Anda di halaman 1dari 80

BACK TO BASICS - ULTRASONICS

1 BASIC PRINCIPLES OF SOUND



Sound waves are vibrations of the particles of solid liquid or gas through which the sound is passing.
Each particle oscillates about a mean position and in doing so causes a similar vibration to be taken
up by its neighbour. The resulting disturbance radiates out from the source as a sound wave.

Sound waves are therefore a form of mechanical energy that can only exist in a solid liquid or gas
and not in a vacuum. Essentially, there are two requirements for sustaining a vibration: there must be
something to vibrate and some force that will always try to return that something to its original
position. In other words, there must be MASS and ELASTICITY. This is illustrated in figure 1.1a
below. A weight is suspended from a beam by a spring. The weight (W) provides the MASS and the
spring provides the ELASTICITY. At rest, the force of gravity (G) acting on the weight is balanced by
the tension (T) in the spring.



















Fig. 1.1a
G
T
W




A
C
B


Fig. 1.1b
Down
Time
Up
Amplitude
One cycle










If the weight is pulled downwards from its rest position (A) to position B, the tension in the spring will
increase. When the weight is released, the weight will accelerate back towards position A reaching its
maximum velocity at position A when the forces T and G are again equal. The momentum of the
weight travelling at speed will cause the weight to overshoot position A. Immediately the tension in
the spring is less than the force of gravity and the weight will begin to decelerate until it comes to rest
at position C. Because force G is now larger than T, the weight will start to descend again;
overshooting position A again until the increasing tension in the spring eventually stops the downward
movement. At this time, the whole cycle of events starts again and continues until friction and air
resistance losses gradually bring the oscillations to a stop.
Figure 1.1b is a graph of the displacement of the weight, during this up and down motion, against
time. In the diagram, two points on the graph are shown where the weight is doing the same thing,
travelling upwards and passing through position A on consecutive passes. The distance (time)
between these two points represent one complete cycle of the oscillation. The number of cycles of
oscillation completed in a given period of time (usually one second) is called the Frequency of the
oscillation. The maximum displacement of the weight from its normal rest position is called the
Amplitude of the oscillation.

One of the best examples of an oscillating source of sound that can be used later in describing the
action of an ultrasonic test probe is the guitar. The strings of a guitar are elastic and pre-tensioned to
produce a particular frequency of vibration. Each string is distorted by the guitarist to stretch the string
and then released. As soon as it is released, the string begins to oscillate about its mean position at
the resonant frequency of that string. Shortening the string using a finger to hold the string against
one of the frets can change the frequency. The human ear recognises the frequency as the Pitch of
the note produced. The Loudness of the note depends on how far the guitarist distorted the string
before t was released, in other words, the Amplitude of that distortion.

The mass of woodwork to which the string is attached amplifies the sound and adds its own harmonic
frequencies to produce a range of notes to give the characteristic richness of tone to the instrument.
The band of frequencies produced is called the Bandwidth of the sound in ultrasonics.

THE ACOUSTIC SPECTRUM

Sound waves are described above as the oscillation of particles of solids, liquids or gases. The
human ear can only detect a small range of possible vibration frequencies, roughly between 16 cycles
per second and 20,000 cycles per second. In theory, however, there is a limitless spectrum of
frequencies and that are possible even if humans cant hear the whole range. The spectrum is
illustrated in figure 1.2 below: -











Typical test range
Subsonic
range
Audible Range Ultrasonic range
20MHz 0.5MHz
0
10 100 1000 10,000 100,000 1,000,000 10,000,000 100,000,000 Hz

Fig. 1.2 Acoustic Spectrum
The unit used to denote frequency is the Hertz, abbreviated as Hz, where 1Hz is one cycle per
second. One thousand Hz is written as 1KHz (Kilo Hertz) and one million Hz as 1MHz (Mega Hertz).
The part of the spectrum from zero to 16Hz is below the range of human hearing and is called the
Subsonic, or Infrasonic range. From 16Hz to 20KHz is known as the Audible range and above
20KHz as the Ultrasonic range. Ultrasonic flaw detection uses vibrations at frequencies above
20KHz.

Most flaw detection takes place between 500KHz and 20MHz although there are some applications,
for example in concrete, that use much lower frequencies and there are special applications at
frequencies above 20MHz. In most practical applications in steels and light alloys, frequencies
between 2MHz and 10MHz predominate. Generally the higher the test frequency, the smaller the
minimum detectable flaw, but it will be shown in following articles that higher frequencies are more
readily attenuated by the test structure. Choosing an appropriate test frequency becomes a
compromise between the size of flaw that can be detected and the ability to get sufficient sound
energy to the prospective flaw depth.

MODES OF PROPAGATION
Sound energy travels, or propagates, outwards from the source of the vibration as the oscillation of a
particle of solid, liquid or gas disturbs the neighbouring particles so that the neighbour takes up the
oscillation. It will take time for the disturbance, called the sound wave, to reach a given distance from
the source. This is a measure of the velocity of sound in a given medium. It will be shown that this
velocity varies with the characteristics of each material and the way in which the disturbance is
transmitted from one particle to the next. The different ways in which the disturbance may be
transmitted are known as the Modes of Propagation.

The different modes of propagation come about because solids, unlike liquids and gases, have a
modulus of rigidity as well as a modulus of elasticity. Solids Liquids and gases all show resistance to
compressing or stretching. In the case of solids we refer to this resistance as Youngs Modulus of
Elasticity (E) for the material. The elasticity of a solid is plotted when a tensile test is carried out and
from the resulting graph the Ultimate Tensile Strength (UTS) of the material can be derived. The
Modulus of Rigidity (G) is the materials resistance to a shear load.

COMPRESSION WAVE MODE

Because liquids and gases have no modulus of rigidity, sound waves can only propagate by using
their resistance to tension and compression. This type of sound wave is called the Compression
Wave. Compression waves can exist in solids, liquids and gases because they all have elasticity.
Compression waves are also known as Longitudinal waves, and sometimes as Plane waves The
individual particles of the solid liquid or gas oscillate about their mean position, and the direction of
propagation of the compression wave is in the same plane as the particle motion as shown in
figure 1.3.


Particle Motion






Direction of Propagation

Fig. 1.3

SHEAR WAVE MODE
Shear waves only exist in solids and rely on the modulus of rigidity of the solid under test, they can
exist on their own or co-exist with compression waves and surface waves. Shear waves are also
sometimes called Transverse waves. Again, the individual particles of the solid oscillate about their
mean position, but the direction of propagation of the shear wave is at right angles to the
particle motion. This is illustrated in figure 1.4.








P
a
r
t
i
c
l
e

M
o
t
i
o
n


Direction of Propagation


Fig. 1.4

SURFACE WAVE MODE
At the surface of a solid, a complex mode of oscillation can exist in which the particle motion is mainly
perpendicular to the direction of propagation as with the shear wave, and partly in the same plane as
the direction of propagation as with the compression wave. This mode of propagation is called the
Surface wave or Rayleigh wave. Surface waves only affect the surface layer of the solid to a depth
of about one wavelength, and have the advantage that they follow the surface contour of the object
and only reflect at an abrupt change such as a corner or crack. For the surface wave, the particle
motion is elliptical with the major axis of the ellipse at right angles to the direction of
propagation. This is shown in figure 1.5.


Fig. 1.5
Direction of Propagation

Elliptical
Particle
Motion




LAMB WAVE MODES
Lamb waves, like Surface waves, propagate parallel to the test surface and have an elliptical particle
motion. They occur when the thickness of the test material is only a few wavelengths at the test
frequency and where the test piece is of uniform thickness. Lamb waves fill the wall thickness and
propagate along the major axis of the component. They can travel several meters in steel, so they
can be used for rapid scanning of plate tube and wire. Recent developments for rapid corrosion
monitoring in buried pipes use Lamb waves under the name Guided Waves. The wall of the
component flexes so that the sound ripples along the material distorting both surfaces. Figure 1.6
illustrates a type of Lamb wave where the crests of the wave on the near and far surfaces coincide.
These are called Symmetrical Lamb Waves. Figure 1.7 shows another type of Lamb wave where the
crest on one side coincides with a trough on the other. These are called Asymmetrical Lamb Waves.



Original plate surface Particle motion



Fig. 1.6



Original plate surface
Particle motion





Fig. 1.7

CREEPING (LATERAL) WAVES
There is a special type of compression wave called a Creeping or Lateral wave. It sneaks along
the surface rather like a surface wave, its use is described later under TOFD techniques.






Reference: - Ultrasonic Flaw Detection for Technicians - Third Edition, June 2004 by J. C. Drury

Published in INSIGHT magazine November 2004
BACK TO BASICS ULTRASONICS

2. PROPERTIES OF SOUND WAVES
VELOCITY
Sound travels at different speeds through different materials. This is noticeable when, for
example, a railroad worker is observed from a distance striking a rail with a hammer. Since
the speed of light is much faster than that of sound, the observer first sees the hammer strike
the rail. If the observer is also close to the rail, the next event is the sound of the blow coming
out of the rail and finally the airborne sound is heard.
This tells us that the speed of sound in the rail is faster than the speed of sound in air. It is
true that sound travels faster in liquids than in gasses and faster in metals than in liquids.
However, it is also true that sound travels at different speeds in different metals. There is a
distinct speed of sound for each material and in ultrasonics this is called the VELOCITY of
sound for that material. This being so, it would be useful to have an understanding of the
reasons for the difference.
Imagine two pairs of identical steel balls, one pair joined by a strong compression spring and
the other pair by a weak spring. If one of each pair is moved towards its partner at a constant
speed, the spring joining the pair will start to compress. Eventually there will be enough
compression in the spring to overcome the inertia of the second ball and it will start to move.
As shown in figure 1, the second ball will move sooner for the pair connected by the stronger
spring.



M
1 P
P =Constant force for both pairs
M
1
=Movement of strong spring second ball
t
1

=Time taken to movement M
1
M
2
=Movement of weak spring second ball
t
2
=Time taken to movement M
2



M
2
t
1
0
Time

P






t
2

Fig. 1

In the analogy, the balls represent the particles of solid, liquid or gas through which the sound
wave is propagating and the springs represent Youngs Modulus of elasticity E. The
suggestion made by the analogy is that the disturbance will pass more quickly from one
particle to the next in a material having greater elasticity. In other words, the velocity of a
compression wave will be higher for greater values of elasticity. This is generally the case but
there is another main factor affecting velocity, and that is the density of the material.
Consider another situation in which pair of aluminium balls and a pair of lead balls replace the
steel pairs in the above analogy but with each pair joined by springs of equal strength. The
inertia of the lead ball is greater than that of the aluminium ball and this time it will take longer
to get the lead ball moving. This suggests that the compression wave velocity will be lower for
high-density materials than for low-density materials. Density and elasticity are the dominant
factors affecting velocity and the expression for the compression wave velocity in a fine wire
is taken to be: -

E
V
co
=
However, when we have a bulk of material, the sample is more rigid than a fine wire giving an
effective increase in Youngs Modulus and we need to modify the expression to take account
of Poissons Ratio. During a tensile test, to measure the strength of a metal sample, the
diameter of the sample reduces as the sample is stretched. The change in diameter divided
by the change in length is Poissons Ratio. Considering all these factors, the velocity of a
compression wave in a bulk material can be calculated from the following formula: -

( )( )

2 1 1
1
+

=
E
V
c

Where
c
V =Compression wave velocity
E =Youngs Modulus of Elasticity
=Material Density
=Poissons Ratio

Shear waves are able to exist in solids but they do not travel at the same velocity as the
compression wave in a given material. This is because it is the Modulus of Rigidity, rather
than Youngs Modulus, that dictates the velocity, and the modulus of rigidity is lower than the
modulus of elasticity. This means that the shear wave velocity is always slower than the
compression wave velocity in a material. In Liquids and gases, the value of the modulus of
rigidity is so low that shear waves cannot propagate. As a rule of thumb, the shear wave
velocity is roughly half the compression wave velocity. The velocity can be calculated from: -
( ) +
=
1 2
1 E
V
s
Or, alternatively

G
V
s
=
Where
s
V =Shear wave velocity
G =Modulus of Rigidity
=Material Density
=Poissons Ratio

Surface (Rayleigh) waves also have their own particular velocity, which is generally taken to
be approximately 90% of the shear wave velocity.

Although the velocity for each of these modes of propagation can be calculated, it requires a
precise knowledge of all the parameters, and these are not usually available to the ultrasonic
practitioner. Parameters such as density and strength vary with alloying, heat treatment,
casting, rolling and forging processes all of which make it difficult to know that the correct
values are being used. Instead, it is more normal to carry out a routine called Calibration
during the setting up procedure for an ultrasonic inspection. In the calibration procedure the
flaw detector time-base is adjusted to give a convenient scale against a calibration sample of
known thickness and made of the same material as the work to be tested. Table 1 at the end
of this chapter lists the compression and shear wave velocities for a number of materials.

WAVELENGTH
While the particles are completing each cycle of their oscillation, the sound wave is moving
outwards in the direction of propagation at the characteristic velocity for that material. It
follows that during the time taken to complete one cycle of vibration, the sound wave will
move a certain distance depending on the velocity in that material. For a given sound
frequency, this distance is relatively small for liquids and gasses compared to that in metals,
because velocities are higher in metals. The distance travelled by the sound wave during
one cycle of vibration is called the WAVELENGTH. In general, if the maximum dimension
of a reflecting surface is equal to or greater than half a wavelength, the reflection will be
detectable. It follows that calculation of the wavelength will help in the choice of test frequency
for a specific application.

Wavelength is given the Greek symbol (lambda) and for any material and sound frequency,
wavelength can be calculated from the equation: -
f
V
=
Where =wavelength
V =Velocity
f =frequency

At ultrasonic frequencies, the wavelength of sound in metals is relatively short and so it is
usual to express the wavelength in millimetres. This is done at the start of the calculation by
changing the velocity from meters to millimetres a second by multiplying the value in M/sec by
1,000.

Example
Calculate the wavelength of a 5MHz compression wave in steel (V
c
=5960 m/sec).
5000000
1000 5960
=
mm 192 . 1 =

ACOUSTIC IMPEDANCE
Acoustic impedance of a material is the product of the materials density and velocity. At the
interface between two materials, the acoustic impedances either side of the interface will
determine what proportion of the incident sound wave will reflect and what proportion will
transmit into the second material. The symbol allocated to acoustic impedance is Z and for a
given material, V Z =

REFLECTION

Incident sound

Reflected sound
Material 2
Acoustic Impedance Z
2
Material 1
Acoustic Impedance Z
1
Interface









Transmitted sound

Fig. 2

Figure 2 shows the interface between two materials whose acoustic impedances are Z
1
and
Z
2
respectively. In the example, part of the energy is transmitted into Material 2 and part is
reflected back into Material 1. The percentage of the incident energy that is reflected can be
calculated from the equation: -

% 100
2
2 1
2 1

=
Z Z
Z Z
RE
Where: -
RE is the reflected energy
Z
1
& Z
2
are the acoustic impedances

Example 1
Calculate the percentage of the incident energy that would be reflected at a steel to water
interface given that Z
steel
=46.7 and Z
water
=1.48.
% 100
48 . 1 7 . 46
48 . 1 7 . 46
2

= RE
% 100
18 . 48
22 . 45
2

= RE
% 88 = RE
Note that the remaining 12% is transmitted into the water.
If the example had been given as a water to steel interface, the second line of the calculation
would have shown a negative value inside the brackets. However, the square term outside
the bracket would restore the answer to a positive value and the answer would have been the
same 88% reflected, this time in the water, and 12% would have been transmitted into the
steel.
When the interface is between two solids, as in the case of a brazed joint between two pieces
of steel, the reflected energy is much smaller, most of the energy passing across the braze
and into the second steel layer. There are also examples of two very different materials that
have the same acoustic impedance such as Ro-cee rubber and water. Sound travelling
through water and then encountering this particular rubber compound will carry on through
the rubber as if the interface did not exist. Table 1 shows the acoustic impedance for a
number of materials.

COUPLANT
Acoustic impedances for metals tend to have high values whereas those for gasses are low.
From the above example it is clear that at a solid to gas interface, the proportion of energy
reflected is going to be very high and the proportion transmitted will be very low. That is useful
because it means that a discontinuity such as a crack or a void in a metal object will reflect
almost all the sound back to the test surface. However, it is also a nuisance because it means
that air between the ultrasonic probe and the test surface will prevent the sound from entering
the component. A couplant is a liquid or paste used between the probe and the test surface
to try to match the acoustic impedance of the probe to that of the test material. It is not a very
efficient process because the best couplants, for example glycerine, only allows about 15% of
the sound to enter the component, and only the same proportion of any energy coming back
to the test surface can enter the probe to give an echo. At best, then, only a little over two
percent of the energy generated at the probe ever gets back to the display.
There are specially formulated couplants for use in flaw detection as well as water, oils,
greases, glycerine and pastes such as wallpaper paste. The most important considerations
when choosing a couplant are firstly that it is not hazardous to the individual and secondly
that it will not adversely affect the component.

References: -
Ultrasonic Flaw Detection for Technicians - Third Edition, June 2004 by J. C. Drury
Ultrasonic Flaw Detection in Metals Banks Oldfield & Rawding ILIFFE 1962
Material Velocity (C) Velocity (S) Density Acoustic Impedance
Units m/s x 10
3
m/s x 10
3
kg/m
3
Air 0.33 - 1.2 0.0004
Aluminium 6.40 3.13 2700 17.3
Barium Titanate A 5.26 - 5700 30.0
Barium Titanate B 5.53 - 5700 31.5
Beryllium 12.89 8.88 1800 23.2
Brass 4.37 2.10 8450 37.0
Cast Iron 3.5 to 5.6 2.2 to 3.2 7200 25 to 40
Copper 4.76 2.33 8930 42.5
Glass (Crown) 5.66 3.42 3000 17.4
Gold 3.24 1.20 19300 63.0
Iron 5.96 3.22 7850 46.8
Lead 2.40 0.79 11300 27.2
Lead Niobate 2.76 - 5800 16.0
Lead Zirconate Titanate A 3.00 - 7600 22.8
Lead Zirconate Titanate B 3.00 - 7500 22.5
Lithium Sulphate 5.45 - 2060 11.2
Magnesium 5.74 3.08 1720 9.90
Mercury 1.45 - 13550 19.6
Molybdenum 6.25 3.35 10200 63.7
Nickel 5.48 2.99 8850 48.5
Oil 1.44 - 900 1.3
Perspex 2.68 1.32 1200 3.2
Platinum 3.96 1.67 21400 85.0
Polystyrene 2.35 1.12 1060 2.5
Steel (Mild) 5.96 3.24 7850 46.7
Steel (Stainless) 5.74 3.13 7800 44.8
Silver 3.70 1.70 10500 36.9
Tin 3.38 1.61 7300 24.7
Titanium 5.99 3.12 4500 27.0
Tungsten 5.17 2.88 19300 100.0
Tungsten-Araldite 2.06 - 10500 21.65
Tungsten-Carbide 6.65 3.98 10000 to 15000 66.5 to 98.5
Uranium 3.37 2.02 18700 63
Water 1.48 - 1000 1.48
Zinc 4.17 2.48 7100 29.6
Zirconium 4.65 2.30 6400 29.8

Table 1

BACK TO BASICS ULTRASONICS

3 REFRACTION & MODE CONVERSION
Previous diagrams have shown the incident sound as if it were a single ray of energy, but of course it
is really a beam that has some width, rather like a torch beam. If the incident beam is directed at an
interface between water and steel at an angle other than normal, the angle taken up by the
transmitted beam in the steel will be greater than the incident angle in water. The advancing wave
front in a sound beam can be defined as the plane in which all the oscillating particles are in phase,
or at the same position in their oscillating cycle. The bottom edge of the beam shown in figure 1
arrives at the interface first and immediately takes up the faster velocity of the steel. As the rest of the
wave front reaches the interface, so the transmitted beam gradually takes up steel velocity. By the
time that the top edge of the beam enters the steel, the sound from the bottom edge has already
travelled four times further than it would have in water. J oining up the in phase points on the wave
front at the instant the top edge enters the steel shows the wave front advancing at a new angle. The
beam of sound is said to have undergone Refraction as it crossed the interface and the new angle is
called the angle of Refraction.

Transmitted
Beam
Water
Steel
Incident
Beam
Refraction takes place in
the time taken for the beam
to cross the interface









Interface









Normal



Fig. 1

The refraction occurs because of the difference in velocity on either side of the interface and the
proportions of energy reflected in the water and transmitted into the steel remain the same as it would
be for normal incidence. Figure 2 shows the incident, reflected and refracted angles. These angles
are always measured from the Normal to the interface. In the diagram, i

is the angle of incidence, r
o
is the angle of reflection and R is the angle of refraction.



Normal













i r
R
Interface
Medium 1


Medium 2
Incident
compression
wave
Reflected
compression
wave
Refracted
compression
wave






Fig. 2

The angles and velocities are related and the relationship is expressed in Snells Law such that: -

2 1 1
V
SinR
V
Sinr
V
Sini
= =
Where: -
i =Angle of Incidence
r

=Angle of Reflection
R=Angle of Refraction
V
1
=Velocity in Medium 1
V
2
=Velocity in Medium 2






MODE CONVERSION
If Medium1 is a liquid and Medium 2 a solid, some of the energy in the solid will change to the Shear
Wave mode. This change is known as Mode Conversion. For small angles of incidence the
proportion of energy changing to shear wave mode is small and can be ignored. However as the
angle of incidence increases the proportion increases and the shear wave becomes significant so that
there can be two types of wave in medium 2 at the same time, both of which can reflect from surfaces
within the object. Since they both travel at different speeds, and Snells Law tells us that they will
refract in different directions, the results can be very confusing. This was a restricting factor in
ultrasonics until Sproule developed the first Shear wave angle probes in 1947. Until then it was
unsafe to rely on angles of refraction greater than about 10
0
since echoes from the compression wave
could not be discriminated from the shear wave reflections. Because of this ambiguity, ultrasonics
tended to be restricted to the detection of discontinuities with surfaces parallel to the scanning surface
such as laminations and cavities. Attempts to detect, for example, weld defects such as lack of
sidewall fusion and root cracks by angling the beam were not reliable.

Sproule realised that the compression wave refracted angle would always be about double the shear
wave refracted angle because the shear wave velocity is about half the compression velocity.
Therefore if the angle of incidence were to be increased progressively there would be a critical angle
of incidence at which the compression wave would refract through 90
0
. Any increase in angle of
incidence beyond this critical angle would leave only a shear wave in medium 2 and the compression
wave would undergo total internal reflection in Medium1. With only a shear wave in medium 2
travelling at a known velocity and at a known angle, the field was open for many new applications of
ultrasonics. The critical angle at which the compression wave is refracted through 90
0
is called the
first critical angle. For a water to steel interface the first critical angle is about 15
0
and for a Perspex
to steel interface the angle is about 28
0
. At these critical angles, the remaining shear wave is at an
angle of refraction just over 30
0
. Increasing the angle of incidence above the first critical angle causes
the compression wave to be totally reflected in medium 1 and the shear wave refracted angle to
increase so that transducers can be produced at a suitable angle to detect particular defect
propagation directions.

Eventually a second critical angle of incidence will be reached at which the shear wave will be
refracted through 90
0
. The shear wave at this second critical angle will again mode convert, this time
to become a Surface (Rayliegh) wave. This new wave travels at 90% of the shear wave velocity, only
penetrates to a depth of about one wavelength, will follow the surface contour of the object and will
only reflect at an abrupt change in surface direction such as a corner or a crack. If the angle of
incidence is increased beyond the second critical angle, no sound will be transmitted into medium 2.
Ultrasonic transducers having refracted angles between 0
0
and 10
0
are likely to be compression wave
probes and those with refracted angles between 35
0
and 80
0
will be shear wave probes. Surface
wave probes have a refracted angle of 90
0
. Between 10
0
and 35
0
, and 80
0
to 90
0
it would be possible
to have two simultaneous modes existing in Medium 2 and so it is unusual to find transducers in
these two ranges exceptions to this rule will be discussed in a later chapter.
0
10
20
30
0 30 60 9
Refracted angle Sreel
I
n
c
i
d
e
n
t

a
n
g
l
e

W
a
t
e
0
r

Shear
Compression
Fig. 3
0
20
40
60
0 30 60
Refracted angle Steel
I
n
c
i
d
e
n
t

a
n
g
l
e

P
e
r
s
p
e
x
90
Shear
Compression
Fig. 4

Figures 3 and 4 show the relationship between the incident angle and refracted angle for water to
steel and Perspex to steel interfaces. The graphs show that the second critical angle for water to steel
is about 28
0
and for Perspex to steel about 58
0
. These values would be different if medium 2 were to
be aluminium or some other solid than steel.


Example 1

An incident compression wave in water meets a steel interface at an incident angle of 19
0
, calculate
the shear wave refracted angle in the steel given that the compression wave velocity in water as
1480m/s and the shear wave velocity in steel as 3240m/s.

From Snells Law

2 1
V
SinR
V
Sini
= Therefore: -

1
2
V
xSini V
SinR =

1480
3256 0 3240 . x
SinR =

7128 0. SinR =

R = 45.46
o


From a practical point of view it is more usual to know the refracted angle needed in the test material
in order to detect a particular discontinuity, and so the calculation would be to find the necessary
angle of incidence, in water for immersion testing, or in Perspex for contact scanning. Example 2
shows this version of the application of Snells Law.

Example 2
Calculate the angle of incidence required in Perspex in order to produce 45
o
Shear wave in steel
given that the compression wave velocity in Perspex is 2680m/s and the shear wave velocity in steel
is 3240m/s.
From Snells Law

2 1
V
SinR
V
Sini
= Therefore: -

3240
7071 0 2680 . x
Sini =

5849 0. Sini =

Incident angle =35.8
o


REFLECTIVE MODE CONVERSION

Mode conversion also takes place when an ultrasonic beam reflects at internal surfaces in solids
whether these are boundary surfaces, machined features, or discontinuities. The relationship
between incident angle of a given beam and the relative amplitude of the reflected and mode
conversion beams for steel is shown in the following graphs. They allow an assessment to be made
of the potential confusion in any given situation and can be used to determine an alternative test
angle to be chosen to avoid the problem.

90
0
60
0
30
0
0
0
40
0
30
0
20
0
10
0
50
0


S
C
0.4
0.6
0.8
1.0
Relative Amplitude
a) Incident Compression Wave










Steel

Air

S
C C










Fig. 5
0.2






A compression wave incident on a steel to air interface will reflect as a compression wave together
with a mode converted shear wave. For example figure 5 shows that at an incident angle () of 30
o

we find that is around 15
o
but the relative amplitudes of the shear wave and compression wave are
90% and 70% respectively. Both will give strong signals if they reach the receiver. In the extreme
case, where is around 60
o
and around 30
o
we find that the relative shear wave amplitude is 90%
but the reflected compression wave amplitude has fallen to only about 10%. For greater angles of
incidence than 60
o
, the shear wave rapidly decreases in amplitude and the compression wave
recovers. Clearly we need to take care in our interpretation of signals if we see that a compression
wave in steel is likely to meet a known reflecting surface in that part of the graph where the shear
wave amplitude is significant.




50
0
40
0
30
0
20
0
10
0
1.0
Relative Amplitude
0.8
0.6
0.4
0.2
0
0

C
S
b) Incident Shear Wave



90
0





Steel

Air

C

S
S





60
0







30
0
Fig. 6









A shear wave will reflect as a shear wave together with a mode converted compression wave. Using
the graph in figure 6 we can see that the most severe case is when the incident shear wave meets a
steel to air interface at about 30
o
. The reflected shear wave amplitude is very low and the mode-
converted compression wave is very strong and almost perpendicular to the test surface.

If the incident shear wave grazes a surface, in other words the incident angle is around 90
o
, there will
be a mode conversion to Rayleigh wave. This can happen when a shear wave grazes the bore of a
machined hole in the specimen. In that case the Rayleigh wave will follow the bore surface and will
reflect if it encounters a sharp changes to the bore such as a keyway. If you are not aware of the
possibility, you may assume that there is a discontinuity in a false position. An example is shown in
figure 7 below. Of course, when you are aware of this phenomenon, you can use it to advantage to
find a crack in an otherwise blind position around a hole or radius.









Assumed reflection point


Rayleigh wave reflects from keyway


Fig. 7







References: -

Ultrasonic Flaw Detection for Technicians - Third Edition, June 2004 by J. C. Drury

Ultrasonic Flaw Detection in Metals Banks Oldfield & Rawding ILIFFE 1962
BACK TO BASICS ULTRASONICS

Part 4. TRANSDUCERS FOR GENERATING AND DETECTING SOUND WAVES

SWINGING THE LEAD

There is an amusing story about a nautical gentleman at the time when wooden ships were being
superseded by iron ones. This sailor thought up a new way to determine the depth of water under the
hull to replace the old lead weight on a rope method. He decided that it should be possible, with a
large hammer and a stopwatch, to bang on the iron bottom of the ship with the hammer and time the
return echo from the seabed with the stopwatch. The measured time could be used to calculate depth
using the speed of sound in water. Fired with enthusiasm, he gathered together a number of marine
dignitaries in the bilges of his ship, passed a large sledgehammer to a muscular boatswain, took out
his stopwatch and ordered the swain to wallop the floor. This he did with such vigour that the hull
boomed for ten minutes and the assembled observers were deafened for a month! Nobody heard an
echo.

There are parallels with ultrasonic flaw detection in the story; we need our sound pulses to be loud
enough to penetrate to the depth of the anticipated flaw and we need the duration of the pulse to be
short so that it does not mask any returning echoes. We also need the sound frequency of the pulse
to produce a wavelength short enough to detect the smallest reflector that must be detected to ensure
safety. In this chapter we will discuss various ways of generating and detecting suitable pulses and
some of the limitations we face in terms of penetration and flaw sensitivity.

ULTRASONIC TRANSDUCERS
A transducer is a device that will change one form of energy into another. Ultrasonic transducers
change electrical energy into mechanical energy (sound waves) or vice versa. There are several
methods used to generate and detect ultrasonic pulses in modern flaw detection and the most
common of these makes use of the Piezo Electric effect found in certain materials. Other methods,
such as the Electro Magnetic Acoustic Transducer (EMAT) and Laser technology will also be
described.

PIEZO ELECTRIC TRANSDUCERS
In 1880 the Curie brothers discovered that slices cut in a particular way from certain crystal materials
would generate an electrical potential across the faces of the slice when distorted by a mechanical
force. They called this phenomenon Piezoelectricity from the Greek words for Pressure and
Electricity. A year or so later Lippman reported that the reverse was true; that a voltage applied
across the slice would produce a mechanical distortion. Quartz was the prime example of a
piezoelectric material, but Rochelle salts and Tourmaline crystals also displayed the same effect.

For the first thirty years of ultrasonic flaw detection from Sokolov in 1929, until the end of the nineteen
fifties, quartz was the most common transducer material. Appropriate slices were cut from a single
crystal. Later new polycrystalline materials were developed that had lower electrical impedance
(resistance to high frequencies) and gave better ultrasonic performance, as much as 60 to 70 percent
more efficient than quartz. These materials have to be Polarised. During polarization the individual
crystals align themselves in the same direction so that their combined effect is coherent. The
polarisation process involves heating the discs in an oil bath to a critical temperature called the Curie
Temperature, applying a strong electrostatic field across the disc and then allowing the temperature
to cool slowly. Figure 1 illustrates the polarising process.



Electrostatic field
Heated oil
- - - - - - - - - - - - -
++++++++++++++


Polycrystalline disc


Fig. 1


The Curie temperature differs for each of the common materials used in ultrasonics, so that the oil
bath will need to be heated to a suitable temperature for the material in use. For Barium Titanate the
Curie temperature is around 120
o
C whereas for various grades of Lead Zirconate Titanate (PZT) the
temperature is from 190
o
to 350
o
C and for Lead Metaniobate (PMN) it is about 400
o
C. If the material
is subsequently heated to a temperature near to the Curie temperature, the disc will depolarise and
lose its piezoelectric properties. It follows that care needs to be taken to avoid depolarisation when
testing hot materials and this will sometimes influence the choice of transducer material.

MODE OF VIBRATION
Whether the transducer disc is made from a naturally occurring piezoelectric crystal, or one of the
polarised polycrystalline materials, we usually refer to the disc as the crystal when talking about
probe construction. The crystal disc or plate may be round or rectangular and for some applications
may be curved plates or concave discs to focus the sound. The way in which the plate vibrates when
stimulated by an electrical pulse depends upon the cut, in the case of quartz, or the direction of
polarisation in the case of polycrystalline materials. Figure 2 represents a typical quartz crystal
showing the three axes defined by crystallographers, and two plates cut from a crystal, one an X-cut
plate and the other a Y-cut plate.




Z
X
Y
X
X-cut crystal
X
X
Y
Y
Y Z
Y-cut crystal


Fig. 2


An X-cut plate is taken from the quartz crystal so that the X-axis is perpendicular to the plate and the
Y-cut plate has the Y-axis perpendicular to the plate. If a voltage is applied across the faces of these
plates, an X-cut crystal will distort in the thickness mode whereas a Y-cut crystal will distort in shear
mode. Figure 3 illustrates the changes in shape when an alternating voltage is applied to an X-cut
crystal and Figure 4 shows the shape changes for a Y-cut crystal. The same two modes of vibration
can be obtained using the polycrystalline materials by polarising across the faces of the plate
(equivalent to X-cut), or parallel to the faces of the plate (equivalent to Y-cut)















The X-cut crystal is the one most commonly used in ultrasonic flaw detection, it can generate and
detect compression waves, and can therefore transmit sound through the liquid couplant we use.
Since shear waves cannot exist in liquids or gases, the only way in which a Y-cut crystal could be
used to generate shear waves in a metal object would be to use a solid couplant or high viscosity
Fig. 4
Fig. 3
liquid such as honey; in other words we would need to almost glue the crystal in position. This is done
in a few very special applications.

METHOD OF PULSING AND FREQUENCY
When we generate a short pulse of sound with our crystal, we dont drive the crystal with an
alternating voltage of suitable frequency; instead, we pluck the crystal with a short sharp electrical
shock and allow the crystal to ring at its natural resonant frequency. This is rather like plucking a
guitar string that also vibrates at its natural frequency. In the case of the piezoelectric plate, the
crystal stretches as the voltage is applied and only produces sound when the voltage is rapidly cut
off. To increase the amplitude (loudness) of the ultrasound we increase the peak voltage (pulse
energy) applied to the crystal. The frequency of our ultrasonic transducer is determined by the
thickness of the crystal. As the crystal is made thinner, so the resonant frequency increases. Quartz
crystals are split in the appropriate and then lapped to the correct thickness for the required
frequency. The polycrystalline materials are made as slurry that is moulded and compacted under
pressure and then sliced and lapped to the required thickness before polarising.

The required thickness for a given frequency can be calculated from the frequency-thickness
constant for the crystal material to be used. Since this depends on the velocity of a compression
wave in that material it can be seen that the thickness for a given frequency will not be the same for
PZT and quartz, for example. The frequency-thickness constant is defined mathematically as: -
2
V
fxt =

Where f =the desired frequency
t =the crystal thickness
V =the compression wave
velocity in the crystal

Example
Calculate the required thickness of a PZT crystal to produce a resonant frequency of 5MHz given that
the compression wave velocity for PZT is 3000M/s.
2
V
fxt =
xf
V
t
2
=
mm x
x
t 1000
5000000 2
3000
=
t =0.3mm


CONTROL OF PULSE LENGTH
In ultrasonic flaw detection we measure the time taken for each echo to arrive at the receiver after
entering the scanning surface of the object. If we know the velocity of sound in the material we can
determine the distance travelled by the sound wave. Suppose that a crack has grown from a bolthole
in the object as in figure 5; some of the sound will reflect from the top of the bolthole, and a little while
later, some will reflect from the crack. The arrival of the two echoes at the receiver will be separated
by a short interval of time (T
2
T
1
). If the ringing time of the crystal (pulse length) is longer than this
interval of time, then we may not be able to distinguish the crack from the top of the bolthole we
may miss the crack. We say that we have not resolved the two echoes or that the resolution is
poor. In order to improve resolution we need to ensure that the pulse length is as short as possible.




Bolthole Crack
T
2
T
1

T
1
T
2
T
1
T
2
Short pulse
Long pulse














Fig. 5


In ultrasonics we shorten the pulse duration by applying a weight to the back of the crystal known as
the damping or backing slug. The damping slug is often made of a mixture of tungsten powder in
an epoxy resin. The amount of damping applied to the crystal will govern the resolution of the probe.
A short pulse probe will have only one or two cycles whereas a longer pulse probe may have from
three to five cycles. An undamped crystal may have twelve or more cycles in the pulse. For a given
number of cycles in a pulse, the duration or space occupied by the pulse will depend on the
wavelength, which in turn depends on the probe frequency and the velocity of sound in the material
being inspected. Thus: -

Pulse length =Number of cycles in the pulse multiplied by the wavelength.

It is obvious that one way to improve resolution would be to increase the test frequency, however,
since the penetration of sound into the object decreases as the frequency increases, this is not
always possible. Choosing a suitable test frequency is often a compromise between resolution
penetration and flaw sensitivity and sometimes ultrasonics will not be able to detect a particular
discontinuity at the critical depth. While resolution is an important consideration in many applications,
it is not always the case and sometimes a longer pulse is preferable. For example, in the examination
of a long shaft such as a railway axle, the screen on the flaw detector may only be 75mm wide and
the display may represent the length of the shaft, say 2.5m long. A short pulse of 2 cycles will occupy
such a small part of the screen that it is too faint to see and it would be better to use a longer, more
visible pulse

PIEZO-COMPOSITE TRANSDUCERS
In a more recent development of the piezoelectric transducer, the active plate in the test probe is
made by slicing piezoelectric crystals into small squares and assembling them into a matrix separated
with an epoxy or a rubber compound as shown in figure 6. The main advantages of this type of
construction are firstly, lower acoustic impedance allowing better acoustic matching. This is an
advantage when testing castings and stainless steel. Secondly, resolution they tend to provide short
pulses without additional damping, allowing the probes to have a low profile.



Plan view



Epoxy






Crystals





Side view




Fig. 6

POLYVINYLIDENE FLUORIDE (PVDF) TRANSDUCERS

PVDF was also found to display Piezo electric characteristics and has been used in ultrasonic flaw
detection. These thin plastic films have advantages and limitations compared with conventional
crystals. On the plus side, they can be easily shaped to focus sound, they produce very short pulses
and they give good transmission into water because the acoustic impedance is similar to water.
Against these advantages, the films are fragile and cannot be used in contact scanning, and the
power output is relatively low compared with ceramic crystals. The main application is in high
resolution immersion testing.

ELECTROMAGNETIC ACOUSTIC (EMAT) TRANSDUCERS
EMAT transducers provide a non-contact alternative to piezoelectric transducers. Sound waves are
generated in the surface of a conductive test object by an electrical pulse applied to a flat coil that is
positioned between a strong magnet and the test piece. The interaction between the magnetic field
generated in the coil by the electrical pulse and the fixed magnetic field of the magnet causes a rapid
shock deformation at the surface of the test piece and an ultrasonic wave travels through the metal
object. The EMAT probe needs to be close to the test surface, but does not need to touch it.
Returning echoes arriving at the scanning surface cause the surface to vibrate in the magnetic field.
This generates eddy currents in the test surface and the coil detects the eddy currents.

EMAT probes can be used with an air gap when testing hot surfaces and on surfaces coated with
non-conducting material such as rubber, paint and fibreglass because the sound wave does not have
to travel through the gap material. The probes can be configured to generate horizontally polarised
shear waves directly into the test object. This is an advantage when testing austenitic welds, castings
and other materials with dendritic grain structure because the shear wave does not mode convert
when it meets a reflecting surface that is parallel to the direction of polarisation. Because shear
waves travel at roughly half the velocity of compression waves and have shorter wavelengths, it is
possible to obtain better near surface resolution and this can be an advantage when testing thin
materials.

However, there are some disadvantages with EMAT probes, they are relatively large and inefficient
compared with conventional probes and they cannot be used on non-conducting test objects unless a
conducting coating is applied.

LASER TRANSDUCERS
Another non-contact method of generating ultrasound uses laser technology. A short burst of a laser
beam on the surface of the test object causes a thermal shock with rapid local expansion of the
surface. The sudden distortion of the surface causes an ultrasonic pulse to travel through the test
object. The returning echo distorts the test surface and this distortion can be detected by a separate
laser interferometer without a couplant, or can be detected with a conventional piezoelectric crystal
and couplant. The gap between the transducer and test surface can be greater than is possible with
EMAT probes and can be as much as 250mm (10 inches). Typical applications include the inspection
of composite materials in the aircraft industry.

References: -
Ultrasonic Flaw Detection for Technicians - Third Edition, June 2004 by J. C. Drury
Ultrasonic Flaw Detection in Metals Banks Oldfield & Rawding ILIFFE 1962

BACK TO BASICS ULTRASONICS

Part 5. PROBE CONSTRUCTION

COMPRESSION WAVE PROBES

Standard compression wave probes can be for contact scanning or for immersion testing. The contact
scanning probes are either single crystal or twin crystal (dual) in construction. The construction of a
typical single crystal contact probe is shown in the diagram in figure 1.


Wear face
Damping Slug
Housing
Cap
Co-axial
connector










Crystal




Fig. .1


The thickness of the crystal determines the operating frequency as we described in an earlier article
and the faces of the crystal are coated in silver to make electrical contact.
The damping slug is cast onto the rear of the crystal and bonds to it as the epoxy sets. The amount of
damping used determines the pulse length. A fine wire is soldered to the back of the crystal, using a
solder that melts at low temperature, before adding the damping slug.
The wear face is glued to the front face of the crystal to protect it during contact scanning. The
thickness of the wear face is important. It is made to be one quarter of the wavelength at the test
frequency for the velocity of sound in the wear face material. This thickness gives maximum
transmission of sound out of the probe into the test sample. Some wear faces are made from shim
steel, others from a hardwearing ceramic material. The steel wear faces can be used to earth the
front face of the crystal to the probe housing and are less fragile if you drop the probe, but are
inclined to stretch and disbond from the crystal with use. If a non-conductive wear face is used, an
alternative earthing method must be used.
The wear face, crystal and damping slug assembly are then fitted into the housing, the other end of
the centre wire is soldered to the centre terminal of the connector and the cap and connector fitted to
the housing. Figure 2 is a photograph of a typical single crystal compression wave probe.











Fig. .2

Twin crystal, or dual probes are used to eliminate the dead zone occupied by the transmission
pulse with a single crystal probe. In this type of probe one crystal acts as a transmitter, the other as a
receiver and the amplifier is isolated from the transmitting crystal. The two crystals are mounted on
acrylic or polystyrene wedges these components are illustrated in figure 3. An acoustic barrier,
usually made of cork, is fitted between the wedges and crystals to prevent cross talk between the
transmitter and receiver. Figure 4 shows a typical twin crystal probe.




Tx Crystal



Acoustic barrier

Wedges
Rx Crystal



Fig. 3















Fig. 4




Immersion probes are similar in construction to that shown in figure 1 except that it is not necessary
to fit a wear plate and so the silvered face of the crystal is usually visible. Probes can be focussed
and this is achieved by fitting a plastic or epoxy lens to the front of the crystal, or by making a curved
sectioned crystal. Figure 5 shows a 20MHz immersion probe with a small diameter spherically
focussed crystal.


Figure 5

The lens or curvature can also be cylindrical as illustrated in figure 6. The cylindrical version is often
referred to as a Paintbrush probe because it allows a wide scan.

Focussed beam
Curved crystal















Fig. 6


Focussing can also be achieved using a technique called Phased Array, although not with
conventional ultrasonic sets. The phased array probe contains a number of small crystals and the
pulsing circuit is designed to be able to apply a pulse to all crystals simultaneously to produce a
conventional zero degree compression wave, or to pulse each crystal separately with a small time
interval between each. In the diagram shown in figure 7, the outer elements are triggered first and a
progressive delay is used to pulse inner elements, the centre crystals being the last to be triggered.
The result is that the ultrasonic wavefront reinforces in the curved way shown in the diagram to focus
at a region determined by the delay intervals. By changing these intervals, the focal length can be
changed. The principles of constructive and destructive reinforcement will be dealt with later in a later
article.













Fig. 7

Single crystal Delay line probes are sometimes used in contact scanning to reduce the Dead Zone
below the beam entry surface occupied by the transmission pulse and probe noise. The delay line is
usually Perspex or a similar material and provides a stand off just like the water path in immersion
testing. The length of the delay line must be sufficient to allow one or more backwall echoes in the
specimen depending on the application. Figure 8 is an example of a delay line probe.


Fig. 8

SHEAR WAVE PROBES
Since shear waves cannot travel through liquids or gases, angled beam probes use compression
waves in the incidence wedge in contact probes. The incident angle will be an angle between the first
and second critical angles (described in the third article in this series) so that we only have the mode
converted shear wave in the test material. Figure 9 is a sketch of the typical arrangement.

Internal Reflection
Incident wedge
Shear wave

Compression Wave Element



Test surface





Fig. 9
We not only get a mode converted, refracted shear wave in the test piece, but we also have a
reflected compression wave in the wedge. If this internal reflection manages to get back to the crystal
face as it bounces around the wedge, we would have a standing echo that would be confusing.
Several methods of avoiding this problem have been used over the years. The earliest probes used a
long Perspex shaped Cusp so that the reflection would be absorbed before it could return to the
crystal. The Cusp made a rather unwieldy probe and the next design used V shaped grooves in the
front and top surfaces of the incident wedge to scatter the internal reflections. Some had a plastic
material moulded onto these grooves to further damp the reflection. In the latest, most compact,
designs the wedge is surrounded by a material that has a good acoustic match to Perspex, but a
much higher absorption of sound. The internal reflections are transmitted easily into this layer and
then absorbed. Figure 10 shows examples of the three designs and illustrates their relative size.













Fig. 10

Figure 11 is a photograph of a sectioned shear wave probe, showing the crystal, incidence wedge
and the blocking medium for the internal reflections.


Coaxial connector
Housing
Crystal
Blocking medium
Incidence wedge
Fig. 11

Phased Array transducers, such as the one already discussed (figure 7), are also used to generate
angled shear waves in the test piece. These transducers have the advantage that the phase delay
between the crystal elements can be varied to give different angles of refraction. The delays can be
swept through a range of values to give a shear wave beam that sweeps through a desired range of
shear wave angles rather as a Radar scanner sweeps the skies.

In the fourth article, we said that EMAT probes could generate compression or shear waves, but that
shear waves were often used because they can be directed perpendicular to the test surface (that is
a 0
0
probe). That has advantages in resolution, because the wavelength for a shear wave is about
half the wavelength for a compression wave and because the velocity of the shear wave is about half
that of the compression wave, we are able to measure thinner sections than we can with conventional
0
0
probes of the same frequency. The EMAT probe shown in figure 12 is a radially polarised shear
wave probe operating broadband between 1-10MHz, with a centre frequency of about 5MHz.













Courtesy of Ultrasonics Group,
Dept of Physics, University of Warwick

Fig. 12




Q FACTOR AND BANDWIDTH
Up to this point we may have gained the impression that our transducer produces a pure note at the
calculated frequency, but this is not true. In fact the sound wave produced contains a band of
frequencies related to the thickness of the crystal, its diameter or length and width plus the effects of
the damping medium. In addition the electrical characteristics of the transducer and associated
circuits affects the overall spectrum of frequencies. We refer to this spectrum as the Bandwidth of
the probe. In a well-designed probe, the centre of this band should be the desired probe frequency
and the lower and upper limits are usually defined as the frequencies at which the amplitude is
reduced by a given factor. Some people use 30% (-3dB) and others 50% (-6dB) as the factor we will
use 50% in the following examples. Figure 13 illustrates the bandwidth of a 5MHz probe in which the
6dB bandwidth is equal to the centre frequency, in other words, from 2.5MHz to 7.5MHz.
0
20
40
60
80
100
0 2.5 5 7.5 10
Frequency (MHz)
A
m
p
l
i
t
u
d
e

(
%
)
-6dB points
Peak
Frequency
Fig. 13


Figure 14 shows the bandwidth for another 5MHz probe, but this time the bandwidth is only from 3.75
to 6.25MHz.

0
20
40
60
80
100
0 1 2 3 4 5 6 7 8 9 10
Frequency (MHz)
A
m
p
l
i
t
u
d
e

(
%
)

-6dB points
Fig. 14

The probe shown in figure 13 can be described as having a broad bandwidth whereas the probe in
figure 14 has a narrower bandwidth. In practice, short pulse probes have a broad bandwidth and long
pulse probes are narrow bandwidth. For a given crystal size, material and frequency damping not
only reduces pulse length, but also reduces pulse amplitude, so the narrower bandwidth probes will
have longer pulses but more amplitude in the pulse therefore giving deeper penetration.

Another way of expressing bandwidth that is also common in other branches of electronics is the
Quality Factor or Q of the probe and is defined by the formula: -
1 2
0
f f
f
Q

=
Where f
0
=the centre frequency
f
1
=the upper 6dB frequency
f
2
=the lower 6dB frequency



Example
Calculate the Q factor for the probe illustrated in figure 13.

1 2
0
f f
f
Q

=
5 2 5 7
5
. .
Q

=
1 = Q

Example 9
Calculate the Q factor for the probe illustrated in figure 14.
1 2
0
f f
f
Q

=
75 3 25 6
5
. .
Q

=
5 2
5
.
Q =
2 = Q
Undamped crystals can have a Q factor as high as 20,000 but for ultrasonic flaw detection the Q
factor is normally in the range 1 to 10.

References: -
Ultrasonic Flaw Detection for Technicians - Third Edition, June 2004 by J. C. Drury
Ultrasonic Flaw Detection in Metals Banks Oldfield & Rawding ILIFFE 1962



BACK TO BASICS ULTRASONICS

Part 6. PULSE-ECHO FLAW DETECTOR

The ultrasonic flaw detector is required to provide the voltage pulse to activate the probe
crystal, to amplify received signals from the probe and to display those signals so that the
relative time of arrival and amplitudes of the signal train can be viewed and interpreted. In
order to display the very short intervals of time involved in testing metals, the early pulse echo
systems used a cathode ray tube (CRT) as the display module. More recently, equipment
manufacturers have turned to digital technology and used LCD panels for the display. The
result has been the manufacture of much smaller and lighter ultrasonic equipment. Ultrasonic
sets in the early 1960s used thermionic valves (vacuum tubes) and weighed 25 to 30 Kg (50
60 lbs). From the late 1960s, transistor technology and smaller CRTs meant that the flaw
detectors became smaller and lighter weighing between 5 and 10 Kg (10 20 lbs). In the new
millennium, the weight has come down to around 3 Kg. Figures 1 to 3 show the progression.










Circa 1960
Fig. 1









Circa 1980
Fig. 2
Circa 2000
Fig. 3


Figure 4 is a block diagram of a typical analogue flaw detector showing the main components
and the controls associated with each component.


Depth range coarse
Depth range fine
Delay
Single/Dual
Switch
Tx Rx
Timebase Generator
Pulse Generator
Frequency
Gain
Reject
Rectification
Receiver
Am

plifier
Clock
CRT Display
Focus
Brightness
X-shift
Y-shift
PRF

























Fig 4

The clock or timer is the heart of the flaw detector. It feeds an electrical pulse to the Pulse
Generator and simultaneously to the Timebase Generator. This timer pulse causes the pulse
generator to send a short, high voltage pulse to the crystal and at the same time triggers the
timebase generator to begin to sweep the electron beam in the CRT tube from left to right
between the X plates at a constant speed.

As soon as the high voltage pulse at the transmitter crystal is cut of, the crystal starts to
vibrate and an ultrasonic pulse propagates into the test piece. While this sound pulse travels
through the material, the CRT sweep continues to track the time as it moves towards the right
hand side of the display. Reflections from internal surfaces arrive at the receiver crystal,
generate a voltage in the crystal and this voltage is amplified and passed to the Y plates
where it causes a vertical deflection of the electron beam proportional to the amplitude of the
received signal.

When the electron beam reaches the extreme right hand side of the CRT it flies back to the
left hand side and waits for the next trigger pulse from the clock. This whole sequence of
events takes place so quickly that we wouldnt be able to see the trace. The clock repeats the
sequence many times a second and the result is a flicker free trace that increases in
brightness the more times we repeat the process each second. The number of trigger pulses
per second is known as the Pulse Repetition Frequency (PRF), or Pulse Repetition
Rate (PRR). It is important that we allow enough time between pulses for all the multiple
echoes within the specimen to die away or we will the tail end of these echoes showing as
Ghosts at confusing positions on the timebase. For this reason the PRF is controlled by the
Depth Range Coarse control in the timebase generator circuit. However, some flaw detectors
have an additional manual control that the operator can use. Ghost echoes are most likely to
be encountered when testing fine-grained light alloy forgings that have very low attenuation of
sound.

The voltages developed in the receiver crystal are very small and need to be amplified. The
Amplifier circuit needs to be tuned to accept the frequency of the ultrasonic pulse and this
can be by way of switched bands for example, 1-3Mhz, 3-7MHz, 7-10MHz & 10-15MHz, or it
could be a wideband amplifier with the range 1-15MHz. If the former, the set will have a
Frequency selector switch that should be switched to the appropriate band for the probe in
use just as you would use the tuning dial to select the desired radio programme.

The Gain or Sensitivity control allows the amplification to be increased or decreased
depending on the strength of the received signals much like the volume control on a radio.
The Gain control is usually calibrated in decibels (dB) and is sometimes called the
Attenuator. Strictly speaking, an attenuator should be calibrated such that increasing the dB
reduces the signal amplitude, but this is seldom the case over recent years. The Bel is a unit
that is commonly used in electronics to compare the ratio between two power or voltage
values and is a logarithmic unit so that large ratios can be expressed concisely. The intensity
of sound in a received pulse is a measure of the power or energy in that pulse, and that
mechanical energy is converted into electrical energy by the piezoelectric crystal. If the power
increases from P
1
to P
2
, then the gain can be expressed as: -

Bels
P
P
Log Gain
1
2
10
=

However, the Bel is too large a unit for the values we shall encounter in ultrasonics and so we
use a unit of one tenth of a Bel or decibel. The equation then becomes: -

dB
P
P
Log Gain
1
2
10
10 =

The CRT measures voltage and electrical power is proportional to the square of the voltage: -
dB
V
V
Log Gain
2
1
2
10
10

= And, removing the brackets: -


dB
V
V
Log Gain
1
2
10
20 =
The height of a signal on the CRT is proportional to the voltage applied to the Y plates and
so we can change the equation so that it is in terms of signal height: -
dB
H
H
Log Gain
1
2
10
20 =
Example 10
Calculate the gain ratio in dB between a signal that is 60% full screen height and one that is
only 30% full screen height.
dB Log Gain
30
60
20
10
=
dB Log Gain 2 20
10
=
dB . x Gain 3010 0 20 =
dB . Gain 02 6 =

When we measure depth or thickness from the timebase, we use the left hand flank of the
signal on the screen. Sometimes surface roughness, material grain size, or electronic noise
create noise signals (grass) that obscure the point where the flank meets the timebase and it
is difficult to make the correct reading. In these circumstances, we can use the Suppression
or Reject control to remove the grass, much as we use a tone control on a radio to cut out
hiss. Because this control can also cut out small relevant signals and make the gain non
linear, a warning light comes on when the control is in use.

The last feature that we need to consider in the amplifier circuit is the one that controls the
degree of rectification and smoothing of the pulse. The received signals are, of course, a few
cycles of alternating voltage. We can display these as they are Unrectified but it is not
so easy to measure amplitude directly from the screen. It is more usual to display these
signals as Rectified and smoothed signals in which the negative half cycles are inverted
and the signal envelope smoothed out. On some equipment, we may also have the choice to
only display the Positive, or Negative half cycles and this may give a sharper flank to the
signal. Figure 5 illustrates the four conditions, but unsmoothed to illustrate the principle.




Unrectified Full Positive Negative

Fig. 5
The Timebase circuit controls the sweep speed and delay functions. The sweep speed will
determines the thickness range that can be displayed on the CRT. A high sweep speed (fast
timebase) may only allow a return path from a 10mm thickness in the test piece and at the
other extreme, a low sweep speed (slow timebase) may allow a return path from 5 metres or
more. Two controls achieve the desired thickness range, the Coarse Depth or Range
control switches the range in steps (10mm, 50mm, 100mm, 500mm, 1m & 5m for example)
and the Fine Depth or Range control is a continuously variable control that allows fine
adjustment during calibration to allow for the specific material velocity. The fine depth range
control is sometimes labelled Material or Velocity.

There are times when we dont want the timebase generator to begin the sweep when the
crystal is pulsed. For instance, when we are carrying out an immersion test we want the
timebase to start when the sound enters the specimen so that the left hand end of the
timebase represents the top surface of the test piece. Another example might be when we are
testing a long shaft and we want to look in more detail at, say, the last 200mm of the shaft. In
either case, we can delay the start of the sweep with the Delay control.

The last component to consider is the display module, the CRT. The image created by the
electron beam (the trace) must be displayed so that the baseline is aligned with the graticule,
extends beyond the left and right hand ends of the graticule, is bright enough to see in the
test environment and is in focus. There are four controls for these functions, the X-shift and
Y-shift controls position the trace, the Brightness control can be adjusted for indoor or
outdoor viewing, and the Focus control sharpens the trace. On many flaw detectors, only
the focus and brightness controls are provided for operator adjustment.

Digital flaw detectors provide the same PRF, Amplifier and Timebase functions but these are
usually controlled using a combination of menu selection and so called Smart Knobs
through the controlling central processing unit (CPU). Figure 6 is a representative block
diagram for a digital instrument.

One of the real advantages of the digital instruments is the facility to store calibrations for a
number of inspection procedures and probes, to store whole traces complete with the
calibration data for each trace and to create databases to store thickness readings. Because
the instruments are based on computer technology, it is possible to connect the flaw detector
to a PC through a serial cable and download stored data, for reporting purposes.

The LCD display also has advantages over the CRT. It consumes less power than the CRT; it
can be backlit for viewing in low light conditions and at the same time is easy to see without
backlighting in daylight. In difficult conditions, the trace can be Frozen so that the operator
can move to a more comfortable position before reading the timebase.



























LCD
Display
Display
Control
Controller
Acquisition
& Digitiser
CPU
Tx Rx
Pulser/Receiver Attenuator
Rectification
& Filtration

Fig 6

Many flaw detectors, both analogue and digital, have gating circuits that allow signals to be
monitored by the instrument and the output used to trigger audible or visual alarms, or to be
connected to chart recorders or computers. The monitor gates may be displayed in one of two
ways. The timebase may be raised over the gate distance as shown in figure 7, or a separate
bar may be used as shown in figure 8.



0 1 2 3 4 5 6 7 8 9 10
Gate
0 1 2 3 4 5 6 7 8 9 10
Gate












Fig 7 Fig 8



There are four main functions controlling the gate, these are: -
Gate Start
Gate Width
Gate Level or Threshold
Gate sense (Rising or Falling Signal)

The gate start control positions the left hand edge of the gate, the first depth that you want
to start monitoring. The gate width control then allows you to set the right hand edge of the
gate, the last depth that you want to monitor. Any signal within that depth range is said to be
in the gate. You may only want signals exceeding a predetermined amplitude to trigger the
gate alarm and you do this using the gate level or threshold control. For those gates that
look like figure 6.7, you set a signal in the gate at the desired amplitude, and adjust the
threshold until the alarm just triggers. For those gates that look like figure 6.8, you simply
adjust the gate level control until the gate is at the desired screen height. For some
inspections, such as when you are using the through transmission technique, you may wish
to monitor for a decrease in signal amplitude. The gate sense can be changed using the
sense control. When falling signal has been selected, the alarm does not trigger as long as
there is a signal in the gate that exceeds the threshold level. Instead, the alarm operates as
soon as the signal drops below the gate threshold.

Some flaw detectors have more than one gate. Two gates can be used in several ways; one
can monitor backwall echo amplitude (falling signal) and one can be used to monitor part of
the timebase for discontinuities (rising signal). The two gates can be used to monitor
consecutive backwall echoes and the difference (gate 1 minus gate 2) can be output as the
thickness of the object. The menu of a digital flaw detector may allow you the choice to
monitor either signal amplitude or time of flight (depth). This is also possible with some
analogue flaw detectors by the appropriate pin selection on the output connecting lead.
Generally, the voltage range for the output signal is about 0V to 5V; this means that the
vertical or horizontal (amplitude or timebase) scales of the display will be proportional to the
output range. If monitoring and recording amplitude, for example, a full screen echo height will
output 5V and a half screen height signal will output 2.5V.

BACK TO BASICS ULTRASONICS

Part 7. THE ULTRASONIC BEAM

The beam of sound waves emerging from an ultrasonic probe is rather like the beam of light from
a torch. The beam will spread out into an elongated cone shape, and the further away you go from
the source, the weaker will be the beam. So in order to know just how this beam affects our in-
spection, we need to study the shape of the beam in detail, and to study the changes in intensity of
the beam along its axis and across the beam.

As a general principle, we have said that the beam gets weaker as we get further from the
transducer. This weakening, or decrease in intensity represents a loss of energy; we say that the
beam is attenuated as it progresses through a material. The sound beam suffers this attenuation for
the following reasons: -
ABSORPTION - of the energy due to moving the vibrating molecules
SCATTER - of sound waves reflecting from the grain boundaries
INTERFERENCE EFFECTS - close to the transducer
BEAM SPREAD - the energy spreads over a larger area with distance

The amount of energy lost through Absorption depends upon the elastic properties of the material
being tested so that steel and aluminium have less absorption than lead, or Perspex. Scatter also
depends upon the material being tested, the larger the grain size, the greater the scatter (see figure
1). Forged and rolled materials generally give less scatter than castings or forgings. Heat treatment
may reduce grain size and therefore reduce scatter, making testing easier. Faced with a material
that presents either, or both, high absorption and scatter, you have to resort to a lower test
frequency to overcome the problem. We can either say attenuation (absorption and scatter)
decreases as test frequency decreases, or penetration increases, as frequency decreases.

















Fig. 1


INTERFERENCE EFFECTS
Point Source: - If we consider a point source of sound energy, then the disturbance (sound
wave) will radiate outwards from the point in an ever increasing circle, just like the ripples on a pond
spreading out when you drop a stone into it. So sound radiates in all directions from a
point source. (figure 2).

Sound wave expanding
outwards from point source
Positive peak wave front
Negative peak wave front

Fig. 2

Finite source: - Our transducer, however, is not a point source, but a plate of
piezoelectric material of finite dimensions. In order to appreciate the way in which sound
spreads out from a finite source, and to help us understand interference effects we will
use Huyghens principle, Huyghens said that you can consider a finite source to be made
up of an infinite number of point sources. When you energise the transducer, sound will
radiate out from each of these point sources, just as it did for the stone dropping into the
pond. Figure 3 shows sound radiating from just one of these point sources and figure 4
shows sound radiating from several point sources.




Time t
1
New wave front








Fig. 4
Fig. 3


It can be seen from figure 4 that a short time (t
1
) after the finite source has been energised, the
disturbances from each of the point sources will have moved outwards by the same amount. Along
a line equal to the radius of the small circles, running parallel to the face of the transducer, these
disturbances re-enforce each other to produce a wave front moving out from the transducer.
Notice also, a little energy diffracts around the edge of the transducer and is lost. A little while
later (t
2
), sound from each point source will have travelled a little further and reinforce at a new
distance in front of the transducer, thus the sound wave progresses from the source (figure 5).

Time t
2










Fig. 5

This wave front may represent the initial expansion of the transducer as it starts to vibrate (a
positive going half cycle). It will tend to push particles of the material away from the source.
Shortly afterwards, the transducer will contract as part of its vibration, and a wave front, drawing
particles into the source (a negative going half cycle) will follow on behind the initial wave front,
followed by another push, then another pull and so on.

In the third article in this series, we discussed refraction of the beam as an angled incident wave
meets an interface. The bottom edge of the beam reaches the interface first and takes up the new
velocity. We can use Huyghens principle to explain what happens. As each point along the
incident wave front reaches the interface, each in turn takes the new velocity and in the new
material, the line of initial wave fronts will determine the direction of the refracted beam. Similarly,
in the fifth article we mentioned phased array probes. The shape of the beam and beam angle will
be determined by the wave front for which the individual wave fronts are in phase.

Now consider a point reflector P just in front of the probe centre. Let us consider how this
reflector is affected by just three of the point sources, one in the centre and one at each edge of
the transducer (figure 6).

P



Fig. 6


We energise the source, and a split second later sound from the middle point source reaches P, and
gives it a push (figure 7). Notice that energy from the edges of the probe has not reached P yet.
This will take longer because P is further from the edges than from the centre.



P





Fig. 7

By the time sound from the edges of the transducer reaches P (figure 8) and tries to push P away
from the transducer, the energy from the centre may be on the opposite half cycle of vibration, and
be pulling P back towards the transducer. The resultant force acting on point P will be the vector
sum of the forces acting from all parts of the crystal. In our example, the result is that P doesn't
move at all (i.e. the sound intensity=0). The distance between the solid arc (positive peak) and the
dotted arc (negative peak) is half a wavelength. If a different frequency had been used, it may have
been that the second positive half cycle from the centre of the crystal reached point P at the same
time as those from the edges of the crystal. In that case, the forces would have reinforced and
point P would have been given an extra hard push.



P






Fig. 8


When two solid arcs cross, the forces from those two parts of the crystal are both pushing at
the intersecting point and when two dotted arcs cross the forces from that part of the crystal
are both pulling at the intersecting point. In both cases we call the effect constructive
interference. When a solid arc cuts a dotted arc, the forces are in opposition and we call the
effect destructive interference. Of course point P will not always be exactly a multiple of
half wavelengths away from the center and the edges, and constructive interference happens
when the relevant point sits anywhere within the same half cycle. Destructive interference
happens when the relevant point is in dissimilar half cycles.

Interference occurs whenever energy arrives at different phase (wavelength) intervals at a
particular point. Whether the interference is constructive, or destructive, is determined by the path
difference between P and the centre, and P and the edges. As P gets further away from the front of
the transducer, this path difference becomes negligible compared to the wavelength (figure 9) and
interference problems become insignificant.



P










Next half cycle



Initial wave fronts

Fig. 9

Variations in intensity due to interference effects occur for some distance in front of the
transducer, as we have just seen. This region is known as the Near Field and the extent of
the near field, known as the near field distance can be calculated from: -
4
2
D
NF = Where, NF =Near Field Distance.
D =Crystal Diameter.
=Wavelength
Example 1.
Calculate the Near Field distance for a 10mm diameter, 5MHz crystal transmitting into steel
(Velocity 5960m/sec. =1.192mm).
192 1 4
10
2
. x
NF =
768 4
100
.
NF =
NF = 21mm (Approx.)

This means that for this probe, in steel, we can expect fluctuations in intensity of sound for the
first 21mm of steel depth due to interference effects. As a result, it is unwise to rely solely on
amplitude as the criterion for acceptance or rejection of the part for discontinuities that are in
the near field region.

The last item on our list of factors affecting attenuation of the sound as it travels through a
material is the Beam Spread. Because the beam spreads out into a conical shape, intensity follows
the inverse square law just as it would for a beam of light or X-rays. If you double the distance from the
probe, the intensity drops to one quarter of its original value because of beam spread. Of course, it
will actually fall to less than a quarter, because we have to add any absorption, scatter losses to
the beam spread losses. We can now plot a graph of intensity against distance from the
probe, to summarise the previous discussions. Figure 10 show amplitude on the vertical axis and
distance on the horizontal axis. Distance is shown in multiples of the near field distance.

0 1NF 2NF 3NF 4NF 6NF 5NF
A
m
p
l
i
t
u
d
e















Fig. 10




D

Fig. 11

o
Near
Field

Far
Field








































The beam profile shown in figure 11 is very much a theoretical beam spread. Alongside there
are three slices through the beam showing that the highest sound intensity is in the centre of
the beam. The sound gradually fades away towards the edge of the beam until there is no
sound left. It is often more convenient to define the beam to a theoretical edge where the
intensity of sound has fallen to one half (-6dB), or sometimes one tenth (-20dB) of the intensity
at the beam centre. We can consider three theoretical edges; one defining the absolute edge of
the beam, another defining the 6dB edge and the third defining the 20dB edge. These three
edges can be expressed mathematically: -

D
.
Sin
22 1
2
= Defines the absolute edge
D
.
Sin
56 0
2
= Defines the 6dB edge
D
.
Sin
08 1
2
= Defines the 20dB edge

It is often convenient to use the theoretical beam shape shown in figure 11 in order to explain
some concepts in ultrasonic flaw detection. However it is not good practice to use a
calculated beam shape for sizing discontinuities by one of the intensity drop methods. This is
because practical beam shapes seldom match the theoretical model closely enough.

Example 2.
Calculate the 20dB beam spread angle for a 5MHz compression wave in steel from a 10mm
diameter crystal.
D
.
Sin
08 1
2
=
10
192 1 08 1
2
. x .
Sin =


10
28736 1
2
.
Sin =


128736 0
2
. Sin =


o
4 7
2
. =


o
4 7 2 . x =
=14.8
o


We have used three terms connected with the beam of sound in the test material, namely
Dead Zone, Near Field, and Far Field. The dead zone is that part of the timebase
occupied by the initial pulse when using a single crystal contact probe. The near field is the
distance in the material that suffers from interference effects and the far field is the rest of the
beam beyond the near field. The trace shown in figure 12 is calibrated for 100mm of steel
return path using a single crystal 5MHz compression wave probe. The three zones are shown
on the trace.


0
1 2 3 4 5 6 7 8 9
10
Dead
Zone
Near

Field Far Field

Fig. 12



















BACK TO BASICS ULTRASONICS

Part 8. CALIBRATION & REFERENCE STANDARDS

As soon as one starts to carry out practical ultrasonics, it becomes apparent that neither the vertical nor
the horizontal scales of the display have any absolute meaning per se. The horizontal scale can be adjusted
to represent a great variety of different time intervals, and these, for a given material and velocity, can be
translated into depth values. The vertical scale gives an indication of the amplitude of signal being detected,
provided you know how much Gain you are using, but it does not necessarily tell you much about the size
of defect causing that reflection. The safest way to get more information about the specimen from the
display is to compare signals from the specimen with those from specially machined blocks. These blocks
we normally classify under one of two headings, depending on the function of the block.

The term Calibration Block is defined in British Standard BS 2704 as: - "A piece of material of
specified composition, heat treatment, geometric form, and surface finish, by means of which ultrasonic
equipment can be assessed and calibrated for the examination of material of the same general
composition." Therefore, a calibration block may be a simple step wedge in a particular material to allow the
timebase to be calibrated for accurate thickness measurement, or it may be a more complex block like the
A.2 block described in BS 2704 which allows calibration of timebase, plus calibration of probe index, angle,
resolution etc.

The second heading, Reference Block is defined in BS2704 as: - "An aid to interpretation in the form of
a test piece of the same material, significant dimensions and shape as a particular object under
examination, but not necessarily containing natural or artificial defects". So, for example, a section of an
aircraft wing forging may be prepared as a reference block so that a technician may become familiar with
the standard signal patterns from the various changes in section and more easily recognise a defect
quickly when examining the component on an aircraft. More usually, the block would contain artificial
defects from which the sensitivity (gain) used in the test could be set.

In either case, the use of the block ensures that there is adequate timebase to display the reflecting
surfaces that are of interest and that the test is carried out at a reproducible sensitivity

CALIBRATION BLOCKS
The BS 2704 A.2 Calibration Block, also known as the International Institute of Welding (I.I.W.) block, or
V1 block, is illustrated in Fig.1.The block can be used for the following assessments: -
- Calibration of the timebase in terms of thickness. - Assessment of Dead Zone.
- Checking linearity of the timebase. - Checking linearity of the Amplifier.
- Assessing overall sensitivity of probe and amplifier. - Checking Resolution.
- Determination of the angle of refraction. - Determination of probe index
- Determination of Beam Characteristics. - Finding the correct Zero Point.



4 0 5 0 6 0
60 70 75
25.00 300.00
1
0
0
.
0
0
100.00R
Perspex
insert
50mm diameter hole
1.5mm hole













Fig. 1

The A.2 block was derived from the original Dutch Block designed by RTD Rotterdam and accepted
by IIW as the IIW V1 Block. In its original form, the deep slot at the center of the 100mm radius was
a scribed line and a 25mm radius slot was positioned as shown in Figure 2. This design is still used in
some parts of the world, and has the advantage that shear waves can be calibrated for ranges other
than multiples of 100mm. In all other respects it is the same as the A.2 block.


Scribed line
25mm radius slot












Fig. 2

The BS 2704 A.4 Calibration block, also known as the 'V2 block', is a more compact form of the 'V1 block'
suitable for site use, although somewhat less versatile in its functions. Figure 3 illustrates the A.4 block.



40 50 60
6
5
7
0
75.00
25.00R
50.00R
12.50










Fig. 3



The Institute of Welding (I.O.W.) Beam Profile calibration block is designed primarily for beam profile
measurement and has four 1.5mm diameter side drilled holes giving eight depths from two scanning
surfaces. These can be examined by direct scan for probes of various angles, and at several more ranges
for each probe, using indirect scans by reflecting from the far surface. There are two series of five holes on
an inclined axis to measure shear wave probe resolution and to simulate an inclined discontinuity. The block
is illustrated in Figure 4.


1
3
3
2
2
5
1
9
4
3
6
2
5
0 5
6
2
2
7
5
5
0
305

Plan view
Side view




















Fig. 4

REFERENCE BLOCKS
Area / Distance reference blocks are mainly used for setting sensitivity levels and accept/ reject levels
for defect sizes by reference to echo height. Blocks are produced in a range of scanning depths and each set
of blocks contains the same diameter flat-bottomed hole in each block. There are three sets of blocks, a set
with 3/64 diameter flat-bottomed holes, a set with 5/64" diameter holes and one with 8/64" diameter
holes. The scanning depths can range from to 22, but at shop floor level, you would only have the few
blocks appropriate to your range of work. Figure 5 shows a typical block, in this case a 3 x 5 block (3" scan
depth, 5/64" flat bottomed hole).

A
B




A =Scan depth

B =Hole diameter





Fig. 5



Distance/Amplitude Correction (DAC) reference blocks are made from the same thickness and grade of
material as the work piece. They contain an artificial flaw (a side-drilled hole). The change of echo height
with changes in scanning distance (multiple skips) is noted and plotted on the display as a "DAC" curve so
that a signal amplitude can be specified to cover all depths within the working range for reporting,
acceptance, or rejection purposes. Figure 6 shows a typical ASME DAC block and Figure 7 shows a DAC
curve.


1/4 T
1.5" or T
4" Always
1.5" Always
1.5" Always

















Fig. 6





0 1 2 3 4 5 6 7 8 9 10
D
A
C

C
u
r
v
e



















Fig. 7


Figure 8 shows an example of a reference block for the examination of a lug in a light alloy structure
such as an aircraft fitting. It is made from the same material as the actual part to be inspected, will
have the same surface finish and, in this case contains an artificial defect to aid the setting of
sensitivity and to help with identification of signals during interpretation.








35
o














Fig. 8



References: -
Ultrasonic Flaw Detection for Technicians - Third Edition, June 2004 by J. C. Drury
BACK TO BASICS ULTRASONICS

Part 9. COMPRESSION WAVE TECHNIQUES
CALIBRATION OF TIMEBASE
The important thing to remember when calibrating the timebase for compression waves is that the left
hand end of the timebase (Zero) must exactly correspond to the entry surface of the beam and the
right hand end represents a known thickness in the material being tested. The exceptions to this rule
are those occasions when you are using delay to expand some distant portion of the material, or
when you are using a multiple echo technique and only noting the decay pattern. For single crystal
probes, the initial pulse contains two elements; the applied square wave voltage pulse and the ringing
of the crystal. The end of the applied voltage pulse represents the top surface at which time the
crystal ringing starts. Unfortunately, the amplitude of this part of the initial pulse is so large that it is
not possible to identify the point at which the ringing starts, nor is it possible to tell from the timebase
line.

There is a similar problem when we calibrate using a twin crystal probe because the initial pulse is at
the start of the Perspex delay line and the sound enters the work piece sometime later. In any case,
because the amplifier is deliberately isolated from the transmitter crystal, there is no signal to mark
the entry surface. Our calibration procedure, whether for a single or twin crystal probe must find some
other way to identify the entry surface or true zero. The most common way to achieve this is to use
two echoes that are a known distance apart, to set one at timebase zero and the other at the right
hand end (10) of the timebase. We do this in the following way: -
- Use the delay control to position the first backwall echo from the desired range on our
calibration block to zero.
- Then use the depth range controls to position the second backwall echo to 10 on the
timebase.
- This may also move the echo from the zero position and so we need to check and adjust this
with delay again.
- These two controls are used alternately until the two echoes are exactly on 0 and 10. We now
know that the timebase is exactly equal to the calibration thickness
- Once we are happy that we have that exact range on the timebase, we lock the depth
controls.
- We then use delay to move the first backwall echo to the right until it is exactly on 10. If the
timebase is exactly a known range and the first backwall echo is on 10; then zero must
coincide with the entry surface.

Figures 1 to 3 illustrate the procedure for calibrating the timebase for 100mm of steel on the A2
calibration block.















4 0 5 0 6 0

0 1 2 3 4 5 6 7 8 9 10
2
nd
BWE
1
st
BWE
Depth
Delay


Fig. 1















0 1 2 3 4 5 6 7 8 9 10
1st BWE
2nd BWE
Depth Delay






Fig. 2



0
1 2 3 4 5 6 7 8 9
10
1st BWE
Delay
only

















Fig. 3


In this example we know that the first and second backwall echoes represent 100mm and 200mm of
steel path time because the A2 block is 100mm wide at this probe position. Therefore figure 2
represents exactly 100mm timebase. This timebase remains constant as long as we do not alter the
depth control, so figure 3 represents zero to 100mm exactly.

CHOICE OF COMPRESSION WAVE PROBES
TWIN CRYSTAL PROBES
For conventional techniques twin crystal probes are generally used on thicknesses below 50mm. They
are also in general use for high temperature thickness measurement, where a thermal insulating
material is used instead of Perspex, to protect the crystals. Because the transmitter section of the
probe is isolated from the amplifier, there is no dead zone and so reflections from surfaces close to the
probe can be identified. However since the Perspex shoe absorbs some sound, less damping is used
and the pulses are longer so that resolution is generally poorer than with single probes.

SINGLE CRYSTAL PROBES.
Single crystal probes are generally used on thicknesses in excess of 50mm. They are also used below
50mm if resolution is an important factor, since single crystal probes usually have shorter pulse
lengths than twin crystal probes. However, for conventional techniques they can only be used when
the transmission noise (Dead Zone) does not encroach upon the useful part of the timebase for that
job. As a guide, you can expect the shortest Dead Zone from high frequency, heavily damped probes.

PROBES FOR MULTIPLE ECHO TECHNIQUE
These are usually single crystal probes, although in some cases twin crystals can be used. When
dealing with thin walled material it is possible to get resonance and anti-resonance conditions that will
affect the decay pattern and may give false indications. This can be avoided if you choose a probe
frequency such that the plate thickness is more than 1.5 x the wavelength of the compression wave in
the specimen material, and a pulse length that is not more than 3 cycles.

THICKNESS MEASUREMENT
One of the most important uses of ultrasonics is that of thickness measurement. It is particularly
useful because it is relatively quick, simple and accurate, and access to only one surface of the
specimen is required. There are many types of equipment and techniques made exclusively for
thickness measurement. It is not intended to deal with all of them here. We will only discuss the use of
the pulse echo system with an A-scan display.

A-SCAN RECTIFIED DISPLAY
This is the most common display presentation for ultrasonic flaw detection equipment. In the sixth
article we described the display for an unrectified trace and various types of rectification.
a) CALIBRATION
The basic calibration of the timebase should be carried out to ensure proper positioning of the
zero and backwall echo as described above. The calibration block should be made of the same
material as the work piece.

For best results the range chosen for calibration should be the shortest range which allows the
first back wall echo to be displayed. For example, if the nominal wall thickness of the work
piece is 9mm. and your flaw detector is capable of showing 10mm across the full graticule,
then the 10mm range should be used. Since the graticule of most flaw detectors can be
divided into 100 small units it follows that a timebase calibrated such that those 100 units
represent 10mm gives you a reading accuracy of 0.1mm per division. If on the other hand you
calibrate such that 100 units represent 25mm, the reading accuracy is 0.25mm per division.


b) AMPLITUDE (GAIN SETTING)
The amplitude of the calibration echo and the amplitudes of thickness echoes made on the
work piece should be adjusted to the same predetermined amplitude. This is normally
between 1/3 and 1/2 full screen heights.

c) READING THE THICKNESS (SINGLE ECHO).
The specimen thickness is determined from the left hand edge of the backwall echo. This is
normally a steep sloping line. If a small half cycle appears at the left hand edge of the signal
that was not present during calibration, this may be removed by inserting a small amount of
suppression or by choosing positive or negative rectification. (See figure 4).



















3 4 5 6 7 3 4 5 6 7
Extra half cycle After suppression or
rectification change
Fig. 4
d) READING THE THICKNESS (MULTIPLE ECHOES)
If the specimen thickness and calibrated range are such that multiple echoes are produced, the
most accurate result can be obtained by reading the thickness of the last multiple echo displayed
and dividing the answer by the number of backwall echoes. In the example shown in figure 5, the
fifth backwall echo shows at 22 mm. so the true thickness is 22 divided by 5 =4.4 mm. In this case, a
single crystal probe has been used and the initial pulse is obscuring the start of the first backwall
echo.










Sometimes the initial pulse obscures the entire first backwall echo and maybe all or part of other
back echoes. Figure 6 shows the same thickness but with the first two back echoes obscured. Care
must be taken to assess the number of echoes that have been obscured.














e) USE OF TIMEBASE DELAY
Apart from its use to correct for Perspex path distance in twin crystal compression wave
probes, "Delay" can be used as an aid to more accurate thickness measurement. For
example, you may want to accurately measure the thickness of a component whose nominal
thickness is 80 mm. If you calibrate the timebase so that 100 scale divisions represents 100
mm of that material, each small division represents 1 mm. If instead, you calibrate the
timebase so that 100 scale divisions represent 25 mm of the test material each division on the
scale represents 0.25 mm instead of 1 mm. The delay control can then be adjusted so that
the third backwall echo from the calibration block is set at O, and the fourth backwall echo at
100 scale divisions. The timebase would represent a thickness range of 75mm to 100mm. The
first back echo from the work piece (80mm) will appear at approximately 1/5th of the
timebase range.



0 5 10 15 20 25
Fig. 5
0 5 10 15 20 25
Fig. 6
A-SCAN UNRECTIFIED DISPLAY.
There are occasions in thickness measurement, particularly if the scanning surface is rough, when
a lot of unwanted signals, "noise " or "grass" appear on the CRT and make it difficult to determine
the point at which the back echo starts. If the ultrasonic set allows an unrectified trace to be
selected, then measurements can be made using the tip of a particular down going half cycle instead
of using the point at which the signal first leaves the timebase.

a) Firstly, let us identify our measuring point. Figure 7 shows a back echo from the 25mm range on
the V1 block with the timebase calibrated for 50mm using the conventional rectified display. The
presentation has then been changed to unrectified and the vertical or 'Y' shift used to raise the
timebase to a level between 1/3 and 175th full screen height. Gain has been adjusted so that the
peak of the longest down going half cycle just meets the graticule. In this case, it is the second
half cycle that is the longest, and we will use the second half cycle as our measuring point.
(Note that sometimes a back echo from the work piece may show the 1st or 3rd half cycle as
the longest - despite this, if you calibrate on the second half cycle you then always measure
from the second half cycle even if it is not still the longest,)






















0
1 2 3 4 5 6 7 8 9
10
Fig.7
b) Having identified the half cycle that you are going to use, you calibrate the timebase so that
this part of the signal coincides with the correct point of the graticule. In the case shown in
figure 7, if we wished to calibrate for 50 mm we would use "delay" to move the second half
cycle from 5.15 to 5.0 divisions and check that the second half cycle of the second back echo
coincides with 10.0 divisions (see figure 8).























0 1 2 3 4 5 6 7 8 9 10
Fig. 8
c) Calibration for other timebase ranges would be done in the conventional way but using the
second half cycle instead of the left hand edge of the pulse, as your measuring point.

LAMINATION TESTING
Lamination testing of plates and pipes that are to be welded or machined is a very common NDT
task. It is also a simple application of compression waves in ultrasonic flaw detection, but one that
might give some problems when examining thinner samples.

STANDARD PROCEDURE
a) Calibrate the timebase to allow at least two backwall echoes to be displayed.
b) Place probe on the work piece and adjust the gain controls so that the second backwall echo is
at full screen height.
c) Scan the work piece looking for lamination indications that will generally show up near half
specimen thickness together with a reduction in back echo amplitude. In some cases, a
reduction in the amplitude of the second back echo may be noticed without a lamination
signal being present. Care must be taken to ensure that this reduction in amplitude is not due
to poor couplant or surface conditions.

MULTIPLE ECHO TECHNIQUE
Lamination testing of plate or pipe less than 10 mm. in wall thickness may be difficult using the
standard procedure because multiple echoes are so close together that it becomes impossible to pick
out lamination signals between backwall echoes. In such cases, we can use a technique called the
"multiple echo technique" using a single crystal compression wave probe. The procedure is as
follows: -
a) Place the probe on a lamination free portion of the work or calibration piece.
b) Adjust the timebase and gain controls to obtain a considerable number of multiple echoes in a
decay pattern over the first half of the time base. A typical example is shown in figure 9.




















c) Scan the work piece, the presence of laminations will be indicated by a collapse of the decay
pattern such as the one shown in figure 10. The collapse occurs because each of the many
multiple echoes is closer to its neighbour in the presence of a lamination.
Fig. 9
0 1 2 3 4 5 6 7 8 9 10
Unlaminated plate
0 1 2 3 4 5 6 7 8 9 10



















Laminated plate

Fig. 10

EXAMINATION OF BRAZED AND BONDED J OINTS
Compression waves can also be used for the detection of areas of non-adhesion in brazed or
bonded (glued) joints,
a) BRAZED J OINTS
If the wall thicknesses permit clear separation between back wall echoes, brazed joints can be
examined using the standard procedure for lamination checking. However, since the braze metal
separating the two brazed walls will have a slightly different acoustic impedance to that of the
parent metal, a small interface echo may be present for a good braze. The technique, therefore, is
to look for an increase in this interface echo amplitude. Figure 11 shows the type of indication
when the unbrazed portion is smaller than the beam diameter. If the two brazed walls are too thin to
permit clear back echoes, a multiple echo technique can be used as described above.












Brazed joint Good braze Bad braze
Fig. 11
b) BONDED J OINTS
These may include metal-to-metal glued joints and metal to non-metal glued joints (e.g. rubber
blocks bonded to steel plates). The technique used is a multiple echo technique. Each time the
pulse reaches a bonded interface; a portion of the energy will be transmitted into the bonded layer
and absorbed. Each time a pulse reaches an unbonded layer, all the energy will be reflected. If we
look at the multiple echo pattern for a good bond, the decay will be relatively short because of the
energy loss at each multiple echo into the bond. However, for an unbonded layer each multiple
echo will be slightly bigger because there is no interface loss, and the decay pattern will be
significantly longer. Figure 12 shows a good bond (probe 1) and poor bond (probe 2).

















1 2
Metal
Rubber
Decay pattern for good bond Decay pattern for bad bond
Fig. 12
References: -
Ultrasonic Flaw Detection for Technicians - Third Edition, June 2004 by J. C. Drury

BACK TO BASICS ULTRASONICS

Part 10. SHEAR WAVE TECHNIQUES
Shear waves at various angles of refraction between 35
o
and 80
o
are used to locate defects whose
orientation is not suitable for detection by compression wave techniques. Some defects, of course,
have volume and their shape enables them to be detected by both compression and angled shear waves.
In this chapter, however, we will be dealing with planar defects whose orientation is such that only
angled shear waves can be used.
Because the beam is travelling through the test piece at a refracted angle other than perpendicular,
we need to distinguish between the beam path length to a discontinuity and its depth below the test
surface. When we encounter a signal, we can measure the beam path length (range) from the
timebase, but we may want to calculate how far in front of the probe (horizontal distance) and how
far below the surface the reflector is located. It is also important when using shear waves to know
where along your probe the beam enters the specimen (beam index). Knowing the beam index
position relative to some datum on the specimen, and the exact beam angle allows you to calculate
the horizontal and vertical distances. There are standard terms for various distances when using
shear waves and these are illustrated in figure 1.


Index.
D
C
A
HSD
FSD




FSD =Full skip distance
HSD =Half skip distance
=Beam angle
ABCD=Beam path length
AB =Half skip BPL








B


Fig. 1


Full skip and half skip distances are measured along the top surface and beam path length (BPL),
along the beam centre. To calculate these, knowing specimen thickness (t) and probe angle () use
the following formulae: -
a) HALF SKIP DISTANCE =t x Tan
b) FULL SKIP DISTANCE =2 x t x Tan 0
c) HALF SKIP BPL =
Cos
t

a) FULL SKIP BPL =
Cos
t 2



Example 1
Calculate the Full skip distance for a 40
o
shear wave beam in a 20mm thick steel plate.
xtxTan FSD 2 =
40 20 2 xTan x FSD =
8391 0 20 2 . x x FSD =
FSD =33.564mm


ESTABLISHING THE TRUE PROBE BEAM INDEX
We need to find the exact beam index for any shear wave probe before measuring the beam angle.
This is because the beam index may not be the one marked on a new probe it may be a millimetre
or so before or after the marked index. Manufacturers use a standard drawing to make probes, but
the velocity of sound in Perspex varies from batch to batch, and with temperature. Also the beam
index and probe angle change as the probe wedge or shoe becomes worn with use. So, the
establishing of beam index and angle will be routine checks throughout the life of the probe. Finding
the beam index is a simple procedure carried out on the A2 or A4 calibration block.
The probe is positioned close to the edge of the calibration block and beaming towards the 100mm
radius (A2 block) or the 25mm or 50mm radius (A4 block) as shown in figure 2 (A2 block) and figure 3
(A4block).



40 50 60
6
5
7
0
45

6 0
45












Fig. 3
Fig. 2


The probe is moved backwards and forwards about the centre mark of the radius with the probe kept
parallel to the edge of the block. As the probe moves, the signal will rise to a maximum and then fall
again as shown in figure 4. When the signal reaches the maximum amplitude, the beam centre is
meeting the tangent to the radius at right angles. This happens when the beam centre is entering the
block at the centre of the radius. The true beam index is now in line with the centre mark of the radius
being scanned. If this does not coincide with the beam index marked on the probe, you would then
either mark the true index on the probe body, or, if the probe body has a millimetre scale, make a
note of the true position in front or behind the marked index.



3 4 5 6 7

















Fig. 4



MEASURING THE TRUE BEAM ANGLE
Once the true beam index is known, the true beam angle can be measured on the A2 or A4
calibration block. The nominal probe angle is marked on the probe and is the refracted angle for steel
unless identified for another material. A 45
o
shear wave probe made for aluminium would be marked
45AL and for copper 45CU. The actual angle for a new probe may be plus or minus two degrees
from the nominal angle because of the batch velocity variations in Perspex, and will change with
wear. Most of us have an inherent tendency to wear the probe in a particular way, just as we do for
shoes. We may wear the heel of the probe down and so increase the actual angle, or wear the toe
and decrease the actual angle. For this reason the beam angle measurement is also a routine probe
check. If the probe is worn down towards one edge, the beam will be thrown off towards that side and
the condition is called squint.

Beam angle is measured on the A2 block by aiming the beam at the 45mm diameter hole and on the
A4 block at the small hole. The probe is positioned on the block at a point near the nominal angle and
the gain adjusted to give a signal amplitude of about 50% full screen height. As the probe is moved
forward and back, the signal rises and falls just as it did when finding the beam index. When the
signal reaches its maximum amplitude, the beam centre is aimed at the centre of the hole and the
beam is hitting the tangent to the hole at right angles. The true beam angle can be read against the
true beam index from the graticule on the calibration block. The example shown in figure 5 has the
beam index opposite an angle of about 43
o
and the nominal angle is 45
o
. With this probe, we would
have to use 43
o
in our distance calculations and for defect sizing.


6 0 7 0
40 50 60
45











Fig. 5


CALIBRATION OF TIMEBASE
The method of calibration of the timebase for shear waves depends on the purpose of the inspection.
If the inspection were to be volumetric, looking for any discontinuities within the scanned volume of
the test piece, then we would calibrate for a suitable timebase range at shear wave velocity. On the
other hand, if the purpose is to look for a specific discontinuity such as a fatigue crack, in a predicted
location, we may well use a Skip method or a Reference Block method. The calibration for a
known range will be dealt with first, using the A2 block and then the A4 block.

USING THE A2 BLOCK
- Place the probe on the A2 block as shown in figure 6.
- Obtain a maximum echo from the 100mm radius.
- Adjust the gain control to peak the signal at about 80% full screen height.
- Use the delay control to position the 100mm signal at zero on the timebase.
- Use the depth controls to place the second reflection from the 100mm radius at ten on the
timebase.
- Check that the left hand edges of the two signals are exactly at zero and ten.
- Lock the depth controls
- Use delay to move the first signal from zero to ten.
- The time base is now calibrated for 100mm at shear wave velocity, and zero represents the
top surface entry point below the beam index.













45
0 1 2 3 4 5 6 7 8 9 10
Delay Depth
100mm echo 200mm echo
Fig. 6
Sometimes you may see part of the initial (transmission) pulse around the zero, this will depend on
the pulse length and gain setting as shown in figure 7.


0 1 2 3 4 5 6 7 8 9 10
100mm echo
Residual initial pulse
Delay















Fig. 7


The slot that marks the 100mm radius on the A2 block is about 4mm deep so that when the probe is
aligned with the edge of the block, the slot makes a corner for the returning echo to reflect part of the
energy back to the 100mm radius. This is why it is possible to obtain repeat echoes from the radius. If
the slot were not there, the reflected energy from the first returned signal would reflect to the rear of
the probe. Figure 8 shows an exaggerated view of the beam path to illustrate the corner effect.


45
4mm deep slot
Exagerated beam path



















Fig. 8

USING THE A4 BLOCK
To calibrate for 100mm using the A4 block: -
- Place the probe on the A4 block as shown in figure 9.
- Obtain a maximum echo from the 25mm radius.
- Adjust the gain control to peak the signal at about 80% full screen height.
- Use the delay control to position the 25mm signal at 2.5 on the timebase.
- Use the depth controls to place the second reflection (from the 50mm radius slot) at ten on
the timebase.
- Check that the left hand edges of the two signals are exactly at 2.5 and 10.
- Lock the depth controls
- The time base is now calibrated for 100mm at shear wave velocity, and zero represents the
top surface entry point below the beam index.



















40
50 60
6
5
7
0
45
0 1 2 3 4 5 6 7 8 9 10
25mm echo 100mm echo
Delay
Depth
Fig. 9
The sound path in figure 9 shows the first echo from the 25mm radius and then the echo from the
50mm radius after reflecting at the scanning surface down to the 25mm radius and back. The total
return path is: -
25mm +50mm +25mm =100mm.
Facing the 25mm radius on the A4 block, signals will arrive at 25, 100, 175, 250mm and so on,
incrementing by 75mm each time. If the probe is turned around to face the 50mm radius, signals will
arrive at 50, 125, 200, 275mm and so on, again incrementing by 75mm.

CALIBRATION USING THE SKIP METHOD
If the purpose of the inspection is to detect surface breaking flaws at the bottom surface or top
surface (typical of fatigue cracks), we know that the echoes will arrive at exactly the half skip or full
skip beam path lengths. We could calibrate the timebase for an exact range using one of the methods
described above and calculate the beam path lengths for half and full skip using the formulae. We
would then know exactly where to look on the timebase for the two conditions. We do use this method
to carry out the critical root scan in weld inspection.

However, in many cases there is a quicker and simpler method. Using a piece of plate of the same
wall thickness as the item to be inspected we can point the probe at the end surface (position 1) and
scan back as shown in figure 10 until we see the echo from the bottom corner (position 2).


45 45 45
F
u
l
l
s
k
i
p
H
a
l
f


s
k
i
p
Scan
1 2 3













Fig. 10



The signal will rise to a maximum as the centre of the beam moves into the corner. We can adjust the
timebase and gain to make sure that we can see that maximum point. As the maximum is reached,
we would adjust the timebase range to position the signal at some convenient part of the trace,
usually about 4. We would then continue moving the probe backwards until the top corner reflection
is seen (position 3). As this signal maximises, we note its position along the timebase. Figure 11
shows a trace with the half skip and full skip positions marked, and in this example, gates positioned
over the two critical locations so that the operator can listen for the alarm rather than watch the
display all the time.

Another point to note from figure 11 is that the position for full skip is at 9 on the timebase and not at
8 (twice 4). This means that timebase zero is not the top surface, and furthermore, we dont know
the exact timebase range. However, for this inspection it doesnt matter because we are only
interested to find out whether or not there is a bottom or top surface breaking corner.

0 1 2 3 4 5 6 7 8 9 10
Gates
Half skip zone Full skip zone















Fig. 11


If the plate to be inspected has accessible edges, you dont need a calibration plate because you can
use the corners on the test piece to set up the two positions on the timebase. However, you have to
be sure that there are no laminations in the beam path because these might reflect the beam back to
the top without reaching the bottom. It is easy to check whether the signal is coming from the
anticipated corner because a shear wave meeting an interface at an oblique angle is easy to damp.
If you put an oily finger on the expected reflecting corner, the signal will be seen to reduce in
amplitude significantly. In figure 10, if you damp the bottom corner when the beam is at the half skip
position (2), the signal will fall and when the beam is at the full skip position (3) you can damp the
signal at the top corner and at the reflecting point on the bottom surface.

PIPE WALLS
If you are going to scan a pipe wall in the longitudinal direction, then you can use any of the above
calibration procedures. However, if you are scanning circumferentially the calculation of beam path
length, and skip distances is more complicated. If you have a segment of pipe of the same outside
diameter and wall thickness as a reference block, you can use the skip method for finding the critical
half and full skip positions on the timebase. If you also need to look for discontinuities in the volume of
the object, you calibrate the timebase on the A2 or A4 block for an exact range, and then put the
probe on the reference pipe segment and note the half and full skip ranges.

The wall thickness for any given outside diameter is important because the normal range of angled
shear wave probes (45
o
, 60
o
,and 70
o
), when used on thick wall pipe may cut across to the outside
surface again without touching the bore. An example is shown in figure 12 where a 45
o
shear wave
only reaches about half way through the wall. In other words, for this outside diameter, the thickest
pipe wall we could test with a 45
o
probe is only half that shown in the diagram. It follows that, when you
are presented with an unusually thick pipe wall for a particular outside diameter, you need to choose
your probe angle carefully in order to inspect the bore properly. For a given angle, the maximum wall
thickness that allows the centre of the beam to reach the bore of the pipe can be calculated from: -

( )
2
1 Sin dx
t

=

Where t =wall thickness
d =pipe outside diameter
=beam angle






Thickest wall that can be
tested with this 45
o
probe








Fig. 12
This formula can be turned round so that you can calculate the best angle given for a wall thickness,
the formula becoming: -


d
t
Sin
2
1 =

Table 1 shows maximum wall thickness that can be tested for three standard angles and a range of
pipe diameters.

Pipe OD

Maximum pipe wall thickness for probe angles
35
o
45
o
60
o
4 (100mm) 21mm 14mm 7mm
6 (150mm) 32mm 22mm 10mm
8 (200mm) 43mm 29mm 13mm
10 (250mm) 53mm 36mm 17mm
12 (300mm) 64mm 44mm 20mm
14 (350mm) 74mm 51mm 23mm
16 (400mm) 85mm 58mm 27mm
18 (450mm) 96mm 66mm 30mm
20 (500mm) 106mm 73mm 33mm

Table 1

Once the correct angle for the pipe size and wall thickness has been chosen, you can establish the
skip and half skip positions using a section of pipe with a drilled hole to produce the required corner
reflectors as shown in figure 13.








Fig. 13



CALIBRATION OF THE GAIN
This is often called "setting the sensitivity", and it means that we adjust the gain so that a significant
discontinuity will give a signal that is large enough to see, but small surface scratches will not. Very
often, we use a reference block, similar in shape and material to the specimen, and containing either a
drilled hole or an artificial (machined) crack. The probe is aimed at this reference hole or crack, to
obtain an echo, this is then maximised by probe movement, and then, the gain is adjusted to give the
required signal height known as the reference level and the gain is then said to be calibrated. This
reference level may be 50% or 75% of full screen height, and is often used as the basis for getting
acceptance standards for the inspection. Hence, you may find that you are working to a specification
that says that any signal equal to, or greater than the amplitude of the reference level is cause for
rejection of the component, whereas any signal lower than the reference level may be ignored.

TESTING FOR OUTSIDE DIAMETER SURFACE FLAWS
Discontinuities that break the top surface such as the crack show in figure 14 will cause a reflection to
occur at exactly the beam path distance for the full skip if a suitable angle that will reach the bore is
used. However, as you can see from figure 15, if you are testing a thick wall tube or a solid bar, the
beam may reach the top surface without first reflecting from any other surface. The beam path length
at which a top surface defect will appear in that case can be calculated from the formula: -

cos D BPL = Where: - D is the outside diameter and is the probe angle.



Fig. 15
Crack














Fig. 14



In the sort of application illustrated in figure 15, if there is no crack, the sound will carry on around the
bar or pipe as shown in figure 16. Provided there is enough sensitivity, you may only need to scan
from position A to position B. The beam will sweep the entire circumference during the short scan
and as long as you have enough timebase and gain, echoes from any discontinuities breaking the
surface will appear at predictable positions.








A
B







Fig. 16









References: -
Ultrasonic Flaw Detection for Technicians - Third Edition, June 2004 by J. C. Drury


BACK TO BASICS ULTRASONICS

Part 11. SURFACE & TOFD TECHNIQUES
SURFACE WAVE TECHNIQUES
Surface waves have been used very successfully for a great number of applications, particularly in
the Aircraft Industry. However, it is not so common in the Steel Industry because surface finishes are
often less smooth, and magnetic particle inspection will find most defects detectable by surface
waves. Nevertheless, there are occasions when the use of surface wave techniques can give the
simplest and most positive results and so, in this section we will discuss some general principles that
can be applied when considering a surface wave technique.

ADVANTAGES OF SURFACE WAVES
Surface waves will follow gentle contours without reflection, but will reflect sharply from a sudden
change in contour. Figure 1 shows a typical example of a component having a complex shape that
would make the use of shear or compression waves difficult, if not impossible. Cracks may develop
anywhere along the leading or trailing edge of the blade out to about two thirds of the blade length, or
in the root radius. A surface wave probe placed at the end of the blade, and directed towards the root
will send a beam along the surface, round the radius and reflect from the edge of the root as shown.
Cracks in the suspect areas will give reflections at an earlier time than the root.



Fig.1
















The fact that surface waves only penetrate to a depth of about one wavelength can be used to
advantage when testing relatively thin wall sections. Figure 2 shows a pipe with a change of section.
We are told that cracks may occur on the inside or outside diameter of the pipe in the necked region.
An angled shear wave probe might be used, but it would be difficult to predict the skip points as the
beam bounces around the section change. However, if we choose a surface wave at a frequency for
which the wavelength is approximately equal to the wall thickness, then the surface wave will fill the
wall thickness, and follow the section change, reflecting for a defect breaking either surface.










SF


Fig. 2


LIMITATIONS OF SURFACE WAVES
The main limitation of the surface wave technique is that the beam is almost immediately attenuated
if the surface finish is rough, covered in scale, or a liquid (such as the couplant), or has any pressure
applied by another object (such as your hand). For this reason it is normal to use grease as the
couplant for surface wave probes (it doesn't run!), and to apply the grease to the probe, place the
probe on the job and scan forward (away from your own grease trail). Ridges left in the couplant
during scanning, and objects resting on the test surface, often give spurious signals that might be
taken to originate from defects so it is normal to test such indication by rubbing a cloth over the area
indicated by the signal. If after this cleaning operation, the signal disappears, then it was a spurious
indication.

CALIBRATION & DEFECT LOCATION
It is not usual to calibrate the timebase for surface waves in the way we would do for shear or
compression waves. This is because we can normally run a finger along the surface in front of the
probe, when we find a defect indication, until the signal is no longer damped. This happens as we
pass over the defect with our finger, however, there are occasions when access is limited and we are
directing surface waves to a region that is out of sight and cannot be reached with the hand. In these
cases, the timebase can be calibrated using the same procedure as for shear waves on an A2 or
similar block.

The sensitivity can be set from drilled holes or spark-eroded slits in suitable reference blocks. In the
aircraft industry, these reference blocks are usually sections of an actual component with a spark-
eroded slit in the critical location.

DEFECT SIZING USING TOFD
The TOFD technique, first used by Silk in 1977, uses tip diffraction to identify the top, bottom and
ends of a discontinuity in one pass. Silk chose to use an angled compression wave for the TOFD
technique rather than a shear wave, for two reasons. Firstly, the tip diffraction signal is stronger than
a shear wave diffraction signal, and secondly, a lateral wave is produced which can be used to
measure the horizontal distance between the transmitter and receiver.

The tip diffraction signal is generated at the tip of the discontinuity effectively a Point source.
According to Huyghens, a point source produces a spherical beam. Figure 3 shows a typical TOFD
transducer set-up on a component with a vertical discontinuity. There are four sound paths from the
transmitter to the receiver. Path A is the lateral wave path travelling just below the surface. Path B
is the tip diffraction path from the top of the discontinuity. Path C is the tip diffraction path from the
bottom of the discontinuity and path D is the backwall echo path.

Figure 4 shows a typical unrectified trace for the four signals. Note the phase relationships, A and C
are in opposite phase to B and D. The important difference to note is between B and C the top and
bottom diffraction signals are in opposite phase. This phase difference allows the practitioner to
identify those points.



R T
Lateral wave
Diffraction echoes
Backwall reflection
from top & bottom
Backwall D
C
B
A










Fig. 3


A B
C
D










Fig. 4


Assuming that the diffracting tip is centred between the two transducers, the depth of the tip below
the surface can be calculated from: -
2 2
2 2

=
HD BPL
Depth
Where:
BPL =Beam path length for the signal in question
HD =Beam path length for the lateral wave.

The distance measurements taken from the ultrasonic trace must be made from the same part of
each waveform. In the trace shown in Figure 4, the largest half cycle would be selected. For signals A
& C this is negative and for signal B positive. The advances in computer technology have made it
possible to carry out all the calculations and plotting to be handled automatically and stored for
subsequent evaluation. The method that has been chosen to display this TOFD data presents the
information in a special B-scan form that is easy to assimilate. The way in which the positive and
negative half cycles are displayed needs explaining.

In a conventional B-scan image, the slice is taken across the weld perpendicular to the centre line.
In the TOFD display, the slice is taken along the weld (figure 5). However, whereas the conventional
B-scan is a relatively thin slice, the TOFD image represents the volume between the probes as they
scan along the weld. The presentation is known as a D-scan.



Cap
Weld length


Root

Fig. 5


An echo arriving at the receiver is a pulse of a certain pulse width and amplitude. In conventional B-
scan displays, this pulse is displayed as a bright spot whose diameter is proportional to the pulse
width and whose brightness is proportional to the signal amplitude. In some ways, it is like a broad
pencil tip that can be used to draw pictures in light or bold broad strokes. The pulse is really a short
burst of a few cycles of alternating waveform. In the TOFD system, the waveform is depicted in
greyscale with positive going half cycles tending towards white, and negative going half cycles
tending towards black (see figure 6). This type of display will allow us to identify phase change so that
we can discriminate between he lateral wave, top and bottom defect signals and backwall.














Fig. 6

This allows particular half cycles to be identified for measurement purposes, and phase changes to
be recognized for determination of top or bottom echo. Figure 7 shows a typical computer screen for
a TOFD inspection. The image shows details of the component (in this case, a weld) as well as the
TOFD D-scan image and an A-scan trace. In this image, left to right represents the component
thickness, and the vertical dimension represents scan length.

The A-scan trace shown corresponds to a slice through the weld at the location indicated by the
cross hairs of the cursor. The striped band on the left of the TOFD image represents the lateral
wave, and the bold striped band to the right of the image represents the backwall echo. The
difference in boldness is due to the different signal amplitudes. Following the horizontal cross hairs
and about half way between the lateral wave and backwall stripes, a series of feint horse shoe
shaped stripes can be seen. These are diffraction signals from a small discontinuity. The A-scan trace
shows the signal clearly.














Fig. 7
In this example, the discontinuity has a very small dimension in the through-thickness dimension, but
close study of the A-scan shows a small phase shift in the last half-cycle of the discontinuity signal.
This tells the practitioner that the distance from top to bottom of the discontinuity is about the same as
the pulse length for this particular discontinuity.









Fig. 8

A much bolder indication can be seen towards the top of the lateral wave line suggesting a
discontinuity at, or just below the surface. In figure 8, the cursor has been moved to this location. The
lateral wave signal can be seen to be longer and stronger than at the previous location. The fact that
the wave shape stays in phase suggests that the diffraction echo, which is extending the signal, has
the same phase as the lateral wave. In other words, it is a bottom tip signal. However, it is not
possible in this case to see where the lateral wave ends and the bottom tip begins, and so it is not
possible to say how deep the discontinuity extends below the surface. The TOFD method is limited in
its ability to size near surface discontinuities when the arrival time difference between the lateral wave
and the diffraction signal is similar to pulse length. Near surface resolution when using TOFD can be
a bit confusing if you look at it from a conventional ultrasonics point of view. Imagine a top surface
crack 4mm deep. At 5MHz, it represents more than two wavelengths at compression wave velocity
and with a reasonably short pulse of two cycles; you might expect to resolve the bottom of the defect.
However, the path difference between the lateral wave and the tip diffraction signals for a probe
separation of 80mm is only 0.4mm and this is about the same as the wavelength for 15MHz (See
figure 9). You would need a 15MHz transmitter with only one cycle in the pulse to resolve the crack.

80.4mm
80mm





Fig. 9


The transducers used in TOFD techniques are angled compression wave transducers. The common
angles used are 60
o
and 70
o
, although other angles may be used if the component thickness makes it
necessary. The design and construction of the transducer is important in order to promote a good
lateral wave. Previous theory has suggested that a shear wave should also exist in the component
and this is true, it does. Figure 10 shows a little more of the trace for the above example. On the
extreme right of both the A-scan and TOFD D-scan, the shear wave can be seen. Since it arrives well
after the other signals, it does not present a problem in this application.



Shear wave










Fig. 10

Scanning with the TOFD system is fast and many scanning systems are motorized. They all require
distance encoders so that the D-scan image can be constructed. The vertical extent of those defects
that can be resolved is many times more accurate than other sizing systems.

References: -
Ultrasonic Flaw Detection for Technicians - Third Edition, June 2004 by J. C. Drury

Anda mungkin juga menyukai