Sei sulla pagina 1di 9

Downloaded from rspa.royalsocietypublishing.

org

Manganin gauge and VISAR histories in shock-stressed


polymethylmethacrylate
N. K. Bourne and Z. Rosenberg
Proc. R. Soc. Lond. A 1999 455, 1259-1266
doi: 10.1098/rspa.1999.0358

Email alerting service Receive free email alerts when new articles cite this article - sign up in the box at the top right-hand
corner of the article or click here

To subscribe to Proc. R. Soc. Lond. A go to: http://rspa.royalsocietypublishing.org/subscriptions

This journal is © 1999 The Royal Society


Downloaded from rspa.royalsocietypublishing.org

Manganin gauge and VISAR histories in


shock-stressed polymethylmethacrylate
By N. K. B o u r n e† a n d Z. Rosenberg‡
PCS, Cavendish Laboratories, Department of Physics, University of Cambridge,
Madingley Road, Cambridge CB3 0HE, UK

Received 7 November 1997; revised 20 February 1998; accepted 30 March 1998

Shock-wave profiles have been measured by many techniques probing the stress or
particle velocity histories of shocked materials. Among the most commonly used
methods are the velocity interferometry system for any reflector (VISAR) and embed-
ded foil stress gauges. It is relatively unusual to employ both these techniques simulta-
neously in an experiment. The gauge is backed using a transparent rear window made
of polymethylmethacrylate (PMMA) through which the VISAR is also shone. Exam-
ples are given of results from the two techniques for metal, ceramic and glass targets,
and the direct comparison validates the accuracy of earlier gauge data. Finally, the
results are used to construct the principal Hugoniot of the PMMA window material
to high accuracy.
Keywords: shock loading; velocity interferometry; stress gauges; shock Hugoniot;
polymethylmethacrylate; phase transition

1. Introduction
Over the past 50 years a range of sensors has been developed to follow the state
of a material that is being dynamically loaded. The aim of an impact experiment is
usually to characterize the variables that describe the state of the impactor or target,
in general to define the pressure, volume and temperature as a function of time.
Invariably these parameters are not measured directly; instead one monitors related
quantities such as stress, strain or particle velocity. An overview of the physical
principals and experimental methods governing shock propagation may be found in
Davison & Graham (1979).
In this application, one of the most commonly used coupled mechanical/electrical
properties is piezoresistivity, which is the change of the electrical resistivity of a
material with applied external stresses. One such piezoresistive material is manganin
(an alloy of 84% copper, 12% manganese and 4% nickel), which has been calibrated
over a wide pressure range (Rosenberg et al . 1980).
The principal advances in dynamic instrumentation since the 1970s have been
in the development of non-invasive interferometric techniques for the monitoring of
the velocity of the external surfaces of impact targets. A major step forward came
with the realization that the Doppler shift of the light could be used to produce
interference patterns from rough surfaces. These velocity interference patterns vary
† Present address: DEOS, Royal Military College of Science, Cranfield University, Shrivenham, Swin-
don SN6 8LA, UK (n.k.bourne@rmcs.cranfield.ac.uk).
‡ Permanent address: RAFAEL, PO Box 2250, Haifa, Israel.

Proc. R. Soc. Lond. A (1999) 455, 1259–1266 c 1999 The Royal Society

Printed in Great Britain 1259 TEX Paper
Downloaded from rspa.royalsocietypublishing.org

1260 N. K. Bourne and Z. Rosenberg

Target

Flyer
impact Longitudinal
position stress gauge
z Spo t σx

PMMA
Window VISAR
fibres
y
x
Figure 1. Experimental arrangement for simultaneous gauge/VISAR experiments.

much more slowly than fringes in a distance-measuring interferometer so that they


can be followed with relative ease. This system (which was dubbed VISAR) was
originally developed by Barker at Sandia (Barker & Hollenbach 1972; Barker &
Schuler 1974), and has now reached a position of dominance for the measurement of
normal velocity of a surface.
In a previous paper we presented an analysis of the ringing that is present in the
traces obtained by manganin gauges placed in the so-called ‘backsurface configu-
ration’ in which a polymethylmethacrylate (PMMA) backing is used to match the
impedance of the epoxy encapsulating the gauge (Bourne & Rosenberg 1996, 1997).
In this work we use this PMMA backing material as a window to make a simulta-
neous VISAR measurement adjacent to the gauge. This simultaneous measurement
allows direct comparison of the two techniques and verifies the accuracy of previously
obtained gauge results as well as yielding a direct Hugoniot for the PMMA window
material.

2. Experimental
(a) Comparison of the techniques
Plate-impact experiments were carried out in the 50 mm bore gun at the Univer-
sity of Cambridge. Stress profiles were measured with commercial manganin stress
gauges placed on the rear face of the specimens and supported with PMMA blocks.
These gauges (Micromeasurements type LM-SS-125CH-048) have been calibrated by
Rosenberg et al . (1980). The signals were recorded using a fast (1 GS s−1 ) digital
storage oscilloscope and transferred onto a microcomputer for data reduction. Impact
velocity was measured to an accuracy of 0.5% using a sequential pin-shorting method
and tilt was fixed to be less than 1 mrad by means of an adjustable specimen mount.
Impactor plates were made from lapped tungsten alloy, copper and aluminium discs
and were mounted onto a polycarbonate sabot with a recessed front surface in order
that the rear of the flyer plate was a free surface.
The VISAR used was an adapted VALYN system (Barker 1996) with polymer
fibres feeding into the enclosure and illuminating the target. The output of the four

Proc. R. Soc. Lond. A (1999)


Downloaded from rspa.royalsocietypublishing.org

Manganin gauge and VISAR histories in shock-stressed PMMA 1261


1.6

Longitudinal stress in PMMA (GPa)


1.2 0.30

Velocity (mm µs–1)


0.8 0.20

0.4 0.10

0.0 0.00

0.0 0.4 0.8 1.2 1.6 2.0 2.4 2.8


Time (µs)

Figure 2. VISAR (dotted line) and gauge (solid line) mounted in backsurface configuration.
Lead glass DEDF impact with a 3 and 6 mm thick Al flyer at 518 m s−1 .

fringe photomultipliers is recorded directly onto a 2 GS s−1 storage oscilloscope. The


design is of standard push–pull form and subtraction of the signals was carried out in
software on a computer that downloaded the recorded data from the scope. The delay
etalons limit the resolution of the system used in this configuration to ca. 2 ns. The
phase data were unwrapped using the method of Hemsing (1979). Where necessary a
flash coating was evaporated onto the rear of the target to aid reflection. A schematic
of the experimental arrangement is shown in figure 1.
The following data represent typical traces selected from ongoing research in the
shock loading of metals, glasses and ceramics. Each series of data from the two
methods is recorded simultaneously from the same experiment. The VISAR traces are
presented as dotted while the manganin histories are solid lines. A range of stresses
is used to illustrate features of each response and the series of experiments presented
is ordered according to the stress induced. In all cases remarkable agreement is seen.
However, some features appear on some traces and not on others, and it is these that
are discussed in detail.

3. Results and discussion


Figure 2 shows two shots conducted on the lead glass DEDF (i.e. type D, extra
dense flint). This has been the subject of extensive investigation in our group over
the past few years (Bourne et al . 1996a, b). Each experiment is conducted close to
the Hugoniot elastic limit (HEL) of the material, which has been determined to be
4.5 GPa. Each DEDF tile was of the same thickness (8 mm) but the aluminium flyer
thickness was varied in order that the releases from the rear of the flyer and the
PMMA backing material superposed at different positions within the tile. In neither
case was any spall fracture seen. The DEDF contains ca. 30% silica and 70% PbO
by weight so that it is poorly conducting under ambient conditions. There is no

Proc. R. Soc. Lond. A (1999)


Downloaded from rspa.royalsocietypublishing.org

1262 N. K. Bourne and Z. Rosenberg

0.6

2.0

Longitudinal stress in PMMA (GPa)

Velocity (mm µs –1)


1.6 0.4

1.2

0.8 0.2

0.4

0.0 0.0
0 0.4 0.8 1.2
Time (µs)
Figure 3. VISAR (dotted line) and gauge (solid line) mounted in backsurface configuration.
Symmetrical impact of a 97.5% alumina flyer of thickness 3 mm and target 6 mm at 688 m s−1 .
The arrowed dip in the gauge trace, A, is discussed in the text and is not seen in the VISAR
trace.

significance in the separation in rise time of the two pairs of traces; they have been
offset for clarity of presentation. It will be noted that velocity and stress histories
follow one another very closely upon loading and release. It will be seen that the
maximum stress in the gauge is ca. 1.5 GPa, which is the HEL of manganin. Thus
the gauge material is behaving elastically throughout the loading. There are slight
deviations between the velocity and stress at the very end of the pulses, but these
may be attributed to the arrival of lateral releases in the gauge, which gives rise to
an extra strain contribution to the signal resulting in an apparent rise in stress.
The most significant differences are ringing on the gauge traces at the bottom and
top of the shock wave where the rate of rise of the pulse is at a maximum. This ringing
is superposed upon a steady response which follows the VISAR signal. This ringing
has been reported before and has been the subject of further analysis (Bourne &
Rosenberg 1996, 1997). In those works we showed that conducting materials showed
an initial dip at the arrival of the shock wave, which corresponded to a capacitive
linkage between the gauge and the conducting target. On the top of the pulse, a
ringing appeared which had a frequency dependence upon the inductance of the
gauge, the capacitance of the gauge and target and of the connecting cabling. Some
authors have attempted to explain such ringing by mechanical arguments (Bourne &
Rosenberg 1997). The figures presented here show conclusively that the ringing has
an electrical origin, and that once deconvolved from such effects the gauges follow
precisely the observed particle velocity history at the PMMA–target interface.
Figure 3 shows simultaneous VISAR and gauge traces from impact upon a 97.5%
alumina. This material and others of differing alumina content have been investigated
in depth by Murray and co-workers (Murray et al . 1996; Murray 1997) and the figure
can be regarded as a typical trace. Again the agreement between the VISAR and the

Proc. R. Soc. Lond. A (1999)


Downloaded from rspa.royalsocietypublishing.org

Manganin gauge and VISAR histories in shock-stressed PMMA 1263

1.6

Longitudinal stress in PMMA (GPa)


0.4

1.2

Velocity (mm µs–1)


0.8
0.2

0.4

0
0 1 2 3
Time (µs)
Figure 4. VISAR (dotted line) and gauge (solid line) mounted in backsurface configuration.
Impact of a copper flyer of thickness 3 mm onto a 12 mm mild steel target at 536 m s−1 . The
arrowed dips in the gauge trace, A and B, are discussed in the text and are not seen in the
VISAR trace.

gauge is very good. The velocity trace is displaced 20 ns before the gauge. The stress
range now extends to 2.4 GPa, which results in plastic deformation of the gauge.
There are several features to be noted. Again there is ringing on the gauge trace
of the type discussed above, which has the same period and source. The initial dip
has, however, disappeared. This is a result of the insulating nature of the ceramic
under ambient conditions. We have recently shown that the break in slope after the
rapid elastic rise is related to a surface fracture zone near the impact plane (Bourne
et al . 1998). It will be seen that the gauge shows the start of ringing at this time,
which may also indicate that the material is conducting as a result of free charge
produced on fracture.
The dip has been used in the past as a measure of the HEL of the ceramic.
This may, however, simply be a lower strain-rate-dependent yield (Grady 1996). In
particular, several authors have shown that this threshold decays with distance (see,
for example, Murray et al . 1996). The comparison with the VISAR signal shows
that the break occurs later and at a slightly higher point in the stress history trace
(arrowed at A). This is to be expected since the gauge rise is limited to 30 ns by its
inductance so that it slightly overestimates the threshold.
Figures 4 and 5 show the response of metals as seen by gauges and VISAR in the
backsurface configuration. In both these figures, traces are presented of the impact
on mild steel of flyers of copper (figure 4) and tungsten alloy (figure 5). In figure 5
the 13 GPa phase-transition stress is exceeded in the steel, pushing the gauge to a
stress of ca. 3.5 GPa. In both figures the first dip is seen after the arrival of the elastic
precursor (arrow A). This has been discussed above for conducting materials.
It will be seen that there is a second dip in the gauge trace (arrow B) when the
target and gauge are pushed still closer together by the arrival of the main plastic
wave. The effect is larger in figure 5 at this point since the acceleration is greater. In

Proc. R. Soc. Lond. A (1999)


Downloaded from rspa.royalsocietypublishing.org

1264 N. K. Bourne and Z. Rosenberg

0.8

Longtudinal stress in PMMA (GPa)


0.6

Velocity (mm µs–1)


2
0.4

1
0.2

0 0

1 2
Time (µs)
Figure 5. VISAR (dotted line) and gauge (solid line) mounted in backsurface configuration.
Impact of a tungsten alloy flyer of thickness 5 mm on a 12 mm mild steel target at 707 m s−1 .
The arrowed dips in the gauge trace, A and B, are discussed in the text and are not seen in the
VISAR trace.

the top of the shock in figure 4 and on the plateau in figure 5 is a Gibbsian overshoot
and some damped ringing giving rise to the dip as expected. Again this may be
analysed using previous work (Bourne & Rosenberg 1997).
There is no ringing in figure 5 when the compression pulse reflected from the phase-
transformed material arrives at the gauge. This is because the wave is slow rising (as
it results from a reflected release fan). This illustrates that a fast-rising pulse (less
than 100 ns) is necessary to observe ringing on gauge histories. Traces for materials
such as soda-lime or borosilicate glasses, whose open structure results in mechanical
ramping of the shock front, show no deviation of the stress and particle velocity
histories and no ringing and have thus not been included here. It is to be emphasized
that the results presented are representative of the worst cases of deviation of gauges
from particle velocity histories. When an embedded gauge is used to measure the
phase transition in this material directly, the wave rises more rapidly and this is
sufficient to result in an overshoot at the bottom and top of the history.
There is deviation of the velocity and stress histories evident in both figures in the
release phase after ca. 2.5 µs. This may be due to permanent plastic deformation in
the gauge, but it is more likely (particularly since the gauges were slightly off-axis to
accommodate the VISAR spot) that lateral releases have reached either measuring
location by this stage, resulting in deviations.

(a) Hugoniot for PMMA


Figure 6 shows the results from the figures plotting stress against particle velocity.
This trace is plotted against previous experimental Hugoniot data from two sources
(Barker & Hollenbach 1970; Marsh 1980). These points were determined by mea-
surements from individual experiments where stress and particle velocity were not

Proc. R. Soc. Lond. A (1999)


Downloaded from rspa.royalsocietypublishing.org

Manganin gauge and VISAR histories in shock-stressed PMMA 1265

Marsh
3
Barker & Hollenbach
Stress (GPa)

0 0.2 0.4 0.6 0.8


Particle velocity (mm µs–1)
Figure 6. Hugoniot for PMMA. VISAR (particle velocity) and gauge (stress) results plotted for
the three flyer/target configurations used in the paper. Comparison with well-known Hugoniot
data determined in an entirely different manner is shown.

measured directly, but values were deduced from measured maximum stress and
impact velocity. The simultaneous measurements in DEDF, alumina and iron allow
a continuous curve to be plotted, which agrees very well with the independently
determined Hugoniot points. The data where the elastic wave rises (faster than the
gauge can adequately respond) have been omitted. Similarly, the release wave data
are largely left unplotted.

4. Conclusions
We have confirmed that manganin gauges follow the stress history of uniaxial strain-
impact experiments and have presented simultaneous VISAR traces that are immune
to electrical interference as a comparison. We have noted several features of gauge
behaviour that must be borne in mind when using them. First, for slow (greater than
100 ns) rising pulses the gauges follow precisely the VISAR signal. For fast-rising sig-
nals, a ringing is observed due to the reactance of the gauge and its associated cabling
and environment. We have shown that this ringing is apparent particularly at the
threshold for elastic behaviour in ceramics and that this point slightly overestimates
that same threshold seen in the VISAR signal. Finally, we have shown that the stress
and strain data may be used to plot a continuous Hugoniot for the backing material,
which agrees well with other published data. A similar method may be used for any
transparent material whose properties are not known.
The authors gratefully acknowledge the support of DERA and, in particular, Dr B. Goldthorpe
for making funds available to purchase the VISAR. We thank Professor J. E. Field, Dr W. G.
Proud, Dr I. G. Cullis and Dr P. Church for useful discussions, and G. Stephens and D. L. A.
Cross for technical support. N.K.B. acknowledges the EPSRC for a fellowship.

Proc. R. Soc. Lond. A (1999)


Downloaded from rspa.royalsocietypublishing.org

1266 N. K. Bourne and Z. Rosenberg

References
Barker, L. M. 1996 User manual, Valyn VISAR Albuquerque, NM: Valyn International.
Barker, L. M. & Hollenbach, R. E. 1970 Shock-wave studies of PMMA, fused silica, and sapphire.
J. Appl. Phys. 41, 4208–4226.
Barker, L. M. & Hollenbach, R. E. 1972 Laser interferometer for measuring high velocities of
any reflecting surface. J. Appl. Phys. 43, 4669–4675.
Barker, L. M. & Schuler, K. W. 1974 Correction to the velocity-per-fringe relationship for the
VISAR interferometer. J. Appl. Phys. 45, 3692–3693.
Bourne, N. K. & Rosenberg, Z. 1996 Fractoemission and its effect upon noise in gauges placed
near ceramic interfaces. In Shock compression of condensed matter 1995 (ed. S. C. Schmidt
& W. C. Tao), pp. 1053–1056. Woodbury, NY: American Institute of Physics.
Bourne, N. K. & Rosenberg, Z. 1997 On the ringing observed in shock-loaded piezoresistive
stress gauges. Meas. Sci. Technol. 8, 570–573.
Bourne, N. K., Millett, J. C. F. & Rosenberg, Z. 1996a The shock wave response of a filled glass.
Proc. R. Soc. Lond. A 452, 1945–1951.
Bourne, N. K., Rosenberg, Z. & Ginzburg, A. 1996b The ramping of shock waves in three glasses.
Proc. R. Soc. Lond. A 452, 1491–1496.
Bourne, N. K., Millett, J. C. F., Rosenberg, Z. & Murray, N. H. 1998 On the shock-induced
failure of brittle solids. J. Mech. Phys. Solids. 46, 1887–1908.
Davison, L. & Graham, R. A. 1979 Shock compression of solids. Phys. Rep. 55, 255–379.
Grady, D. E. 1996 Shock-wave properties of brittle solids. In Shock compression of condensed
matter 1995 (ed. S. C. Schmidt & W. C. Tao), pp. 9–20. Woodbury, NY: American Institute
of Physics.
Hemsing, W. F. 1979 Velocity sensing interferometer (VISAR) modification. Rev. Sci. Instrum.
50, 73–78.
Marsh, S. P. 1980 Shock Hugoniot data. Berkley, CA: University of California Press.
Murray, N. H. 1997 PhD thesis, University of Cambridge, UK.
Murray, N. H., Bourne, N. K. & Rosenberg, Z. 1996 Precursor decay in several aluminas. In
Shock compression of condensed matter 1995 (ed. S. C. Schmidt & W. C. Tao), pp. 491–494.
Woodbury, NY: American Institute of Physics.
Rosenberg, Z., Yaziv, D. & Partom, Y. 1980 Calibration of foil-like manganin gauges in planar
shock wave experiments. J. Appl. Phys. 51, 3702–3705.

Proc. R. Soc. Lond. A (1999)

Potrebbero piacerti anche