You are on page 1of 22

NDT.net - June 1999, Vol. 4 No.

6 Ultrasonic Inspection 2 Training For Nondestructive Testing Variables Affecting Test Results Ed Ginzel Email: eginzel@mri.on.ca Homepage: http://www.mri.on.ca Home study correspondence courses (UT, ECT, LPI and MPI) including NDT Problem Solver Software (solving standard equations in UT, RT and ECT)

TABLE OF CONTENTS Introduction Ultrasonic tests can provide information about several aspects of a material such as: thickness, attenuation, shape, presence of defects, size and their orientation. These rely on two main measurements: amplitude of signal and time of the signal arrival. To a lesser extent the frequency content of the signal can also provide useful information but its application is not so common.

Introduction Instrument Performance Transducer Performance Material Variations


Surface Roughness Surface Coatings Couplant Condition Part Size and Geometry Internal Structure

Defect Variation

Defect Size and Geometry Location with Respect to Adjacent Surfaces

We make certain Orientation of Major Axis assumptions about the test conditions and Type of Discontinuity and Conditions of Reflection presume that changes in time or amplitude are caused by variation in the parameter of interest. The assumptions made are based on all parameters being constant except the one we are interested in measuring changes in. For example, when performing a thickness measurement we assume the acoustic velocity of the test piece we are measuring is the same as the acoustic velocity in the calibration piece. We further assume that the temperature at which tests and calibrations are made are not important. Yet either or both of the parameters assumed fixed (materials velocity and temperature) can affect our test results. Variables affecting the test results will be divided into 4 groups: 1. instrument performance 2. transducer performance 3. material variations 4. defect variations

Another factor relating to the results of an inspection is the Human Factor, this is a widely debated subject. The subject is not discussed in this chapter nor is the related subject of Probability of Detection. For more information on POD please refer to earlier publications on NDTnet Instrument Performance In Chapter 7, performance verification of instrumentation was discussed. Scope (display) and pulser/receiver variations can affect time, amplitude and frequency content of ultrasonic signals. Scope - The primary variable in the scope is the linearity of the time base. Verification methods will usually require a tolerance in accuracy to a percentage of the total screen range (typically +/- 2%). This ensures no distance measured will be in error by more than 2%, e.g. for a 250mm range it may be possible to have an error of +/-5mm maximum in steel. Pulser-Receiver - Amplitude uncertainties will result from variations in the linearity of the vertical deflection of the scope or due to inaccuracies in the amplitude control. Scope vertical linearity ensures that the relationship between two signals of different amplitudes is maintained over the entire range of the screen height. This is done by comparing the relative height of two echoes at different screen heights. e.g. setting two echoes 6 dB apart starting with one at 80% FSH, the other at 40% FSH adjustments are made to first increase the 80% FSH signal to 90% and 100%. The lower signal should be 45% and 50% respectively. Reducing the higher signal in 10% FSH increments, the lower should continue to be half its height. Tolerance for this parameter is +/-5% of the screen height. This ensures that the signal ratio of two different amplitudes truly indicates the size or distance effects. This would be most important for DGS type comparisons. The other aspect of vertical linearity variability is the amplitude gain control. This applies to the calibrated gain control usually found in dB increments on a flaw detector. Since the dB is derived from (dB = 20 log A2/A1 changing the dB gain by a fixed amount should change the ratio of the signals. This allows us to expect a signal at 50% FSH to increase to 100% FSH when 6dB is added to the receiver gain. ASME code requires scanning of a weld to be done using 14dB over reference. This means a signal that was 20% of the reference amplitude at reference gain would then come up to the reference level denoted by the DAC. If the receiver gain is not linear the smallest recordable indication may be greater or less than the intended level. This will be another source of incorrectly sizing a defect with respect to a reference. The effect of bandpass filters on displayed signal amplitude has been discussed in Chapter 6. The effect on amplitude is to reduce signal height if the centre frequency of the received signal is beyond (greater or less than) the bandpass region of maximum response. This can be a factor when it is known that the reflected or re-transmitted signal off a defect has a frequency spectrum determined by the characteristics of the defect (this is the basis of acoustic spectrum analysis or frequency analysis). Transducer Performance As with the pulser/receiver, transducer performance is checked and monitored for change. Codes and standards such as those discussed in Chapter 7 cover the details for verifying instrument, transducer and

system performance. But in addition to ensuring these aspects are within tolerances allowed initially, they must all be monitored on a regular basis to ensure no significant changes occur. BS 4331 Part 3*, recommends the following probe/system performance checks; ITEM MONITORING FREQUENCY daily on rough surfaces, such as castings, twice daily

(squint) monthly and when large changes in probe angle or index are observed monthly and whenever repairs have been made to either probe or instrument and if one instrument is replaced with another

frequency

-tonoise ratio daily and after repairs or replacement as above

gain

power

monthly and after repairs or replacement as above

* BS 4331 Part 3 will be replaced by EN 12668 -2, Non-destructive testing - Characterization and verification of ultrasonic examination equipment- Part 2: Probes. The above monitoring items apply to contact testing probes. The wear experienced by movement on metal surfaces tends to accelerate changes in performance. Some of the changes introduced by wear can alter test results significantly. As an example, consider beam angle change. If at the start of the day a nominal 60 probe was found to have an actual angle of 62, an indication is found with a soundpath of 150mm in a 100mm thick plate butt weld. A quick calculation made by the operator allows plotting to position the defect 132.4 mm along the surface from the exit point and then 70.4 mm below the test

surface. After several hours of heavy scanning the operator has unwittingly worn the probe down on the nose and the actual angle changes to 58 (both 62 and 58 are within acceptable tolerances for a nominal 60 probe). When the same indication is investigated again the operator finds an indication in that same area but it does not seem to be the same. The soundpath is only 133mm for the "new" indication. The operator, believing the refracted angle is 62 now plots a defect 62mm below the surface. This situation is illustrated in Figure 8-1.

Figure 8-1 Similar errors in lateral positional plotting can result from skewing of the beam. When plotting an indication with an angle beam we assume the beam extends directly ahead in-line with the probe housing but wear on one side or the other of the wedge will steer the beam away from the centre-line. If we use the previous example with the 62 probe finding an indication at 150mm soundpath, the plan view plot would indicate it to be about 132mm directly in front of the probe from the exit point. If a 5 skew existed that was not accounted for, the error in locating this indication would be about 11.5mm (see Figure 8-2).

Figure 8-2

Index point or beam exit point for an angle beam probe is easily established on the 11W block. This is used to establish the actual refracted angle so its accuracy within +/-1.5mm is essential. For longer soundpaths (>25mm the effect on positioning a flaw forward or backward would not be too critical. e.g. If the flaw plotted in Figure 8-1 was 1mm forward or backward due to the exit point being off 1mm from its scribed position it would have little bearing on the evaluation of the defect. If, however, the weld tested was on a 6mm thick pipe with a TIG root the root width might be about 2mm. An error in exit point placement could plot the defect on the wrong side of the weld. In the chart of items checked as per BS 4331, the first three items are unique to contact probes, but the remaining items could be considered by any transducer evaluation, including immersion probes. Handling and aging can cause changes to the element's backing, degree of polarization, lensing material shape, lens material bond to the element or degree of loading (for the thin gold face on PVDF elements).

These changes result in changes in both amplitude and frequency. The effect on performance is multistepped. For example: if aging has resulted in a slight disbonding of the element from the high density backing of a standard ceramic element, its damping will be reduced. This will lead to an increased ringing. More ringing reduces resolution and increases the extent of dead zone due to the rattle. Decreased damping due to the disbondment, however, allows vibrations to be larger so sensitivity is increased. The reduction in backing load tends to change the centre-frequency to a higher value but the increased sensitivity, afforded by more and larger vibration displacements, reduces the bandwidth. The higher frequency increases the near zone as it is a function of wavelength. The angle of divergence is also changed (decreased ) as it too is a function of wavelength. Operating probes in warm water (>50C) or high radiation fields (several MegaRads) can cause blistering or disbonding of the epoxy material used for lensing. This could have similar effects to those noted for backing disbondment as well as distorting and redirecting the beam centre-line. In addition to aging and environmental causes of alterations to the transducer performance, handling can also cause changes to occur. A sharp jolt from dropping a probe may result in similar disbond problems. With the availability of different pulse shapes it may be possible to deteriorate polarization in an element. A negative going pulse voltage is normally applied to probes but polymer elements tend to perform better with a positive going pulse. Polymer probes will show no deterioration if pulsed with negative going spikes but ceramic elements may experience depolarization over extended periods of time. Depolarization will reduce sensitivity and the increased gain required will manifest itself in a lower s/n ratio. Sources of variation in transducer performance are many. Establishing a baseline with tolerances and then monitoring for changes in any of the parameters checked will help to ensure reliability of test results. Material Variations When considering the variables of the test material that affect test results we can group them into three areas of concern: 1. entry surface 2. part size and geometry 3. internal structure. Entry surface variables include: 4. surface roughness 5. surface coatings

6. couplant condition. Surface Roughness Surface roughness will have several possible effects on the inspection of a test piece. In contact testing roughness on a gross scale results from: weld spatter, plate scale, dirt (sand) and rough cast surfaces from sand casting. These irregularities will cause some points of contact to push away the couplant and force it into the lower areas around the probe. If the couplant is not sufficiently viscous it will drain away quickly and fail to couple the probe to the test piece. See Figure 8-3.

Figure 8-3 In addition to reduced coupling, which will reduce signal amplitudes, the rough surface increases the rate of wear on the probe. On an otherwise smooth surface isolated protrusions such as weld spatter can hinder or stop probe motion or in the case of mechanized systems there may be sufficient force to move the probe past the obstruction but this could result in damaging the probe by either tearing it from its mounting or severely scoring the plastic wedge. When the dirt on the test piece is very fine (similar to a flour texture) coupling can be prevented due to surface tension preventing the liquid couplant penetrating to the metal. Unless a transfer value has been established between test piece and calibration piece, this could go undetected. In addition to affecting coupling, surface roughness tends to reduce signal amplitude by scattering and focusing the beam. This applies to both contact and immersion testing. Whether uniform or irregular, a rough surface has the potential to present a scattering effect at an interface where a beam impinges. The degree of scattering is based on the ratio of roughness to wavelength. When roughness is less than about 1/10 a wavelength, scatter will be negligible. To reduce signal losses due to scattering an operator can select a lower frequency probe. With a wavelength of 0.37mm in water for a 4MHz probe, signal loss due to scatter can occur for irregularities as small as about 0.04mm. In addition to signal reduction another effect of surface irregularities is to redirect and mode convert some energy which when returned to the probe can be the source of spurious signals. In contact testing false indications from standing waves resulting from scatter on rough surfaces will normally have short soundpaths. They can be eliminated as true flaws by failing to locate any trace of indication from the full skip or from the opposite side.

Unless done properly, removal of surface roughness by mechanical means can result in further scattering problems. Small curved gouges left by a grinding wheel used to remove spatter or machining grooves can form small lenses. The affect of grinding can be unpredictable. Some of the lensing may concentrate the beam thereby increasing signal amplitude, or, the lens effect may be a de-focusing of the beam, again resulting in lower than expected signal amplitudes. Uniform surface preparation by sand or shot blasting usually provides a good surface for ultrasonic testing. Removal of excess metal by a hand held grinding wheel is commonly used on weld caps and roots. When a pipe weld has had its root ground flush and inspection can only be performed from the outside diameter, quality of grinding can result in unnecessary repair calls if grinding has been along the weld axis. The small grooves made by the grinding wheel run parallel to the root edge and are easily confused with lack of fusion, missed edge or undercut defects. Surface Coatings Surface coatings are added to protect a surface from corrosion or to enhance its appearance. Thin films, such as oxide layers, anodizing layers or electroplated finishes, and the slightly thicker coatings of paint or lacquer are usually well bonded to the surface. Quality of bond may be compared to the uncoated reference block by a simple transfer value. Even a slight loss due to the coating may be preferable to removing the coating and trying to inspect on the rough surface it hides. When thickness testing is done on a painted surface the paint thickness can add error to the reading. For example: A nominal 25mm steel plate has a cellulose paint coating of 0.5mm. Vsteel = 5980m/s, V paint = 2600m/s. If a digital thickness meter is calibrated on a 25mm thick piece of the steel plate without the paint coating and then placed on the painted surface an error will occur. The coating is sufficiently thin that its interface with the metal will occur in the dead zone but the duration of time spent in the paint is added to the travel time to the opposite wall of the plate. If the true plate thickness at the point of measurement is 25.16mm and the paint coating is 0.5mm thick, the time in the paint is 0.5 ( 2.6 x 10 6 = 0.19s. 0.19 microseconds is equivalent to 1.15mm in steel. The reading on the digital meter would combine the two thickness as though all travel was in steel. This results in 25.16 + 1.15 = 26.31mm as the indicated thickness. This problem can be overcome by using an A-scan display and measuring the interval between the first and second echo instead of the main bang and first echo. This is shown in Figure 8-4.

Figure 8-4 Couplant Condition Both contact and immersion methods utilize intervening media to transfer sound from the probe into the test piece and back to the receiver. With immersion methods it is accomplished by a single fluid medium. In contact testing there are nearly always at least two intervening media; the delayline or protective face and the thin film of coupling fluid or grease. Attenuation and acoustic velocity are the two main properties that dictate the performance of a couplant. Attenuation affects amplitude of the signal and velocity will determine both transit time and refracted angles. But attenuation and velocity of couplants are not independent properties. Each is a function of other parameters. Unless these parameters are controlled or in some way compensated for, gross variations from the reference value or calibration conditions can result. Attenuation of couplants varies with material composition as would be expected. Published attenuation values are available for many materials as indicated in the table below. Attenuation coefficients are often quoted in nepers which allow for frequency dependence. 1 Np = 8.686 x f2 = dB/cm. Table 8-1 indicates attenuation of some common liquids. Table 8-1 Material water Attenuation (Np x 10-15) 25.3

silicon oil castor oil mercury ethylene glycol methanol

6200 10100 5.8 128 30.2

In more practical terms, for water, this would mean an attenuation of about 5 dB per metre. Since such long water path lengths are not normally used the 0.005 dB/mm attenuation is considered negligible. But for the heavier oils attenuations 200 to 500 times greater can have significant effects on signal amplitude and frequency content. For the fixed delaylines or wedge materials used in contact testing attenuation variations can be far more pronounced and variation between manufacturers can cause considerable response differences. For example the plastics listed in table 8-2 are typical wedge materials selected by manufacturers and based on velocity for refraction purposes, but attenuation differences would result in noticeable amplitude response variation and frequency content of transmitted waveforms. Since the operator rarely knows what wedge material a manufacturer has used, little can be done to correct for potential variations in periodic inspections where results of tests taken with one or more years separation are compared. Table 8-2 Material Attenuation dB/cm @ Acoustic 5 MHz Velocity 2.75 to 2.61 2.30 2.32 to 2.48 2.6 to 2.77

Plexiglas (acrylic) 6.4 to 12.4 lexan (poly carbonate) polystyrene nylon 32.2 1.8 to 3.6 2.8 to 16

Attenuation is not a material constant. Under changes in conditions it can change. For example attenuation in water is inversely proportional to both temperature and pressure. At standard pressure and temperature (1 atmosphere and 20C) attenuation in water is 25.3 x 10-15 Np. When temperature is 0C and water still liquid attenuation is 56.9 x 10-15 Np and at 40 it is 14.6 x 10-15 Np. At 1000 atmospheres attenuation drops to 12.7 x 10-15 Np and increases to 18.5 x 10-15 Np in a vacuum (zero atmospheres) when the temperature is held at 30C.

Attenuation of couplants need rarely be considered when calibration and test conditions are the same couplant material, temperature and pressure. However, mechanical actions can add to variations in attenuation under some conditions e.g. liquid soap is often used in contact testing. Under static conditions it provides reasonable coupling, ease of probe movement and clean hands. When a part is inspected with more rapid probe motion than may be used for static calibration it is possible to lather the soap. As bubble density builds in the couplant attenuation will increase. Far more pronounced on test results are the affects of velocity changes. As with attenuation of couplants, velocity is normally considered fixed for a given material. Providing all parameters affecting velocity are controlled the assumption is valid but subtle changes in parameters can have significant results. Just as plastic compositions change in velocity so too does water. Pure water at 20C and 1 atmosphere pressure has a velocity of 1480m/s. But water is not normally pure. As salinity increases as in sea water, acoustic velocity increases. At 20C in sea water with a 3% salinity the acoustic velocity increases to about 1515m/s. For work done on off-shore structures, where immersion work would occur using the surrounding sea water, any calibrations done on-board ship must use water of the same salinity as will occur at the depth of test or probe placement and refracted angle will be mis-calculated. As well, acoustic velocity increases with pressure in water so with increasing water depth velocity also rises. This is relatively insignificant but may be corrected for by the equation. Vd = Vo + (d x 0.018) where Vd = acoustic velocity at depth d d = depth in meters Vo = acoustic velocity at the surface. For our example of a 3% salinity the correction for work at 50m would be 1515 + (50 x 0.018) or 1515.9 m/s. Temperature is undoubtedly the most significant parameter affecting acoustic velocity. As such, its control or knowledge of its change is essential to ensure inspection accuracy. Change in acoustic velocity with change in material temperature is termed acoustic temperature dependence. Strangely, temperature dependence is not always the same sign, i.e. increasing material temperature will increase acoustic velocity in some materials and decrease it in others. Table 8-3 illustrates some of the variations in temperature dependence ( V/ t) for some materials. Table 8-3 Material caster oil Acoustic Velocity @25C 1470 (V / t (m/s C) -3.6

ethylene glycol kerosene methanol water (pure) water (sea) polystyrene polymethyl methacrylate polyvinyl chloride (hard)

1658 1324 1103 1498 1531 2400 2690 2380

-2.1 -3.6 -3.2 +2.4 +2.4 -4.4 -2.0 -8.0

By comparison most metals have a temperature dependence of between -0.5 to -5.0 m/s/C depending on mode and axis of propagation. When dimension verification is accomplished by determining Table 8 - 3a: travel time in a coupling liquid, as in Figure 5-20, variation in Velocity change with Temperature temperature could easily affect accuracy. Even more Temperature Velocity Delta challenging is correcting for change over a period of time. C m/s m/sC e.g. temperature at the start of the test is 20C, but over a half hour period it increased to 35C. For water this increases 5 1440 3,14 the acoustic velocity to 1520 m/s, it does not change linearly with temperature (see Table 8 - 3a). 15 472 2,65 When applied to angle beam testing the problems become more obvious. Snell's Law allows us to predict refracted 25 1498 2,15 angles in a material based on knowing the acoustic velocities 35 1520 1,67 involved and the incident angle. The ratio of the velocities determines the degree of refraction. When a standard 45 1536 1,18 contact testing wedge is purchased it is marked with the nominal refracted angle it will produce in a steel with shear 55 1548 0,68 velocity assumed to be 3250 m/s. A plastic wedge having an acoustic velocity of 2600 m/s is machined at 43.9 to 65 1555 0,20 produce a nominal 70 refracted beam in steel. This assumes 75 1557 20C operating temperature. If a test piece is inspected that is warm to the touch (~40C) the velocity of both steel and plastic will be less than assumed. If the plastic changes by -4.0 m/s /C, and steel by about -2 m/s C the velocities to use would be 2312 for the plastic and 3210 for the steel. The actual refracted angle would be closer to 74.

Figure 8-5 illustrates the effect of temperature on refracted angles in steel for three common fixed wedge angles.

Figure 8-5 Incident angles indicated in the legend in Figure 8-5 are those to produce nominal 45, 60 and 70 refracted angles under standard conditions. The plastic calculated for is UVA II with an acoustic velocity of 2760 m/s. The affect on angulated longitudinal waves is even more pronounced, this due to the greater ratio difference in velocities between the plastic and steel in compressional mode. Part Size and Geometry Test results may vary if the test piece differs from the calibration or reference piece. In this way both shape and size will contribute to potential variation in test results. Particular interest in this variable exists for contact testing on curved surfaces. When a flat probe is used on a convex curved surface only a portion of the probe makes contact. This will reduce the amount of sound that can be transferred to and from the test piece. As a result sensitivity compared to coupling to a flat piece is reduced. The proportion of sound reduction compared to a flat piece is a function of the curvature of the part, the crystal diameter and the coupling ability of the couplant via its viscosity. Some sources also consider the relative hardness of the probe face with a greater coupling or contouring available from softer material such as plastics and virtually no contouring available. To avoid machining calibration blocks for every possible radius and surface condition compensation is made by adding gain to the receiver. The amount of compensating gain can be determined by a simple

transfer value or it can be calculated using formulae and charts. Examples of the charts used for convex curvature compensation are found in Figures 8-6 and 8-7. These are from Australian Standard 2207 Methods For The Ultrasonic Testing of Fusion Welded Joints in Steel. Two conditions are considered. In the first figure a nomogram is used to correct for losses when the probe contact is made on the curved surface. The test part radius is located on the left-hand scale and a line made through the appropriate probes diameter on the middle scale. The point on the right-hand scale this line intersects is the amount of gain to add as a correction factor. In the second figure the probe makes contact on a flat surface but the beam reflects off a convex curve thereby redirecting portions of the beam away and reducing the maximum possible reflected energy. The graph used does not consider probe diameter, instead, ratio of surface curvature to metal path thickness is used. moving vertically up from the ratio axis (horizontal axis) at the appropriate ratio for your work piece, the point on the vertical axis where the curve is intersected gives the necessary correction factor.

Figure 8-7

Figure 8-6 Other codes and specifications may use different equations or graphs but the intent is the same. Attempts to compensate by simply adding gain may not be adequate. Improvement is had by contouring the plastic wedge to the test piece. However, since the probe can no longer be calibrated on a flat reference piece this makes machining of a reference piece of the exact same geometry a necessity. For very small parts even this may prove unsatisfactory due to the production of surface waves and other spurious signals associated with large time differences of the beam in the wedge or delayline. If contouring probes proves too noisy then immersion methods or even another NDT method may have to be considered. Geometry is not only a consideration as a potential source of signal variation but also of feasibility. Consideration must be given to beam shape when interaction with a boundary occurs. Formation of

unwanted surface waves and mode converted waves will result due to finite extents of a beam. The single ray drawn from the probe exit point at the refracted angle is merely a convenient presentation of the principle of the test. In reality portions of the beam will impinge at greater and lesser angles due to divergence or side lobes (in the near zone). A beam impinging on a curved surface intended to generate a high refracted angle in the test piece is most prone to surface wave generation. See Figure 8-8. This occurs more often for curved surfaces than Figure 8-8 for flat surfaces because not only does the beam divergence increase the incident angle but the point of incidence on a curved surface is always receding so further increase in incident angle results. Curved surfaces make plotting more difficult than the simple trigonometry for flat surfaces. Compare similar conditions for a 45 beam in flat and curved plate. Using a thickness of 20mm the signal obtained at 35mm for an inspection of a 100mm radius pipe occurs 19.2 mm from the OD test surface and 31.7mm along the test surface from the exit point to the point over the indication. On a flat piece this would indicate an indication 15.2mm down and 24.8mm from the exit point. See Figure 8-9.

Figure 8-9 Geometry can also limit inspections. Again, concentric geometries are a common problem in this regard. A critical ID/OD ratio exists that will not allow an angle beam from the OD to intersect with the ID. This is shown in Figure 8-10.

Figure 8-10 This critical angle occurs when the ray representing the centre of beam is tangential to the inside surface of the pipe. Two situations may exist; either the angle is fixed (as in contact testing) and the critical ratio is to be determined or the ratio of wall thickness to the OD is known and the maximum angle that can be used is sought. The equation relating wall thickness t, pipe outside diameter D and refracted angle can be written D(1-Sin ) t=------------------2 t is the maximum thickness for which the beam centre will still glance off the ID. The other parameter being sought, maximum angle, would use the following equation: 2t Sin = 1 ----D Figure 8-11 plots the t/D ratio that satisfies the above equation. When the thickness increases to one half the diameter this is the same as a solid cylinder and no ID exists to bounce off.

Figure 8-11 Although the equations and graph indicate that high angles of refraction may be used to ensure both ID and OD are seen, the actual size of the pipe will limit the practical maximum angle due to beam width and the resultant surface waves that occur. For tubular products where wall thickness approaches the wavelength the wall is flooded with sound and plate waves discussed in Chapter 5 result. Some authors have indicated that contact pipe inspection in the circumferential direction can be accomplished by simply moving the probe over the pipe surface by a distance equal to one full skip. The principle involved is shown in Figure 8-12.

Figure 8-12 This draws well for the centre of beam ray but when applied to real conditions this can supply little more than a go/no-go inspection. Beam spread, mode conversions and attenuation will not permit accurate locating of any defects occurring several skips away. In fact attenuation will probably limit detectability

of defects by this method to pipes under 10-20cm diameter. Such a technique may be useful when access is limited to only half of the circumference. The presence (but not necessarily the location) of a flaw may be detected by first scanning in one direction to the obstruction then the other, as in Figure 813.

Figure 8-13 Internal Structure The final aspect of material variations affecting test results is the structure of material under test. Material parameters are a function of makeup and environmental conditions. Makeup is determined by design and processing. Whether the material under test is steel, aluminium or fibre-composite, variations can occur by design. Proportion of resin to fibre will vary in composites and metals may have many alloying variations. In addition, metal grain structure can be varied by alloy, heat treatment and working. All these factors will provide differences in the results of ultrasonic tests manifested as variations in velocity or attenuation. Also, just as temperature and pressures were noted to change velocity and attenuation in couplants so too will the material under test be similarly affected by these externally controlled conditions. In cooling molten metals, solidification begins at many sites throughout the melt. At each site the growth pattern is determined by the surrounding liquid material and the surrounding temperature gradients. Crystals form as growth progresses. Eventually the crystals' growth is halted when another crystal is encountered and all the liquid has been consumed. Although we generally consider the metal to be homogeneous, on a microscopic scale the boundaries formed by the edges of the grains make it inhomogenous. For an ultrasonic wave each crystal presents a different acoustic impedance depending on orientation and degree of inter grain bonding. As well, there may be pores, gaps and non-metallic inclusions. All these factors will cause scattering. Just as with surface roughness, scatter will be a function of wavelength. Krautkramer points out that for grain sizes up to about 1/100th of a wavelength scatter can be considered negligible. However, as grain size increases beyond that, it can become a significant factor adding to decreasing signal amplitudes. As grain sizes increase to greater than 1/10th the wavelength, inspection may not be possible by

ultrasonics. Austenitic stainless steels are typical of metals with large grain structures. In the production of austenitic steels manufacturers often attempt to control or limit grain size. This is done by : a) introducing small amounts of grain refining elements b) limiting the temperature the steel is heated to c) by hot working the steel to break up the austenite grains. In spite of these effects the stainless steel product is not always consistent in its grain structure. When testing stainless steel forgings, it is possible to have areas of higher attenuation than others. In cases such as this it will require observant operators to recognize the increase in grass level. Velocity changes with material and condition as well. Contact angle wedges are normally made for steel so the refracted angle indicated on the wedge assumes it will be used on a steel with a longitudinal velocity of about 5900m/s and a shear wave velocity of about 3250m/s. When the same wedge is used on aluminium plate with longitudinal velocity 6320 m/s and shear at 3100 m/s the 'nominal' refracted angle indicated will not indicate the true angle. But one need not move to a completely different metal to illustrate velocity changes. Rolled plate for pipeline construction shows variations of 8-10% in shear wave velocities. This is attributed to rolling and heat treat differences and the resulting differences in grain elongation and orientation. Even alloys of steel show marked variation in acoustic velocity; 4340 steel has a shear velocity of 3240 m/s while in 4150 steel it is 2770 m/s, a difference of nearly 17%. These variations point out the importance of material specific calibration pieces. Finally, as with couplants, acoustic velocity of a test material varies with temperature. Most published values will indicate velocities determined at 20C. For work at much higher or lower temperatures corrections will need to be made. This will require the temperature dependence for the material to be established and this will have to be in addition to similar corrections made for couplant changes. Defect Variation The fourth major factor affecting test results is the defect or reflecting surface of interest. In evaluating a signal an operator will use three items; soundpath, probe position and amplitude. The change in relationship of these three aspects is called "echo dynamics". Therefore, investigating the echo dynamics of a flaw allows the operator to build up an image (mental in manual scanning and possibly visual if automated) of the shape of the flaw. Four factors are significant in the response obtained from a defect; 1. size and geometry 2. location with respect to adjacent surfaces 3. orientation of the major axis 4. type of discontinuity and conditions of reflection. Defect Size and Geometry Both defect size and defect shape have a significant affect on signal amplitude. Principles of the AVG system showed how signal amplitude is a function of the ratio of reflector area to element area. Generally small defects provide smaller amplitude signals than larger flaws. However, an irregular flaw

shape may mean not all of the flaw reflects the sound back to the receiver. Irregular facets of a crack or close proximity of pores in clusters of porosity can result in sufficient losses due to scatter that very small signals are received in spite of the fact that a large volume of metal is missing; i.e. signal amplitude is no guarantee of defect size. Defect sizing using the AVG method can result in errors of 3:1 to 6:1. With the SAFT method sizing accuracy can be almost 100% (published by IZFP). Location with Respect to Adjacent Surfaces Defect position with respect to adjacent surfaces presents several causes of variable results. Simple attenuation accounts for reduced signal amplitude by increasing the soundpath (in the far zone) to the flaw. If the flaw is close to another reflecting surface confusing signals may result or signals may be lost. Figure 8-14 shows how a mode converted signal may arise due to a reflection off a planar flaw near a flat surface.

Mode converted wave off defect at 65 strikes radius providing large amplitude signal that when plotted gives a virtual location of a defect different from its actual position Figure 8-14 If the adjacent surface in 8-14 had been flat the reflected shear and mode converted signals would both have been undetected. Such signals produced by mode conversion can be differentiated from flaws by using a couplant wetted finger to rub the test surface. The compressional mode is noticeably damped. Mode converted longitudinal and re-directed shear waves can be seen when drawing distance amplitude correction curves on some calibration blocks. The typical ASME calibration block uses a side drilled hole at quarter thickness (1/4t). The convex radius of the hole allows both situations to occur (see Figure 815).

Figure 8-15 Orientation of Major Axis When the major axis of a defect is not exactly perpendicular to the beam reflection causes the returned signal to be directed away from the simple return path back to the transmitter. For small angles this will not cause a total loss of signal as beam dimensions are sufficient that the off-centre portions can still be detected by the probe. Even small angles off normal(e.g. +/-5) can result in significant signal reductions. When expected flaws are planar and no convenient pulse-echo angle can be arranged to ensure the beam will strike the flaw at right-angles tandem probe arrangements are preferred. Type of Discontinuity and Conditions of Reflection To some extent this has been addressed by the other aspects. Defect size and geometry is usually determined by its type; e.g. porosity is usually small and spherical, slag is irregular in shape and size, and non-fusion is usually planar. However, reflectivity of defects is not a simple matter of incident angle. For very fine porosity there may be no noticeable back reflected signal but the scatter such a dispersive defect would cause would reduce the transmitted energy. Maximum reflection occurs off a free boundary. This is effectively the situation for non-fusion and cracks where the void is air. However, when a dissimilar material fills the void, as would be the case in a slag inclusion or tungsten inclusions in a TIG weld or carbide inclusions in castings or forgings, part of the sound incident on the boundary is transmitted. This will reduce the reflected signal. Added to the loss due to transmission into the next medium is the associated losses due to the reflection at any angle other than 0. Reflection and transmission coefficients discussed earlier show how quickly amplitudes can change due to boundary material differences and incident angles. Finally, ambient conditions of pressure can have profound effects on a signal amplitude. Studies have shown a signal from a surface breaking crack to be reduced by 20dB when the cracked sample was placed in compression. This effectively "closed the gap". |Top|

NDT.net NDT.net, info@ndt.net /DB:Article /DT:tutor /AU:Ginzel_E_A /IN:MRI /CN:CA /CT:UT /ED:1999-06

You might also like