance attached to apparent shifts of 20% or so. Normal surface vibrations on the opposite side of the sample were measured with a confocal Fabry-Perot (CFP) interferometer of the type described by Monchalin and Heon. [5] The implementation was similar to that described by Wagner [6] and Tittmann et al. [7] The output of such interferometers is proportional to the surface velocity. The interferometer was used in the transmission mode, which limited the frequency-range of sensitivity to a half-maximum band between 0.2 and 2.5 MHz. The detection system used a single-mode solid-state laser providing a 0.7-mm diameter beam of continuous 532 nm light at a power level of 200 mw. As shown in Fig. 1, the beam is first divided by a variable beam splitter with the majority of light directed towards the sample. A small fraction is directed back into the CFP to provide a feedback signal for a servomechanism which keeps the operating point near the midpoint of a CFP resonance. The Doppler-shifted light scattered from the sample is collected by a series of optical components which direct the light into the CFP, where it is demodulated to provide a photodiode output proportional to the velocity of the moving surface. The waveforms were recorded on a digital oscilloscope coupled to the lab computer. The oscilloscope was triggered by the generating laser pulses, so that time t = 0 on all waveforms corresponds to the beginning of the ultrasound generation. In a typical experiment, through-transmission waveforms were recorded as a function of source laser power. Eight waveforms were averaged at each laser power level. Because the samples studied were fairly absorptive, changes in the absorption owing to surface damage were not evident in a series of measurements. DEPENDENCE OF SIGNALS ON LASER POWER Figure 2 shows a series of waveforms generated with increasing source-laser power, for a 2.9-mm thick quasi-isotropic graphite/peek sample. The signals clearly gain complexity as the power level is increased, suggesting additional generation mechanisms. Analysis of the waveforms is complicated owing to the numerous modes possible in anisotropic materials. A pattern of echoes consistent with the sample thickness and typical propagation speeds is evident. An initial attempt to quantify the dependence of the signal strength on source power is shown in Fig. 3, which shows the peak-to-peak amplitude of the first longitudinal signal as a function of pulse energy for three independent runs. Because of changes in operating conditions and the energy-calibration of the laser, moderate vertical or horizontal displacements of the curves are not significant. A pattern of generally increasing signals with increasing energy is observed in each case, with a trend toward a plateau well above typical ablation thresholds for the material, followed by less consistent patterns, including ranges of decreasing signal strength. Figure 4 shows a similar plot for a 64-layer, 8.2 mm thick quasi-isotropic graphite/epoxy sample. In this experiment ablation was visually observed (through protective goggles) at energy levels near 100 mj. The region of uniformly increasing signal strength clearly extends well beyond the point at which ablation is observed, before a plateau is reached at higher energy levels. 579

101 mj Signal (arbitrary units) 98 mj 90.5 mj 85 mj 79.5 mj 71.5 mj 0 Figure 2: Ultrasonic waveforms transmitted through a quasi-isotropic graphite/peek sample as source laser power is increased. The traces are displaced vertically to avoid overlap. The scale is arbitrary. Further evidence of the onset of ablation is provided by photoelectric detection of the ablation plume. A fast photo detector was placed near the generation side of the graphite/peek sample, oriented at 90 to the surface to collect mainly light emitted by the hot ablation products. A 730 nm filter was used to limit the response to the 532 nm generating radiation. The lowest trace on the left side of Fig. 5 1.5 Amplitude 1.0 (arbitrary units) 0.5 Ablation visible 80 100 120 Laser pulse energy (mj) Figure 3: Dependence of ultrasonic amplitude on laser intensity for the quasi-isotropic graphite/peek sample. The open circles correspond to the data in Fig. 2. During that measurement series, ablation was observed through protective goggles at the point indicated. 580

0.6 Ablation visible Amplitude 0.4 (arbitrary units) 0.2 0 %~0--------~80~------~1~00~------~12~0------~ Laser pulse energy (mj) Figure 4: Dependence of ultrasonic amplitude on laser intensity for the quasi-isotropic graphite/epoxy sample. represents the residual sensitivity to stray scattered 532 nm light. It is also shown magnified to better display the response characteristics of the photodection circuit. It is clear that the signal rises rapidly after laser impact, with a decay characteristic of the electronics. Signals of a different shape arise from the ablation plume. The curves increase in magnitude with laser energy. Each curve reaches a maximum and decays 0.2 1... '''~... " 234 t (/1s) 5 140 Laser pulse energy (mj) Figure 5: Left: Photodetector output, in arbitrary units, showing optical activity associated with the growth and decay of the ablation plume. The signals were recorded with a fast photo detector directed at 90 to the normal of the generation surface. The lowest curve is the response of the system to stray 532 nm light scattered from sample. It is shown magnified as a dotted line. Right: The maximum photo detector signal as a function of laser pulse energy. The line is a linear fit. 581

at a rate presumably associated with the cooling and weakening of the plume. The maxima are reached at times which increase with laser energy over the range 0.2-0.7 MS. The right side of Fig. 5 shows the maximum photodetector signal as-.a function of laser pulse energy. The data appear to extrapolate back to a threshold consistent with other indications of an ablation threshold. Again, owing to shifts in calibration of the laser, the energy scale here should be regarded as nominal. In future work laser power levels will be measured more consistently in order to compare measurements of the plume with ultrasonic signal strength. SEPARATION OF THE WAVEFORMS Further analysis of the waveforms was carried out in an attempt to differentiate the thermoelastic and ablatic generation mechanisms. The first approach was point-by-point subtraction of a scaled thermoelastic waveform (measured at low energies when no ablation was visually observed) from higher-energy waveforms. The scaling factor was determined from the measured signal at 1.16 MS, approximately 0.1 MS into the first longitudinal pulse, assumed early enough to precede any waves generated by ablation mechanisms. A typical thermoelastic wave was multiplied by the scale factor and subtracted from the combined waveform. Figure 6 displays the thermoelastic wave, the combined waveform and the difference between the combined wave and the scaled multiple of the thermoelastic wave. This difference will be referred to as an ablatic wave. The ablatic waveform is similar to the thermoelastic waveform, with small differences possibly related to the different generation mechanisms. Its first positive peak typically occurs between 0.2 and 0.4 MS after the peak of the thermoelastic waveform. Kukreja and Hess observed time delays of this magnitude in studies of ablation generation in pure polymer films. [1] Using very different techniques, they observed dependences on source laser wavelength and power (Arbitrary units) o 2 6 t (MS) Figure 6: Isolated ablatic wave (dotted line), derived from the 112-mJ waveform by subtracting appropriatedly scaled values of a thermoelastic waveform measured with 73-mJ laser pulses. 582

I I I I 0.10 r- - 0.08 r- A - 0 0 (Arbitrary 0.06 r- - units) 0 0 0 0.04 r- - 0 0 B 0.02 r- - 0 0.00 r-0 6 0 0.0.0 00 - I I I I 80 100 120 laser energy (m] /pulse) Figure 7: Amplitudes A (.) and B (0) of the thermoelastic and ablatic basis functions, as functions of laser power. which have not yet been explored in the current work. The next step was to analyze the full waveforms using linear regression, in an attempt to separate out the amplitudes of the thermoelastic and ablatic components. The low-energy waveform was used as the "basis" thermoelastic waveform f(t). The waveform was arbitrarily scaled to unit amplitude at its first maximum near 1.2 /LS. The highest energy waveform was similarly scaled to unit amplitude at the first maximum and f(t) was subtracted from it to remove the thermoelastic component. The difference was then rescaled to unit amplitude at the first positive maximum attributed to the ablatic signal. This waveform was used as the ablatic basis waveform g(t). Waveforms generated at intermediate laser pulse energies were assumed to be a linear combination Af(t) + Bg(t). The coefficients A and B, determined by linear fits to the data, are shown as functions of laser power in Fig. 7. The fits were generally of acceptable quality; fits limited to the range 0 :S t :S 2 /LS had correlation coefficients in the range 0.933 to 0.997. Fits over the full recorded range had poorer correlation coefficients but the values of A and B so determined were not very different. A variable time delay T was introduced into the model by replacing g(t) with g(t + T), and determining T by non-linear least squares. The best fits were obtained with T = 0 ± 50 ns. No systematic changes in T with laser power were observed. Figure 7 shows a clear threshold for the ablatic wave amplitude. In future work this will be correlated quantitatively with the threshold of the ablation plume photo detection signal by making frequent laser power measurements. Also, improvements in the signal-to-noise ratio should make it possible to follow the thermoelastic signal to lower levels. DISCUSSION The application of laser ultrasonics to the study of the transition from thermoelastic to ablation generation of ultrasonic waves has been demonstrated. A 583

limiting condition on the maximum amplitude achieved by entering the ablation regime for CFR composites has been found and attributed to interference from a second (ablatic) wave. The ablatic waveforms were found to have a first positive peak on the order of 0.2-0.4 fj.s later than the thermoelastic wave, qualitatively consistent with the time scale observed for formation of the ablation plume. Linear fits to model thermoelastic and ablatic waveforms determined the dependence of the amplitudes of the thermoelastic and ablatic waves on laser power. Calculations of the first maxima of the waveform simulated from the model showed a trend to saturation similar to that directly observed. This suggests that the interference between the thermoelastic and ablatic waveforms limits the amplitude of the first peak. This point will be investigated more fully in future work. ACKNOWLEDGEMENT The authors gratefully acknowledge the support of this work through the Army Research Office/University Research Inititative grant, "Multidisciplinary Program in Manufacturing Science of Polymeric Composites." REFERENCES 1. L.M. Kukreja, and P. Hess, "Photoacoustic detection of the decomposition kinetics of polymers: interpretation of acoustic signals," Applied Surface Science, Vol. 79, 399-402 (1994). 2. See e.g. C.B. Scruby and L.E. Drain, Laser Ultrasonics, (Adam Hilger, New York, 1990), and references therein. 3. G.S. Taylor, D.A. Hutchins, C. Edwards, and S.B. Palmer, "TEA-C0 2 laser generation of ultrasound in non-metals," Ultrasonics, Vol. 28, 343-350 (1990). 4. A.D.W. McKie and R.C. Addison, Jr., "Practical considerations for the rapid inspection of composite materials using laser-based ultrasound," Ultrasonics, Vol. 32, 333-345 (1994). 5. J.-P. Monchalin and R. Heon, "Laser ultrasonic detection and optical detection with a confocal Fabry-Perot interferometer," Mat. Eval., Vol. 44, 1231-1237 (1986). 6. J.W. Wagner, "Optical detection of ultrasound," in Physical Acoustics XIX, A.D. Pierce, ed., 201-206 (Academic Press, New York, 1990). 7. B.R. Tittmann, R.S. Linebarger, and R.C. Addison Jr., "Laser-Based Ultrasonics on Gr/Epoxy Composite," Journal of Nondestructive Evaluation, Vol 9, 229-238 (1990). 584