Вы находитесь на странице: 1из 20

J. Sound vz%.

(1967)

5 (I), 173-192

THE MEASUREMENT IN METALS

OF APPLIED

AND RESIDUAL WAVES

STRESSES

USING ULTRASONIC
D. I. CRECRAFT~

School of Engineering Science, The University (Received 25 February

of Warwick,
I

Coventry,

England

966)

The results of measurements of the stress-induced velocity variations of both longitudinal and shear ultrasonic waves in the megacycle region are quoted for steel, aluminium and copper, and the corresponding third-order elastic constants calculated. The methods of measuring the small velocity changes due to both stress and preferential grain alignment are discussed, and some attempts at separating the two effects so that residual stress can be non-destructively detected are described.

I. INTRODUCTION Residual stresses are becoming of ever greater importance to the engineer subjecting metallic materials to loads approaching their yield stresses, since, under these conditions, an unfavourable residual stress can add to the applied stress and cause failure of a critical component in service. Whilst the material is, in general, taken through many stressrelieving heat processes before attaining the final shape of the component, the final heat process will seldom involve sufficiently high temperatures for complete stress relief, and the final machining process will almost certainly introduce further residual stresses. In this context the critical stages in the manufacture are, therefore, the final stages after which no further work can be done on the component to relieve residual stresses. There is thus a need for a non-destructive means of detecting residual stresses in finished components. The only established technique depends on the use of X-rays to measure the Bragg reflection from, and hence the strain in, the metallic lattice of the first few atomic layers from the surface. The residual stress is then calculated from the residual strain, using the accepted values of elastic constants. As a method of measuring residual stresses within metallic materials, rather than on their surfaces, the use of ultrasonic waves has been suggested. It is known that ultrasonic wave velocities in materials are sensitive to stress within the materials, and hence the technique would be analogous to photoelasticity, which is used in the analysis of stress distribution in transparent materials. Because of this analogy the words sonoelastic (Elion (I)) and acoustoelastic (Benson and Raelson (2)) have been used to describe the effect, and sonoelasticity is used here to describe the use of ultrasonic shear (transverse) stress waves in close analogy with the polarized transverse light waves of photoelasticity. 2. PHOTOELASTICITY AND SONOELASTICITY

In photoelasticity plane polarized light is obtained from a polarizer and passed through a parallel faced t ransparent solid at normal incidence. Figure I shows how the wave splits into two components, each polarized parallel to the axes of the principal stresses in the t Dr. Crecraft was at the Department while most of this work was done.
of Electronic & Electrical Engineering, University of Birmingham

173

I74

D. I. CRECRAFT

plane normal to the propagation direction. In general, the two components travel with different velocities, since the effect of stress (except in the special case when the principal stresses are identical) is to render the material electrically anisotropic: i.e., the permittivity varies with direction, giving rise to principal permittivity axes coincident with the principal stress axes, and the material is said to be optically birefringent. The emerging light is viewed through a second polarizer, often crossed with the first, and Figure I shows that this passes components of the two waves which are equal and opposite in space, but which have a relative time-phase shift due to the difference in their velocities in the stressed material. This interference effect causes the light intensity at a given point

Axisof transmitter poloriser

I ,

Principal stress axis (Xl

emerging from moteriol

Figure I. Section through XY plane of material under stress, showing the space intensities of light waves in photoelasticity and ultrasonic shear waves in sonoelasticity, with crossed polarizers.

to reach a maximum whenever the stress is sufficient to cause 180~ phase shift, and to reduce to zero when an integral number of cycles of phase shift have been encountered. Points over the surface of the second polarizer at the same intensity appear to be joined by bands or fringes as they are commonly called, and these represent isobars in the stressed material. Since at any point the relative retardation, as it is called, between the two wave components is the result of all the stresses encountered in passing through the solid, and since the velocity changes are very small, the velocity difference indicates the average stress over the thickness of the solid at this point. Moreover, the two components passed by the second polarizer are maximum when the principal stress axes lie at 45 to the polarizer axes, and zero when either stress axis coincides with either polarizer axis; thus the directions of the stress axes can be found by rotation of the solid or of the polarizers. The relationship between velocity difference and stress is linear, and Table I (Heywood (3)) sh ows values of two coefficients used in photoelastic studies for three transparent materials representing the wide range encountered. The stress-optical coefficient, S, is defined as the fractional velocity difference per unit stress: i.e.,

where c, and c2 are the two velocities under stressed conditions, cO is the velocity in the unstressed material, and P and Q are the principal stresses in the plane normal to propagation. The fringe-stress coefficient, F, is defined as the stress required to produce one

STRESSMEASUREMENT BY ULTRASONICS

17.5

fringe: i.e., one cycle of phase difference between the two waves, in unit thickness of material. Thus
F =

P-Q)x
N

where z is the thickness of the material, and N is the number of cycles of phase difference. It is readily shown that F=h/S, and must therefore be quoted at the wavelength A. The elastic stress in a material also affects the velocity of an elastic wave. In the case of a solid, both compressional (longitudinal) and shear (transverse) waves will propagate, and the analogy with photoelasticity is in the use of shear waves. The velocity is given by c = (C/p), where p is the density and C is the appropriate elastic constant, which for shear waves is the shear modulus I*. Whilst the change in density due to strain has a small effect, the velocity change caused by the effective elastic modulus changing with
TABLE I

Photoelastic

and sonoelastic coeficients


Material S(i~~.~lb-l) F (lb in.-2)

Photoelastic values at h= 5893 A (Heywood)

Glass Celluloid Gelatin Nickel-steel Copper Aluminium

1.87 x 10-s 9.45 x 1 0-8 1.66 x I O-~ 594 x 10-s 9.0 x 10-s 2.41 x 10-7

1.24 x 1 0 3 2.45 x 102 1.4 x 1 0-l 475 x 105 2.15 x 10~ I.08 x 105

Sonoelastic values atf= 4.7 MC/S (Crecraft)

in the stress-strain relationship, within the so-called elastic region. This will be discussed later. The complete analogy with photoelasticity occurs when a shear wave is propagated through a stressed solid. The latter becomes elastically anisotropic, with the axes of principal elastic constants coincident with the principal stress axes, and the wave splits into two components polarized parallel to the stress axes, so that Figure I applies to this case also. The material is then said to be sonically birefringent. Continuing the analogy, the stress-sonic coefficient and fringe-stress coefficient may be defined as in the case of photoelasticity, and some values determined by the author are shown in Table I for comparison. It is interesting to note that the stress-optical coefficient for Celluloid has practically the same value as the stress-sonic coefficient for copper, but that Celluloid is over one thousand times as sensitive to stress in terms of the fringe-stress coefficient. The difference is attributable to the difference in wavelengths, since F = X/S. The photoelastic values in Table I are quoted for a wavelength of $393 A, or about 6 x IO- m, whereas the ultrasonic waves at 4.7 MC/S have X = 6 x IO-~ m in copper. Thus in photoelasticity the low S values are compensated by small wavelengths, and so large numbers of fringes can be obtained for small stresses. In sonoelasticity, on the other hand, the larger wavelengths give low phase shifts, which means that small phase angles must be measured accurately if stresses are to be detected at all in metals. The use of higher ultrasonic frequencies with smaller wavelengths leads to excessive attenuation. Furthermore, while the resolution across the field of view in the photoelastic case is limited only by the very small wavelengths, in the sonoelastic case

stress can be much larger, which implies non-linearity

176

D. I. CRECRAFT

considerations of beam shaping and attenuation at present lead to the use of sources and receivers several wavelengths across, which in practice means a resolution of several millimetres.
3. METHODS OF MEASUREMENT OF STRESS-INDUCED

VELOCITY

CHANGES

The direct measurement of the phase shift between the two shear wave components propagating in a metal under applied stress will yield the values of S and F and these can then be used in determination of residual stresses. However, it is not possible to tell from this measurement how much the individual velocities have changed. Consequently, methods of measuring small changes in the velocities of both shear and longitudinal waves were considered so that the changes in the elastic constants under stress could be estimated. Earlier work on the non-linear stress-strain relationship was carried out by Bridgman (4) and others, using very high hydrostatic pressures. Hughes and Kelly (5) propagated both shear and longitudinal waves through solids stressed both hydrostatically and uniaxially, and showed that a set of three different wave velocity change measurements was sufficient to calculate the three third-order elastic constants 1, m and n introduced by Murnaghan (6) in his second-order elastic theory. For velocity change measurements in Polystyrene, Armco Iron and Pyrex Glass, Hughes and Kelly used the pulse-echo technique entailing measurement by oscilloscope time-base of the time taken for a pulse of ultrasonic waves to travel through the specimen. This method lacks precision, and a refinement used by Bergman and Shahbender (7) and Elion (I) involved simultaneous transmission of a pulse through the specimen and through an ultrasonic delay line set to give the same delay as the specimen in the unstressed condition. This enabled them to make comparative measurements of transit time differences with higher time-base speeds, and led to resolution of velocity changes of IO-~ in 4 in. thick aluminium. McSkimin andAndreatch((8), (9)) used the p u 1se superposition method, wherein the pulse repetition frequency (p.r.f.) is adjusted so that the period is close to an integral multiple of the transit time. It is then possible for waves within succeeding received pulses to arrive in phase, with a consequent increase in the amplitudes observed on the oscilloscope. The transit time is calculated from the p.r.f., the echo number and interpolation from a number of readings. Quoted accuracies were of the order of one part in 104 for measurements on fused silica and single crystals of quartz and germanium. Various interferometer techniques have been used wherein the electrical signals from a train of pulse echoes produced by multiple reflections in the specimen are added to those from another train, with velocity differences leading to phase cancellation of particular echoes. Rollins ((IO), (I I)) used an arrangement, based on the technique of Espinola and Waterman (IZ), in which echoes from a stressed specimen were added to those from an identical unstressed specimen. This technique is quite sensitive, Espinola and Waterman claiming velocity resolution of 5 X I O -6 in their measurements of temperature induced variations in the elastic constants of steel, fused silica, and alkali halide crystals. Rollins estimated a stress resolution of 20 lb ine2 in his measurements on 2s aluminium. Rollins (10) also used an arrangement in which two shear wave echo trains occur within the tested specimen, generated by one transducer shearing at 45 to the applied stress axis, so that the two components of Figure I are generated. Smith (13) used separate transducers, one each side of a nickel-steel specimen, to generate the two principal shear waves, with calibrated variable electrical delay between them so that exact echo cancellation was possible. The transducers were placed opposite one another so that both launched waves through the same volume of the specimen. Their axes of polarization were aligned one

STRESS MEASUREMENT

BY ULTRASONICS

I77

parallel to, and one normal to, the applied stress axis so that each produced a principal wave polarized perpendicular to, and hence unaffected by, the other.
4. THE SING-AROUND SYSTEM

Since the accurate measurement of ultrasonic velocity in a convenient-sized specimen requires accurate measurement of a small time of a few microseconds, a natural step is to convert this time to frequency, which can be measured with the required level of accuracy. The technique, described by Cedrone and Curran (14) is known as the sing-around method and basically involves transmitting a pulse through the specimen from one side

Figure 2. The sing-around system.

and using the electrical signal from the receiving transducer on the other side to retrigger the transmitter, so that a pulse of sound energy sings-around the system at a p.r.f. determined by the ultrasonic transit time. A refined version devised by Myers, Mackinnon and Hoare (I 5), allows the ultrasonic path length to be effectively increased by triggering from an echo among the multiple reflections. This has the additional advantage that the method will work in materials with low attenuation, where triggering on the first received p&e would be disrupted by succeeding echoes. A very sophisticated version of this by Forgacs (16) is quoted as measuring velocity with a precision of one part in IO'. The method is not intended primarily for the accurate measurement of absolute velocity, since the measured transit time includes delays in the electronic triggering system. If the delays are sufficiently constant, however, very small changes in transit time may be detected. The form of the sing-around system used in the present work is shown in Figure 2. Operation is started by first closing the supply switch to the astable multivibrator so that the transmitter is triggered at a comparatively low p.r.f. Figure 3(a) shows the first three received pulses, representing one, three and five traverses of the specimen, with, on the lower trace, the echo pulses triggered by those in the echo amplifier and fed to the gate, The position of the gate pulse, which is set by the manual control on the monostable multivibrator providing the variable delay, is shown in Figure 3 by marker pips. The delay control is then turned to bring the gate pulse into coincidence with the required echo pulse, whereupon the latter is fed to the transmitter drive-pulse generator, causing retriggering of the transmitter. Providing sufficient loop gain then exists in the complete
12

178

D. I. CRJZCRAFT

feed-back loop, and this can be achieved by adjustment of various controls, then the selected echo pulse will sing-around the loop. Figure 3(b) shows retriggering taking place from the second received pulse [the centre pulse in both Figure 3(a) and (b)]: i.e., after the transmitted pulse has made three trips through the specimen. The extra pulse which appears between the first and second received pulses [the third from the left in Figure 3(b)] is the third received pulse from the previous transmission. Finally, the astable multivibrator is turned off, and the sing-around p.r.f. is read on the digital

0.5 V/cm t

O~tiV/Crn T

IO V/cm t

IOOVkm t Gat:-&e 5 pseckm Cdl I.0 flsec/cm -----+

(a)

O,SV/cm

05Vkm

IOVhll T

IOV/cm t 5 flseckm W

(b)

(e)

O.fjV/cm t IO V/cm t l.Opsec/cm __) IO psedcm ----+

(cl

(1)

(b) singing-around after Figure 3. Sing-around oscillograms: (a) not singing-around; 3 trips; (c) (b) expanded x 5 ; (d) as (c) but showing the transmitter pulse (lower trace) ; (e) after 7 trips. Upper traces, output from after 5 trips; (f) singing-around singing-around receiver crystal; lower traces, output from echo amplifier (except in (d)).

frequency meter of the counter-timer type. With a counter gating time of I set, and a typical total ultrasonic transit time of zo psec for three traverses of the specimen, the p.r.f. will be 50 kc/s and the resolution 2 x IO -5. Greater precision can be obtained by extending the gating time to, say, IO set, but it is often more convenient to set the counter to measure the time required for, say, 10~ sing-around cycles. In this example, this time will be 0.2 set, but a I MC/S counter-timer will measure this in units of I psec, giving a resolution of 5 x I0 -j. The limit to the resolution is determined by the stability of the equipment and is about IO-~ under laboratory conditions.
4.1. FACTORS AFFECTING THE SING-AROUND P.R.F. IN A STRESSED BAR

If a long metal bar is stressed longitudinally by tensile or compressive loading at its ends, the distribution of stress will be practically uniform throughout the path of an ultrasonic wave propagated across the bar through its centre. As stress is applied to the system will change inversely as the transit time bar, the p.r.f. u) of the sing-around (t), the latter change being caused both by velocity (c) change and path length change

STRESS MEASUREMENT BY ULTRASONICS

79

due to strain. The velocity change will, in turn, be caused by changes both in the elastic constants (C) and in the density (p) due to strain. The conventional definition of strain is AL eL = -_j L dL AL --as--0, L L

where L is the length of the bar and AL is a small increase in length due to applied stress. In analogy with this, we may define the fractional change in any quantity x as AX dx as--o, Ax e, = --em-.-.X X X so

that, for instance, ef = -e,.

The total path length nT = ct, where T is the thickness of the bar and n is a factor depending on the echo used for retriggering. The total derivative of this gives the lateral strain as

giving the fractional change in velocity as e, = eT-et,


= -peL+ef

(1)

where p is Poissons ratio. Since


c = (qpy, e, = *ec - &ep,

and the fractional change in density, from linear elastic theory, is ep = -eL(I--2p), giving the fractional change in elastic constant
ec =
2ef-eL.

as

(2)

Equations (I) and (2) show that if both the p.r.f. and the length of the bar appear (to a first approximation) to vary linearly with stress, then both the velocity and the elastic constant must vary linearly with stress. Since the linear elastic theory is valid only to a first approximation, the analysis is strictly true only to a first approximation, but the inaccuracy will only be of the same order as the departure from linearity in the stress-strain relationship, which in the elastic strain region is very small.
5. MEASUREMENTS 5. I. EXPERIMENTAL
DETAILS

OF STRESS-INDUCED

VELOCITY

CHANGES

The sing-around system was used to make measurements of the changes in velocity, due to applied uniaxial stress, of longitudinal and shear waves propagated across the centre of bars of nickel-steel (S/NTV), high conductivity copper (999%) and commercial purity aluminium (99.5%). Th e s h ear waves were polarized, first parallel to, and then perpendicular to, the applied stress axis which was parallel with the axis of the bar in each case. The applied stress was confined to the elastic region in each case. Compressive stresses were applied to a 12 in. long, I in. square section nickel-steel bar, a 6 in. long, I in. square copper bar, and a 2.7 in. long, 4 in. x Q in. section aluminium bar, by a hydraulic press,

180 the 12

D. I. CRECRAFT

load being measured by the hydraulic pressure gauge. Tensile stress was applied to a in. long, I in. x Q in. section tensile test piece, machined from the same stock as the above nickel-steel bar. This test piece was also subjected to compressive loading. Measurements were made with I mm thick X-cut quartz crystals driven at 2.5 MC/S for the longitudinal waves coupled by oil films, and 0.5 mm thick Y-cut quartz crystals driven at 45 to 47 MC/S for the shear waves. The latter were coupled by phenyl salicilate (salol ) in early measurements, and later by Aroclor I 254.
5.2. MEASUREMENTS

In Figure 4, the sing-around p.r.f. in the I in. square section nickel-steel bar is plotted versus compressive stress for the three wave types (17), retriggering of the transmitter occurring on the second received pulse in each case. The spread of one count in the readings is due to the ambiguity in the counter-timer technique, both possible readings being obtained at each point.

k*

2.5 me/s longitudinal


: wave I\ %

\ k

v S

41440-

N:-*2*_

4.5 mcksheor wave z 4, 430 polorised perpendidular _ - to axis of stress P 5 E 41,420i S s 41,41041,400 ,<.5 /

3W)_
~*~:-~,

r/ I 1 1

me/s shear wave polarised parallel to oxis of stress

IO I;

14 16 18 20 22 in:*)

Compressive

stress (I031b

Figure 4. Variation of sing-around

p.r.f. with stress in nickel steel S/NTV.

Figure 5 shows measurements made on the nickel-steel test piece in both compression and tension. In this case triggering was on the fourth received pulse, after seven traverses of the 4 in. thickness by the ultrasonic pulse. The measurements on aluminium and copper were made with an improved countertimer with which period, or decades of periods, could be measured, and are shown in Figures 6 and 7. Almost all the measurements show a linear variation of sing-around p.r.f. or period with stress and, together with linear variations in longitudinal strain measured on the specimens, indicate linear variations in ultrasonic velocities and elastic constants when equations (I) and (2) are considered. It might appear from Figures 4,6 and 7 that in each case the velocities of the two shear waves were different at zero stress, since their initial sing-around p.r.f.s differ. This has no significance, however, since retriggering did not necessarily occur at the same

STRESSMEASUREMENT BY ULTRASONICS

181

part of the received pulse in each case, although it will be seen later that their zero-stress velocities do differ, due to preferred grain orientation. Figure 5 confirms expectations that the velocity change would be of opposite sign for tensile stress. The slight difference in slope between the compressive and tensile regions
t
p.r.f. k/s) 34,8808: 8 34,8?02,O , 16 IO 864$/ in-) */ --- 34,050 Sheor wove polarised parollel ,/ *I to stress axis r 1/ Tension,(ton 12 8 4 _ // r 34,090-

/ 8/ 4

, q , I? , 19,2p

Campression,(i031bin-2)

--34,840

--34,830

Figure 5. Variation of sing-around

p.r.f. with compressive and tensile stress in nickel-steel.

19.720 -\

19,710 -

Shear wove polorised parallel

H 3 H k 0

19,700-

1?690-

Longitudinal

wove

1500 2000 Applied compressive load (lb)

2500

Figure 6. Variation of sing-around

period with stress in aluminium.

(about 4%) is probably attributable to calibration differences between the hydraulic press and the tensile testing machine. The compressive results show curvature at higher loads, which is probably due to the slight curvature of the test piece observed at these loads. Since the I in. square section bar of the same material showed no such effect, it is most probable that the curvature of the specimen caused an increase in the effective crystal bond

182

D.

I. CRECRAFT

thicknesses, giving rise to a longer path length. The linear part of the compressive curve has the same slope as that for the I in. square bar, confirming the reproducibility of measurements made by the sing-around technique.
35.030 r
Shear wove polorised parallel

Shear wove polorised perpendicular, to stress 4.5 me/s 3 trips

4 2 Compressive

6 stress (IOlb

6 in?

IO

Figure 7. Variation of sing-around period with stress in copper.

The absolute velocities in the three bars were measured by the pulse-echo and the Lame constants, X and CL, were calculated from p = p& and h = p(&where cl0 and czo are thelongitudinal Poissons ratio was calculated from
2&J,

method,

and shear wave velocities in the unstressed material.

A
* = z(X+u)

and equations (I) and (2) were then used to find the fractional changes in velocity and in elastic constants , shown in Table 2. The values of the stress-sonic coefficient quoted in Table I were computed from these results, with the convention that S is the fractional velocity difference between the shear wave polarized parallel to the stress axis and that polarized perpendicular to the stress axis, per unit compressional stress. Using this convention, the values of S calculated from other workers and the authors work on metallic materials are all positive. In the present case it can be seen from Table 2 that this is because the wave polarized parallel to the stress axis suffers by far the larger velocity change, and always increases in velocity under compressive stress. Confirmation of this is not possible from the measurements made by others of birefringence, due to the difference between the two wave velocities. On the macroscopic scale, the much larger velocity change of the parallel-polarized wave is attributable to the fact that the applied uniaxial stress causes shear strain in the plane in which the wave motion shear occurs, so that the velocity is determined by a shear

STRESS MEASUREMENT BY ULTRASONICS

183

modulus which has changed due to applied shear. On the other hand, the perpendicularpolarized wave shears in the plane of the applied uniaxial stress where, in an isotropic body, there is no resultant shear deformation. The fact that in practice the velocity does change in the latter case shows that this macroscopic viewpoint does not provide a complete description of the situation.
TABLE 2

Sing-around
Specimen

measurements of stress-induced changes


Steel Aluminium Copper

Fractional change

in S.A.

period (t)
Fractional change in velocity S = e,-e, Fractional change in elastic constants Tolerances

t1,
tzz

I7
12

95
00

8.2

t,

Clu CZZ
C2Y

-47 -0.6 -0.2 57 59


01 10

Cl, C2Y
C2Z

12.9
20.1

-24.1 -6.1 34 275 24. I -9.2 9.8 579 + 02

44 -46 -5.3 - 1.5 75 9.0 -7.9 -0.3 17.6 rto.2

Figures are parts in 10s per IOOO lb in-z. Subscripts: all waves propagated in x direction, and ~y=longitudinal wave propagated transverse to stress axis, y; zy =shear wave polarized parallel to stress axis; zz= shear wave polarized normal to stress axis.

5.3.

THIRD-ORDER

ELASTIC CONSTANTS

elastic theory derives from an expression for the strain energy in terms of strain products of the first and second degree, with corresponding first-order and second-order coefficients. This expression is differentiated with respect to strain in order to produce an expression for stress in terms of the first-order coefficient (corresponding to initial hydrostatic strain) and the products of the strains and the second-order coefficients. The latter are thus the familiar elastic constants of linear theory. If the expression for the strain energy is now considered with additional third-order strain products, then the first differentiation yields, in addition to the above terms, second-degree strain products with third-order coefficients. A further differentiation with respect to strain
gives the elastic constants in terms of the second-order coefficients and the products of the third-order coefficients and the strains. Murnaghan (6) introduced three third-order constants, I, m and n for the case of a normally isotropic body, taking the Lame constants, X and CL,as the second-order coefficients. Hughes and Kelly (5) derived expressions for elastic wave velocities in terms of these five constants in seven cases, two of hydrostatic pressure and five of simple uniaxial stress with the stress axis, wave propagation direction, and polarization (in the shear wave cases) coinciding with one or other of the axes of an orthogonal system. Three of these expressions, of interest in the present case, are

The linear, or first-order,

PO&J = A+2p-go [22-$(m+X++)

184
PO& =

D. I. CRJZCRAFT

, m+;;+h+zp CC-T I 3Ko [

PO&

= p-$i-[m-(%)n-zh].

In these, p. is the density and Ko = A+ 2~/3 is the bulk modulus, under zero stress conditions. T is the uniaxial compressive stress, which is applied in the direction given by the second subscripts on the velocities. Of the first subscripts, I refers to longitudinal waves and 2 refers to shear waves, the latter being polarized in they direction. Propagation is in the x direction in all cases. In general, c=co(I -e,), and since e, is very small (the largest observed value being about IO-~) to the required level of accuracy c2 = c$(1 - ze,), so that the velocities given by equations (I), (2) and (3) should vary linearly with stress, as is confirmed by the present measurements. The Lam6 and Murnaghan constants derived from the three wave velocities measured during the present work are shown in Table 3 together with those calculated by Hughes
TABLE 3

LamC and Murnaghan constants (bars x 10~) Lam6 constants


Material x II Mumaghan constants 1 m n

Source

Polystyrene Armco-iron Pyrex Nickel-steel S/NTV Copper (999%) Aluminium (99 %)

0.2889 * 000 I
11 00 + 0.04

I353 _c0.003 10.90 _co*1 10.40 +0.1 6.10 ko.1

0.1381 + 0001 8.20 ko.10 275 * 0.03 8.17 f 002 4.60 + 001 2.49 * 001

- 1.89 f. 0.32 - 34.8 +6.5 I4 +4-o -5.6 f 20 542 +3o -4.7 f 2.5

- I33 + 0.29 - 103.0 L!z 70 9.2 f5.0 -67.1 f 0.6 - 372 + 0.5 - 342 + 10

- 1.00 +0*14 110 +I10 42 f35 - 78.5 + 0.7 -40.1 +0.5 - 24.8 rt 10

Hughes and Kelly

Crecraft

and Kelly (5) from their measurements on three other elastic materials. It should be noted that n is given by equations (2) and (3) as

-(@+

I)&

where S is the stress-sonic coefficient, and thus n can be determined from measurements of birefringence alone. The values for nickel-steel differ somewhat from those published previously by the author (17), since they include corrections for slight errors in the absolute values of velocities upon which previous calculations were based.

STRESS MEASUREMENT

BY ULTRASONICS

1%

6. RESIDUAL

ANISOTROPY

6. I. THE PROBLEM

OF LAUNCHING

SHEAR WAVES

It has been seen that shear wave velocities are affected by stress to an extent determined by their direction of polarization with respect to the principal stress axes in the plane normal to their direction of propagation. This is extremely important in the determination of residual stresses, since, in analogy with photoelasticity, the difference in velocity between the two principal shear waves indicates the magnitude of the stress, and their directions of polarization indicate the principal stress axes. Moreover, an accurate measurement of ultrasonic path length is unnecessary. The use of longitudinal waves would involve the measurement of path length and transit time to within one part in ten thousand even to detect large stresses, would not indicate the stress axes, and would be very sensitive to slight changes in material properties from sample to sample. For these reasons, the use of shear waves is imperative, and these waves must enter the specimen at normal incidence so that, when the transducer is turned to change the angle of polarization, the path through the specimen is unaltered. Thus it is not possible to use the familiar shear-wedge, used in ultrasonic flaw detection, wherein shear waves are generated by mode conversion of a longitudinal wave incident at an angle to the boundary between the specimen and an oil film interposed between the wedge and the specimen. An attempt to launch shear waves directly from an appropriate transducer will result in transverse slip if a liquid couplant is used, and is only completely successful if the transducer is cemented to the surface. This latter arrangement is impracticable since changing of the polarization direction involves different bond thicknesses, apart from the time involved. There are some very viscous liquids which have sufficient dynamic shear stiffness to couple shear waves through a very thin film. Aroclor 1254 (a chlorinated diphenyl; Monsanto Chemicals Ltd.) is one of these and at 3 MC/S its use involves about 6 dB loss per transmission compared with a solid cement, phenyl salicylate ( salol ). In an attempt to launch shear waves at normal incidence without cementing to the specimen and to improve on the loss involved with liquid couplants, a transducer assembly was devised (IS), relying on interface pressure for coupling. A Y-cut, or rotated Y-cut, quartz crystal, or lead-zirconate-titanate disc electrically polarized for thickness shear, is cemented by phenyl salicylate to a metal coupling piece. This is pressed hard against the surface of the specimen by the application of force to the cover or cap as shown in Figure 8. In this way it is possible to apply an interface pressure high enough for almost complete transmission of the incident energy without the application of injurious stresses to the crystal. The loading force is applied via the steel ball, which ensures symmetrical pressure distribution at the interface. Figure 9 shows how the shear energy transmitted through the interface increases with apparent interface pressure for a steel coupling piece on a ground nickel-steel surface. It can be seen that an apparent pressure of about 20,000 lb inW2 is sufficient to cause complete suppression of reflections from the interface and almost lossless transmission. This requires a load of IOOO lb, since the front face has 4 in. diameter. In order to measure the difference in the velocities of the two principal waves over a given path through a specimen, the transducer is pressed onto the surface and used as a common transmitter-receiver to observe the transit time of a pulse reflected back from a parallel opposite face. The load applied to the transducer affects both velocities equally and has no effect on the birefringence due to the principal stresses in the plane normal to propagation, which are those resolved by this method. Since the velocity difference is usually very small, the relative phase of the received echoes being less than one cycle, birefringence as such is not apparent and as the transducer is rotated the velocity appears

186

D. I. CRECRAFT

simply to have maximum and minimum values 90 apart. Under these conditions it can be shown that, if the relative phase difference is 4, then the apparent phase E of the received $ in.
sieel boll \

Holes for connecting wires

Figure 8. Single-crystal transducer.

10.000 lnterfoce

20,000 pressure (lb/in*) 1000 load (lb) 1500

500 Transducer

Figure 9. Pressure-coupling characteristics of 4 in. steel transducer on polished steel surface. Curve A, polished transducer with liquid; curve B, polished transducer on dry surface; curve C, transducer face as turned.

wave will vary from zero to +, as the transducer polarization is rotated from a direction parallel to one anisotropy axis to the other at right-angles to this, following the law

where 0 is the kgle turned through. The accuracy of the stress reading does not, therefore, depend critically on the angle of polarization, since da/d@ is small when 0 is near zero and 90. On the other hand, the location of the principal stress axes is not very precise, and

STRESS MEASUREMENT

BY ULTRASONICS

87

consequently there is no need for more than three or four readings at 45 intervals under these circumstances. In use, the transducer is loaded by either a small hand press or an hydraulic press. The latter has the advantage that the load can be monitored by measurement of the hydraulic pressure, ensuring the same interface conditions and strain in the specimen at every angular position of the transducer. The velocity measurement process is simplified further by the use of a four-crystal transducer, which generates two shear waves with orthogonally disposed polarizations, obviating the need for rotation in many cases. All four crystals are polarized along radii of the transducer and in use one of a pair of diametrically opposite crystals is used to generate a shear wave and the other to receive the reflection from the back-wall of the specimen. The transmitter and receiver are then switched to the other pair of crystals, which transmit and receive a wave polarized in a direction at 90 to the first. Since the sing-around system is not readily applicable to a common transmit-receive transducer, velocity differences have been observed on a delayed time-base oscilloscope with a time-base calibration of IOO nsec cm- l. Taking the display resolution as I mm gives IO nsec time resolution which, in the case of the I in. thick nickel-steel specimen with a round-trip time of 1542 psec, is a fractional velocity difference of 6.32 x IO-~, corresponding to a stress resolution of 485 ton inF2, or about 5 ton inm2. 6.2. PREFERRED ORIENTATION Early measurements with the pressure-coupled transducer revealed elastic anisotropy in all the specimens used previously for measurements of stress-induced changes. The anisotropy was very small and constant throughout each specimen, with a principal axis parallel to the bar axis, and evidently due to preferential grain alignment induced by rolling or extrusion. The fact that the anisotropy was constant throughout each specimen meant that it could not be attributed to residual stress, since the forces in a closed system must sum to zero. The fractional velocity differences observed were o-27$ in both nickel-steel specimens, 0.55~h in the copper bar, and 3.3% through the thickness (0.5 in.) and 2.6% through the width (0.75 in.) of the aluminium bar. In addition, a Duralumin bar gave a difference of 0*3% whilst a Nimonic alloy piece exhibited a difference of 1.4% and another (NIM 8oA) of 17*oO/~. In all cases except that of the copper bar, the wave polarized parallel to the bar axis propagated with the greater velocity. Preferred orientation in the direction of working is well known, and is readily measurable in extreme cases, such as drawn wires, by static measurement of elastic constants parallel to, and perpendicular to, the extrusion or rolling axis. Surface effects are measurable by X-ray techniques, but the proportion of grains lined up in a given direction must be at least a few per cent to be noticeable. Ultrasonic birefringence is certainly the most sensitive method for bulk measurements, since velocity differences of the order of 0.1 y. are readily observable. Sullivan and Papadakis (19) have also observed small preferred orientation in rolled steel by ultrasonic birefringence with crystals cemented to the surface with polarization at 45 to the rolling direction. Imagining a model with some grains aligned parallel to the rolling direction, embedded in a sea of randomly-orientated grains, they derived the expression

ts-ty = L&rc,)(c,c,)-1,
where t, and tY are the transit times of the two waves polarized parallel to and transverse to the rolling direction, L is the path length, K is the fraction of grains aligned preferentially,

188

D. I. CRECRAFT

and c, and cY are the shear wave velocities in a single crystal, or grain, given by

with polarizations along a [IOO] and a [I IO] direction respectively. expression in terms of fractional velocity change it becomes
k _

Rearranging

the

-C.

co cy-cx

'xcY

If the values of the elastic constants of iron are used, this gives k = 0.4% as the proportion of grains in the nickel-steel bar with their [I IO] direction parallel to the bar axis. Sullivan and Papadakis observed o* I o/o alignment. Even for surface measurements an ultrasonic method is the most sensitive, the ultrasonic goniometer developed by Bradfield (20) measuring accurately the surface wave velocity, and hence the surface elasticity, in various directions. 6.3. PLASTIC DEFORMATION In the final sections work is described on specimens containing residual stresses induced by inhomogeneous plastic deformation. In order to investigate the effect of plastic strain in the absence of macroscopic residual stresses, the nickel-steel test-piece described previously was subjected to a few per cent uniaxial tensile plastic strain. Unfortunately, the strain measurements were upset by necking of the specimen in the region of the extensometer. The last reading obtained was 3.25% under load and from extrapolation of the stress-strain curve it seems likely that the residual strain in the parallel region after removal of the load was about 4%. Before strain, the preferred orientation caused a fractional velocity difference of between 0.20% and 0.23%, the spread representing + I nsec tolerance on the velocities across the 3 in. section, with the parallel polarized wave propagating faster. After plastic deformation, and removal of the stress, this difference was increased to between 0*39O/~ and 0.42O/~.Since strain of the order of 50% is necessary to cause appreciable change in preferred orientation (Doan (zI)), it seems likely that this increase in velocity difference was due to dislocation movement alone. The wave polarized parallel to the stress axis propagates as shear in a plane parallel to the deformation direction, and must attempt to shear slip planes where some dislocations are pinned following the plastic strain. Since the preponderance of slip planes lie at 45 to the principal stress axis, the wave polarized in the plane normal to the deformation direction must have a similar action, but since the plastic strain in this plane is half the principal strain, the increase in macroscopic shear modulus due to pinned dislocations is less for this wave than for that polarized parallel to the specimens axis. The above argument applies equally well to tensile and compressive plastic strain. This is of practical importance in stress analysis, since it means that the velocity of the parallelpolarized wave always increases with respect to that of the other wave regardless of the sign of the plastic strain, but under elastic conditions is sensitive to the sign of the strain. This is demonstrated in the next section. Rollins (II) found no further birefringence induced by plastic strain of 0.36% in a CIOI~ steel specimen. Evidently this order of strain causes a fractional velocity difference of less than o-01 %, the apparent resolution of Rollins measurement. 6.4. RESIDUAL
STRESSES IN .4 BENT BAR

In an effort to learn something more of the combined effects of residual stresses and grain orientation, the I in. square nickel-steel bar was subjected to simple beam loading sufficient to cause permanent bending. Under these conditions the upper and lower halves

STRESS MEASUREMENT

BY ULTRASONICS

189

of the beam suffer plastic deformation near the centre of the beam, where the bending moment is greatest (see Figure IO(a)). The stress parallel to the beam axis caused by the bending moment falls off from the yield stress in the region of plastic flow to zero at the centre of the section (the neutral axis) and reverses sign on the other side of the centre (Orowan (22)). When the load is removed, the inner part of the bar, in an effort to straighten, exerts elastic stress on the outer layers, with the result that the latter become stressed elastically with the opposite stress to that which caused their deformation, whilst the adjacent layers which suffered only elastic strain are now elastically strained by stresses in the same sense as before, as shown by Figure IO(b).
Compressive

1Tensile I *
(0) Stress during strain distribution plastic

; .5 = 2 g z 0 k 9

(a) -5.0 i-7.5 +10 0 -10 +5 0 t5.0 (b) Stressdistribution after removal of load

@ (b)

Sonoelostic reodmgs (compression I-0*5ton in-)

Readings after subtraction of 3 Odue tooriginal preferred orientation (d)

Cc)

Figure

IO.

The bent bar.

Measurements made across the centre of the bar are recorded in Figure IO(C), these representing the difference in round-trip times of the two waves, that polarized parallel to the bar axis being the faster. The original anisotropy gave rise to a difference of 30 nsec, and this figure subtracted from the readings gives those shown in Figure IO(d). This last set is not balanced and evidently does not represent the residual stress system. Subtraction of a further 25 nsec from all but the centre reading, to allow for the plastic deformation, yields a more nearly balanced set, which is shown in Figure IO(b) after conversion to stress. (IO nsec represent 5 ton in2, see section 6. I .) This, then, confirms the fact that plastic strain causes a greater increase in the shear modulus in the direction of the greatest strain, regardless of the sign of the strain. 6.5. RESIDUAL STRESSES IN A DISTORTED RING Some residual stresses may exist in circular parts in the form of hoop stresses. Figure I I shows a specially prepared nickel-steel specimen of approximately 2 in. square section and originally forged as a circular ring from the cast billet in such a way that the grain flow was expected to be radially symmetrical, with the flow across the section as shown. The ring was then plastically deformed, hot, between the jaws of a drop forge. The system of residual stresses and plastic strains is not expected to be simple, therefore.

190

D. I. CRECRAFT

Wave round-trip time differences observed across the section at eight positions round the ring are shown in Figure I I, in units of zo nsec, representing approximately 5 ton in.-2 in the 2 in. section. The wave polarized in a direction tangential to the ring was the faster in every case. Positions A, C, E and G are those of greatest deformation and must contain most residual anisotropy. The positions B, D, F and H lie midway between these and, since each lies between positions of opposite bending moment, are assumed to be stressfree, or neutral positions. It is also assumed that no plastic deformation took place here, and so the grain orientation should be as before distortion of the ring.
7

1416

(b)

Figure I I. The distorted ring: (a) sonoelastic readings ; (b) expected grain flow in section. Although the readings vary considerably across the section at the neutral positions, all four positions are similar, confirming the fact that grain flow during manufacture was radially symmetrical. The readings have been averaged for each radius and the curve in Figure 12 shows the way the average anisotropy varies across the section, indicating greater preferential orientation near the outside edge, as shown by the section in Figure I I. Positions A and E are similar since, in both, the outside edge of the ring has been plastically stretched and the inside compressed. The two similar sets of readings at these positions have been averaged, and the averages of the neutral positions at corresponding radii have been subtracted, giving the curve shown in Figure 12. This procedure was repeated for the readings at positions C and G, which are also similar, giving the final curve in Figure 12. The two curves for the residual stress regions cannot be true representations of the residual stresses across the sections, since the bending moments do not sum to zero, and in order to achieve a balanced distribution it is necessary to subtract further figures in both cases. Whilst this again confirms the fact that induced birefringence is independent of the sign of plastic deformation, it is difficult in this case to find a simple procedure to yield balanced distributions. If the bending operation were of a simple nature, then no plastic deformation would occur at the centre of the section. The curve for positions C and G shows this to be so for these positions, but this is certainly not true for positions A and E.

STRESS MEASUREMENT

BY ULTRASONICS

191

It appears that overall plastic strain has taken place in the latter case, and any further attempts at analysis would be pure conjecture unless some independent means of separating residual stress from the other effects can be found.

(al

OutsidsGcdge

,,I 1 + in ihtervals (cl

I Inside edbe

Figure 12. Anistropy in distorted ring: (a) neutral positions, B, D, F and H; (b) positions C and G ; (c) positions A and E. 7. CONCLUSIONS

Although the problem of residual stress analysis is by no means solved, it has been shown that elastic anisotropy of small orders can be reliably measured, and that in cases where anisotropy due to preferred grain orientation is of a regular or predictable nature so that its variation from point to point can be allowed for, then residual stresses can be estimated from the remaining anisotropy not accounted for by grain alignment. The application of more precise measurement techniques, such as the sing-around system, to the pressure-coupled transducer will lead to simple measurements of greater resolution, so that if a way is found for reliable separation of the effects of residual stress and preferred orientation, ultrasonic residual stress analysis will become a useful technique.
ACKNOWLEDGMENTS

The majority of this work was performed in the Electronic and Electrical Engineering Department of the University of Birmingham, and grateful thanks are due for the help and advice of the Head of Department, Professor D. G. Tucker and of Dr. L. Kay (now Professor at the University of Canterbury, N.Z.). Thanks are also due to Rolls-Royce Ltd., Derby, for their generous financial support of this project, and in particular to Mr. J. D. Hislop of this company for his interest and advice and his help in supplying specimens for testing. The plastic strain measurements on the tensile test piece were carried out in Lanchester College, and Mr. G. A. Johnson of the Mechanical Engineering Department is to be thanked for his kind assistance in operating the tensile test machine.

192

D. I. CRECRAFT

REFERENCES I. H. A. ELION 1960 Non-destruct. Test. 18, 180. Recent advances in new areas of non-destructive testing. 2. R. W. BENSONand V. J. RAELSON1959 Product. Engng. 30. Acoustoelasticity. 3. R. B. HEYWOOD1952 Designing by Photoelasticity. Chapman and Hall. 4. P. W. BRIDGMANI948 Proc. Am. Acad. Arts Sci. 76, 55. The compression of 39 substances to I oo kgjcm2. D. S. HUGHESand J. L. KELLY 1953 Phys. Rev. 92,s. Second-order elastic deformation of solids. 2: F. D. MURNAGHAN 195 I Finite Deformation of an Elastic Solid. London : J. Wiley. 1958r. Appl. Phys. 29, 12 , 1736. Effect of statically 7. R. H. BERGMANand R. A. SHAHBENDER applied stresses on the velocity of propagation of ultrasonic waves. 8. J. MCSKIMIN 1961 J. acoust. Sot. Am. 33, I. Pulse superposition method for measuring ultrasonic wave velocities in solids. 1962 J. acoust. Sot. Am. 34, 5. Analysis of the pulse super9. J. MCSKIMIN and P. ANDREATCH position method for measuring ultrasonic wave velocities as a function of temperature and pressure. IO. F. R. ROLLINS Dec. 1959 W.A.D.C. tech. Rep. 59-561. Study of methods for non-destructive measurement of residual stress. tech. Rep. 61-42 Part I. Ultrasonic methods for nonII. F. R. ROLLINS May 1961 W.A.D.C. destructive measurement of residual stress. 12. R. P. ESPINOLAand P. C. WATERMAN19583. appl. Phys. 29,718. An unltrasonic interferometer for measurement of velocity changes in solids. acoustoelastic 13. R. T. SMITH 1963 Ultrasonics I, 3. Stress-induced anisotropy in solids-the effect. and D. R. CURRAN1954J. acoust. Sot. Am. 26,963. Electronic pulse method for 14. N. P. CEDRONE measuring the velocity of sound in liquids and solids. IS. A. MYERS, L. MACKINNONand F. E. HOARE 1959J. acoust. Sot. Am. 31, 2. Modifications to standard pulse techniques for ultrasonic velocity measurements. 16. R. L. FORGACS196oJ. acoust. Sot. Am. 32, 12. Improvements in the sing-around technique for ultrasonic velocity measurements. 17. D. I. CRECRAFT1962 Nature, Lond. 195, No. 4847. Ultrasonic wave velocities in stressed nickelsteel. 18. D. I. C&CRAFT 1964J. Sound. Vib. I, 4. Launching ultrasonic shear waves into solids at normal incidence by pressure coupling. 1962J. acoust. Sot. Am. 34. Ultrasonic double refraction in 19. P. F. SULLIVANand E. P. PAPADAKIS worked metals. 20. G. BRADFIELD 1964 Use in Industry of Elasticity Measurements with the Help of Mechanical Vibrations. N.P.L. Notes on Applied Science No. 30. London: H.M.S.O. 21. G. E. DOAN 1953 The Principles of Physical Metallurgy, 3rd Ed. London: McGraw-Hill. 22. E. OROWAN 1959 Internal Stresses and Fatigue in Metals, Gen. Motors Res. Labs. Elsevier. Pub. Co. Causes and effects of internal stresses.

Вам также может понравиться