Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

ALLTEM UXO Detection Sensitivity and Inversions for Target Parameters from

Yuma Proving Ground Test Data

David L. Wright, Craig W. Moulton, Theodore H. Asch, Philip J. Brown II, U.S. Geological Survey
Misac N. Nabighian and Yaoguo Li, Colorado School of Mines
Charles P. Oden, Earth Science Systems, LLC

Abstract

ALLTEM is a multi-axis electromagnetic induction system designed for unexploded ordnance


(UXO) applications. It uses a continuous triangle-wave excitation and thus measures target step
response rather than the more common impulse response. In May, 2006, we operated ALLTEM over
the Army’s UXO Calibration Grid and Blind Test Grid at the Yuma Proving Ground (YPG), Arizona.
The system multiplexed through all three orthogonal (Hx, Hy, and Hz axes) transmitting loops and
recorded a total of 19 different transmitting (Tx) and receiving (Rx) loop combinations. This was
accomplished while in continuous motion with a spatial data sampling interval, after waveform
averaging, of 15 cm to 20 cm. ALLTEM records data at a constant 100 kilosamples/s rate at 24-bit
precision rather than in a small number of time gates.
Maps produced from differences between early- and late-time unfiltered signal amplitudes are of
high quality, but exhibit ground response and system thermal drift effects. We find that by exploiting
the high density of the time-series data to digitally filter the data and by moving the early-time pick to
275 µs, late enough that the step response of an analog low-pass filter has settled, instead of the 75 µs
early time we had previously used, we can produce amplitude difference data and maps that are almost
free of ground response and system drift effects while retaining good sensitivity to UXO. The
improvement in the signal-to-noise ratio (SNR) greatly enhances the ability to detect small or deep
targets and also produces more consistent inversion results from ALLTEM data.
An inversion algorithm has been developed and applied to data from various sets of the available
19 Tx-Rx combinations over a number of targets. We present results that show that the inversion
algorithm produces accurate parameters for some known targets in the Calibration Grid. This suggests
that it is possible to obtain good multi-axis system target inversions from moving platform data even
with some position “noise.”

The ALLTEM System

ALLTEM is one of several multi-axis electromagnetic and magnetic prototype systems being
developed for UXO applications (Barrowes et al., 2006; Snyder and George, 2006; Smith, J.T., et al.,
2004; Smith, D.V., et al., 2004; Wright et al., 2006). The motivation for using multi-axis excitation and
observation of target response is that the additional information aids in going beyond simple detection of
metal objects in the ground to the critical task of discriminating between probable UXO and harmless
scrap metal that can be safely left in the ground. The system was described in Wright et al., 2006, but
since that time some additional Rx coils have been added and additional Tx-Rx polarization
combinations have been implemented. The system in its present configuration is shown in Figure 1.
System control and data acquisition software, written in National Instruments LabVIEW®, provides the
flexibility of selecting a single Tx polarization or multiplexing through all three polarizations. All data
shown in this paper were acquired while the system was multiplexing.

1422
Figure 1.: This annotated photograph of the ALLTEM system shows the three orthogonal square Tx
loops and the square Rx loops of 1 m and 34 cm size.

An important system change made between 2005 and 2006 was the adoption of a Leica 1200 real
time kinematic global positioning system (RTK-GPS) because it has a 20 Hz position update rate that is
fast enough to greatly reduce or perhaps even eliminate the previous requirement to interpolate position
data. Dynamic tests indicate that the positions for the rover GPS antenna are accurate to within 2 cm
when the solutions are in “fixed” mode. Although there are additional position errors due to cart roll,
pitch, and yaw, these errors were often small enough that the inversions provided good estimates of
target position, depth, and orientation, and reasonable and reproducible values for dipole moments of
these targets, even though the system was moving.
In Wright et al., 2006 we discussed advantages of using a triangle-wave excitation as was
pioneered by the UTEM system (West et al., 1984). One advantage is that the responses of ferrous and
non-ferrous metal objects have opposite polarities. Another advantage is that the response of a ferrous
target has a non-zero late time response.

Primary Leakage Suppression


An excitation or primary signal that is always on, however, has practical consequences similar to
those of frequency domain systems. In particular, leakage of the primary signal into the received signal
and unwanted response to the ground are two factors that we have to address. We suppress primary

1423
signal leakage in three ways. The first reduction is geometrical; the receiving loops are symmetrically
placed and recorded as opposite gradient pairs. Secondly, since it is very difficult to place coils with
enough precision to completely eliminate primary imbalances and almost impossible to maintain
complete balance as temperature changes, we also use programmable differential gains on the
gradiometer loop outputs. The gains are set under software control. Finally, in processing we typically
select a few waveforms at the beginning and/or end of each line (file) where we know there are no
targets and subtract the average or an interpolated value of the averages of these waveforms from each
of the other records along the line.

Rejecting Ground Response


ALLTEM also responds to the conductivity and magnetic susceptibility of the ground. The
ground conductivity response dies very quickly, much faster than the responses from metal objects
buried in the ground. After the short ground conductivity transient, the response to the magnetic
susceptibility in the ground is a square wave whose amplitude depends on the ground susceptibility and
on the cart height and attitude relative to the ground. Because of the varying amplitude of this response,
we do not use simple signal amplitude for target detection, but rather a difference between the waveform
amplitude at a relatively early time and at a later time as illustrated in Figure 2.

Figure 2.: This figure shows a waveform (black curve), a late-time cursor (red) and two early-time
cursors (green and blue). Amplitude difference maps are produced from differences between voltages at
the vertical red cursor time and either the vertical green cursor or vertical blue cursor time.

Initially we used voltage differences between a late time, such as at the vertical red cursor at
5095 µs, and an early time pick at 75 µs (vertical green cursor) because the resulting amplitude
differences are large. An example amplitude difference map produced from data over the YPG
Calibration Grid using this early time pick is shown in Figure 3. Although this map shows most of the
targets, considerable instrument drift and ground response are evident in this map and some targets
cannot be seen. Subsequently, we performed additional filtering to reduce noise in the data and made
maps over the Calibration Grid and the Blind Test Grid using the 275 µs time (vertical blue cursor) for
the early pick. A map from the same data, but with the background subtracted, additional filtering and

1424
using the later time pick is shown in Figure 4. Although we lose some target signature amplitude,
instrument drift and ground response are almost entirely suppressed.

Figure 3.: This raw data amplitude difference map uses the earlier (75 µs) time pick. The open black
circles show target locations. No background removal or other processing has been applied.

The maps shown in Figures 3 and 4 are produced from vertical polarization data only and thus
correspond most closely to maps of single polarization data from baseline instruments such as a Geonics
EM-61. A comparison between the maps of Figures 3 and 4 makes it clear that because the ground
response, instrument drift, and general noise level have been well suppressed, several more targets are
now easily seen in the data. For example, the targets in cells H1 through M1 in Figure 4 are all clear,
but in Figure 3 most of them are not evident.
There are some other particularly difficult targets in the Calibration Grid. At cell locations A2
and B2 there are two horizontally buried 81 mm M374 mortar rounds, one nose north and the other nose
south. These are deep, 1.5 m or 18.5 times the target diameter, and spaced only 1 m apart. However, in
a close examination of Figure 4 we conclude that ALLTEM detected them, but their signatures are
merged because their depth of burial exceeds their separation and thus the peak amplitude response is
between the two targets and not directly over either of them. At such a low SNR it is not clear that an
algorithm could reliably determine that there are two targets rather than one. Other difficult targets
include I10 and H10, I11 and H11, and I12 and H12. These are, respectively: horizontally buried 2.75”
M230 rockets buried at 1.2 m, or 17 times the diameter, oriented nose grid north and grid south; 60 mm
M49A3 mortar rounds buried at an indicated depth of 0.4 m (we think this is really 0.91 m), nose north,
and 0.91 m nose south; and 57 mm M86’s buried nose north and nose south at 0.91 m deep. Although
the SNR is low it appears that these targets were also detected, though in pairs, rather than individually.

1425
Figure 4.: This map was produced from the same data as the map of Figure 3, but the data have been
filtered and the later (275 µs) early-time pick was used. A grid and the Army lane designations have
been added to the figure. The open black circles indicate target locations, as in Figure 3. Several
additional targets, as compared to those of Figure 3, are clearly visible in this map.

Maps from Other Polarizations

The other excitation and observation polarizations provide additional information about buried
targets. We show examples of two orthogonal horizontal polarization maps in Figures 5 and 6. If the
buried objects all had a vertical axis of rotational symmetry, all target signatures in Figures 5 and 6
should ideally be identical except for a 90 degree rotation. For rotationally symmetric UXO items that
are not buried with the axis of symmetry vertically oriented, however, this is not the case.

1426
Figure 5.: This figure shows an amplitude difference map produced from data where both the excitation
and observation directions (x-direction) are horizontal and perpendicular to the direction of travel.
Figure 7 includes a schematic top view of the cube that illustrates the x and y directions.

Figure 6.: This is an amplitude difference map when excitation and observation directions (y-direction)
are horizontal and parallel to the direction of travel.

1427
For example, cells B4 and D4 (see Figure 4) contain 155 mm projectiles buried horizontally in
the magnetic north-south direction. The response is stronger when the excitation and observation
directions are parallel to the long axis of the projectiles (x-direction case of Figure 5). These differences
are used by the inversion algorithm as discussed in the following section.

Inversion

Comparisons we made between ALLTEM and an EM61-MK2 indicated that ALLTEM, after
processing the data as discussed above, detected the targets in the Calibration Grid at least as well as the
EM61-MK2, but detection alone is not sufficient. Discrimination between probable UXO items and
harmless scrap metal that could be left in the ground is the goal. We developed an inversion algorithm
for ALLTEM data and have applied the algorithm to a number of known targets in the Calibration Grid.
The algorithm is part of a processing and inversion package called GP Workbench (Oden, 2006; Oden
and Moulton, in review) and uses a physics-based non-linear inversion method. It minimizes an
objective function that is a measure of the difference between experimental data and forward modeled
data using a Gauss-Newton method (Gill et al., 1986). ║Yp-Y║/║Y║, where Yp is the predicted data
vector and Y is the measured data, is the normalized mean squared error (MSE) and measures the
difference between the final forward model and the data. ALLTEM inversions are not overly sensitive
to noise and in our experience an MSE below 0.1 suggests a good inversion. Failed inversions typically
have an MSE well above 0.2, but there can be exceptions. Because the inverse problem is non-linear, an
inversion might find a local minimum and not the desired global minimum. This possibility can be
mitigated by a number of methods, including constraining the solutions by various means and by using
numerous initial models. Zhdanov, 2002, and Oldenburg and Li, 2005, discuss geophysical inversions.

In order for the physics-based forward/inverse modeling to work, all the relevant ALLTEM
system parameters must be determined either by calculation or calibration. We use a mixed method in
which some parameters are calculated from first principles while others are derived from field
calibration data. GP Workbench calculates the magnetic intensity produced by each of the three
orthogonal Tx loops (see Figure 1) using the Biot-Savart law. On the other hand, GP Workbench uses
field calibration data to derive the electronic gain of each Rx channel since these gains can be changed
or could drift and accurate gain values for each Rx channel are needed by the inversion.

Figure 7 shows a comparison between measured data and modeled data for six of the possible
polarizations. The measured data are in the inside panels and the data calculated by the forward model
are on the outside. The target is a 60 mm mortar round and the direction of travel was perpendicular to
the long axis of the mortar as shown in the inset. Four of the panels are from data recorded by four 34
cm antenna pairs on the bottom and top faces of the cube. The locations of the antennas are shown at
the top center which is a schematic of a top view of the cube. The other two panels are from horizontal
gradient data recorded parallel to and perpendicular to the long axis of the target. Each plot in Figure 7
is a surface fit to individual data points. The locations of the measurement points are given by the dots
in the measured data panel labeled “ZZF.” For most of the cases the visual agreement between the
measured and final forward-model calculated data is good so that a small MSE is expected. A small
MSE is favorable, although that does not by itself guarantee accurate calculated target parameters.

In the tables that follow, we show inversions for several targets in the Calibration Grid. For all
the cases we consider in this paper the two time picks were 275 µs and 5095 µs as shown in Figure 2.

1428
Figure 7.: This figure shows measured data (central panels with gray background) compared to
calculated data from the forward model (outside panels). Visually, the agreement is quite good. The
target is a 60 mm mortar round as shown at the upper right. The scale on the color bars is in mV.

1429
Because of the triangle-wave excitation that ALLTEM uses, not only are transient eddy currents induced
in metal objects, but ferrous targets are being actively, continuously, and presumably linearly,
magnetized. Since the open-circuited Rx loops produce voltages proportional to the component of dB/dt
normal to the plane of each coil, where B is the magnetic induction, the responses from time-varying
target magnetization go to non-zero constant voltages at late times. We consider two cases for induced
moments. The first case we call the “magnetodynamic” or early-time case that is strongly influenced by
the induced and decaying eddy currents. The second we call “magnetostatic” and that case is dominated
by the late-time magnetization of the targets. For large ferrous targets we do not reach the magnetostatic
limit within our 5.55 ms time window. However, in the dominantly magnetostatic case, i.e. when most
of the transient response is over, and for ferrous rotationally symmetric bodies that are longer than they
are thick (many UXO objects), we expect that there would be one larger polarizability moment and two
smaller ones and that the two smaller ones should be the same. For spherical isotropic bodies all three
calculated moments should be the same. Tables 1 and 2 show results from two boundary marker
spheres. The inversion finds target locations in X and Y. We have removed the most significant digits
to reduce the size of the cell entries. The full Universal Transverse Mercator (UTM) coordinates are
X(m) = 757xxx.xxx, and Y(m) = 3638yyy.yyy. In the tables that follow the first data row contains the
given ground truth values and the second data row contains the calculated values. Asterisks indicate that
the parameter is either not meaningful, or not known. For example, in Table 1 azimuth and inclination
have no meaning for a sphere although the inversion will calculate numbers for those parameters since
the calculated dipole moments are not precisely the same. M1, M2, and M3 are the three calculated
orthogonal target polarizability dipole moments. The last column MSE is the final mean squared error.

Table 1.: Inversion over 8 pound steel shot at grid location N4.
X (m) Y (m) Depth Azimuth Inclination M1 (m3) M2 (m3) M3 (m3) MSE
(m) (degrees) (degrees)
681.916 392.56 -0.20 * * * * * *
681.749 392.392 -0.24 105.56 47.0 0.55 0.41 0.51 0.010

Table 2.: Inversion over 8 pound steel shot at grid location N5.
X (m) Y (m) Depth Azimuth Inclination M1 (m3) M2 (m3) M3 (m3) MSE
(m) (degrees) (degrees)
685.601 391.636 -0.20 * * * * * *
685.388 391.422 -0.23 168.93 53.1 0.52 0.51 0.46 0.008

The three calculated dipole moments are approximately equal as they should be for a sphere. We
are assuming that the magnetic susceptibility is isotropic in the steel shots. The MSE in both inversions
is low, indicating a good fit between the measured and forward modeled data.
In Tables 3, 4, and 5 the targets are 60 mm M49A3 mortar rounds.

Table 3.: Inversion over 60 mm M49A3 at grid location F10 with lines run east to west.
X (m) Y (m) Depth Azimuth Inclination M1 (m3) M2 (m3) M3 (m3) MSE
(m) (degrees) (degrees)
695.636 373.609 -0.25 10.60 0.0 * * * *
695.387 373.766 -0.27 15.75 0.0 2.19 0.38 0.39 0.035

1430
In Table 3 the calculated depth, azimuth and inclination are quite good and the M’s show one
large polarizability dipole moment and two smaller and almost equal moments. This is expected for
rotationally symmetric bodies such as the M49A3.

Table 4.: Inversion over 60 mm M49A3 at grid location F10 with lines run south to north.
X (m) Y (m) Depth Azimuth Inclination M1 (m3) M2 (m3) M3 (m3) MSE
(m) (degrees) (degrees)
695.636 373.609 -0.25 10.60 0.0 * * * *
695.636 373.543 -0.26 16.80 2.9 1.99 0.43 0.35 0.026

A comparison between Tables 3 and 4 is instructive because the target is the same one, but the
two inversions used different data sets acquired along lines run in orthogonal directions. In addition, the
data for Table 4 were acquired at low speed and with 0.25 m line spacing instead of the ordinary 0.5 m
line spacing. Both inversions are reasonable and the calculated parameters are not far apart, but the
additional data obtained at the lower speed did not significantly improve the inversion result in this case.

Table 5.: Inversion over 60 mm M49A3 at grid location M11.


X (m) Y (m) Depth Azimuth Inclination M1 (m3) M2 (m3) M3 (m3) MSE
(m) (degrees) (degrees)
700.225 387.001 -0.48 10.60 45.0 * * * *
700.225 386.877 -0.47 2.88 49.8 1.98 0.42 0.31 0.045

The calculated depth in Table 5 is very close to the actual depth and the inversion has also done a
very good job of getting the azimuth and inclination. The dipole moments are close to those of Tables 3
and 4. Since the target is deeper, the SNR is lower, but the MSE is still small.
In Tables 6, 7, and 8 we show inversions for 20 mm M55 projectiles. These are among the
smallest items in the UXO inventory. They are often difficult to detect, and are therefore even more
challenging to invert for target parameters.

Table 6.: Inversion over a 20 mm M55 at grid location I15


X (m) Y (m) Depth Azimuth Inclination M1 (m3) M2 (m3) M3 (m3) MSE
(m) (degrees) (degrees)
706.565 377.595 -0.20 10.60 0.0 * * * *
706.218 377.260 -0.18 192.13 0.0 0.08 0.01 0.01 0.297

The MSE in Table 6 is significantly higher than in the previous cases, but the inversion has
nevertheless produced very reasonable results. The calculated azimuth is very close to correct except
that the nose and tail of the projectile are reversed. The M’s indicate that the object is axisymmetric and
small.

Table 7.: Inversion over a 20 mm M55 at grid location H15.


X (m) Y (m) Depth Azimuth Inclination M1 (m3) M2 (m3) M3 (m3) MSE
(m) (degrees) (degrees)
706.191 375.584 -0.20 190.60 0.0 * * * *
706.248 375.558 -0.18 13.87 1.1 0.08 0.01 0.01 0.235

1431
The result for the inversion in Table 7 is also very reasonable. Once again, the inversion got the
nose and tail reversed. In this case the calculated polarizability dipole moments are the same as for the
previous case and the MSE is similar between the two cases.

Table 8.: Inversion over a 20 mm M55 at grid location J15.


X (m) Y (m) Depth Azimuth Inclination M1 (m3) M2 (m3) M3 (m3) MSE
(m) (degrees) (degrees)
706.968 379.574 -0.14 * -90.0 * * * *
706.986 379.531 -0.17 344.87 88.0 0.12 0.01 0.01 0.030

The inversion results shown in Table 8 calculated an almost vertical orientation for this projectile
which is correct except once again getting the nose and tail swapped. The largest calculated dipole
moment is higher in this case than for the previous two cases, but the results still indicate that the object
is small and axisymmetric. Vertically oriented targets produce higher responses to vertically oriented
fields and observation directions. In addition, this target is slightly shallower than the previous two.
This may explain why the MSE is much better for this case.
Tables 9, 10, and 11 show inversions of a 57 mm M86.

Table 9.: Inversion of a 57 mm M86 at grid location I12.


X (m) Y (m) Depth Azimuth Inclination M1 (m3) M2 (m3) M3 (m3) MSE
(m) (degrees) (degrees)
700.714 378.775 -0.91 10.6 0.0 * * * *
700.298 379.000 -1.86 194.32 64.0 5.72 12.23 17.96 0.276

The inversion shown in Table 9 has failed. None of the calculated parameters are reasonable.
The cause for this failure is that the target is so deep that the SNR is too small. There is a rule of thumb
that UXO targets can be reliably detected to a depth of about 10 or 11 times their diameter. Although
many targets can be detected well beyond this rule of thumb, successful inversion needs some minimum
SNR. This target is, in fact, one of a difficult pair of targets at I12 and H12 (see Figure 4) discussed
above. The MSE is no higher than for some of the 20 mm inversions that produced reasonable results.

Table 10.: Inversion of a 57 mm M86 at grid location K12.


X (m) Y (m) Depth Azimuth Inclination M1 (m3) M2 (m3) M3 (m3) MSE
(m) (degrees) (degrees)
701.449 382.654 -0.49 * 90.0 * * * *
701.148 382.241 -0.38 119.27 71.4 0.79 0.64 0.38 0.403

The results shown in Table 10 are not very good, but are clearly better than for the much deeper
target of Table 9, even though the calculated MSE is worse.

Table 11.: Inversion of a 57 mm M86 at grid location M12.


X (m) Y (m) Depth Azimuth Inclination M1 (m3) M2 (m3) M3 (m3) MSE
(m) (degrees) (degrees)
702.15 386.624 -0.46 10.60 45.0 * * * *
701.871 386.279 -0.47 5.99 43.7 1.76 0.47 0.42 0.163

1432
The target considered in Table 11 is only slightly shallower than the previous one, although
oriented differently, but the results are dramatically better. All of the calculated parameters are
reasonable and the MSE is considerably lower, though not as low as for some of the shallow 60 mm
targets. These results suggest that fairly sharp boundaries can exist between some cases in which the
inversions are quite good and other cases in which the inversions are not good, but MSE alone is not
sufficient to define that boundary.

Inversion Issues
The inversions we have presented in this paper are generally very encouraging, but much
remains to be done. For instance, we do not know whether there are parameter ranges where good
inversions are likely, and whether certain results, in addition to the MSE, can be used to indicate when
the inversions can or cannot be trusted. By studying the topography of the objective function, it may be
possible to determine which basins of attraction are relevant for certain classes of targets. This might
increase confidence in target classification and perhaps lower the minimum required SNR. We currently
are examining how much each additional polarization adds to the accuracy and reliability of the
inversions. We used only six of the nineteen recorded polarizations in the inversions shown in this
paper and relied upon multiple locations of the system to provide sufficient data to calculate all the
parameters for which we were solving. The required position and orientation accuracies to obtain good
inversions are also not precisely known for ALLTEM. It has been reported that multi-axis data can
tolerate more position error than single-axis data, perhaps up to about 5 cm (Collins, 2006; Tantum et
al., 2006). Our results tend to support such a conclusion. The positional accuracy of our ALLTEM data
at the Calibration Grid was usually within 5 cm when the cart was on level ground, but the cart was not
always on level ground. A 10 degree cart tilt induces about a 21 cm position error. Future plans include
adding roll, pitch, and yaw measurements to the data stream to allow correction for topography-induced
position errors.
The three calculated orthogonal polarizability dipole moments are functions of time and we will
use the time dependence of these moments as another parameter to aid target identification. As yet we
have not obtained a low noise and precisely controlled set of data over a suite of known targets.
However, we plan to make such measurements on an outdoor elevated test stand away from ground
effects and indoor electromagnetic noise. These measurements should help us relate our data and
inversion-derived dipole moments to particular target shapes, sizes, and materials and help answer some
other remaining questions, for example, the sensitivity to position and orientation errors and how close a
large target can be to the ALLTEM sensors before the single dipole representation is inadequate.

Conclusions

The 2006 data sets acquired over the Blind Test Grid (not shown) and the Calibration Grid at the
Yuma Proving Ground are of high quality and many targets can be seen in amplitude difference plots of
raw data, although system thermal drift and ground response are evident in such raw data maps.
We have found that by filtering the data during post-processing and by a judicious choice of time
picks from which amplitude differences are calculated, the system drift and ground response are almost
entirely eliminated. Without such processing the ground response may reach levels of +/- 20 mV or
more, whereas the processed YPG data commonly show remaining noise of only about +/- 1 mV and
sometimes even less. Maps of processed amplitude differences are among the best we have seen for
electromagnetic induction data at the Calibration Grid. The effect of the noise reduction is very
significant at the detection level and appears to be equally important for inversions to derive target
location, depth, orientation, and polarizability dipole moments that provide information related to target

1433
size and shape. The examples we have shown support the view that good target parameter inversions
using moving platform data are possible. Results should further improve as position and orientation
errors are reduced. Continuing work will focus on test stand measurements and further characterization
and improvement to the inversion algorithm. It may prove beneficial to constrain or confirm the results
with other techniques. Examples are the use of the Extended Euler Method to derive depths for UXO
(Davis et al., 2005) and monitoring the total transient field rotation with time.

Acknowledgments and Disclaimer

This research was supported wholly by the U.S. Department of Defense, through the Strategic
Environmental Research and Development Program (SERDP). We also acknowledge the support of
Yuma Proving Ground and Aberdeen Test Center personnel who made this work possible.
Any use of trade, product, or firm names in this paper is for descriptive purposes only and does
not imply endorsement by the U.S. Government.

References

Barrowes, B., O’Neill, K., Snyder, D.D., George, D.C., Shubitidze, F., 2006, New man-portable vector
time domain EMI sensor and discrimination processing, in Proc. of the UXO-Countermine-
Range Forum, July 10-13, 2006, Las Vegas, NV, (Power Point presentation, 18 slides).

Collins, L.M., 2006, Statistical and adaptive signal processing for UXO discrimination for next-
generation sensor data: SERDP Project MM-1442 In-Progress Review, Feb. 22, 2006, (Power
Point presentation, 49 slides)

Davis, K., Li, Y., and Nabighian, M, 2005, Automatic detection of UXO magnetic anomalies using
extended Euler deconvolution: 75th Annual International Meeting, Society of Exploration
Geophysicists, Expanded Abstracts, pp. 1133-1136.

Gill, P.E., Murry, W., and Wright, M.H., 1986, Practical optimization: Amsterdam, Elsevier, 401 p.

Oden, C.P., 2006, Calibration and Data Processing Techniques for Ground Penetrating Radar Systems
with Applications in Dispersive Ground: Ph.D. Dissertation, Dept. of Geophysics, Colorado
School of Mines, Golden, CO, 249 p.

Oden, C.P., and Moulton, C.W., in review, GP Workbench Manual: Technical Manual, User’s Guide,
and Software Guide: U. S. Geological Survey Open-File Report.

Oldenburg, D.W., and Li, Y., 2005, Inversion for applied geophysics: a tutorial, in Butler, D.K, ed.,
Near-surface geophysics: Tulsa, Society of Exploration Geophysicists Investigations in
Geophysics No. 13, pp. 89-150.

Smith, D.V., and Bracken, R.E., 2004, Field experiments with the tensor magnetic gradiometer system
at Yuma Proving Ground, Arizona, in Proc. of the Symp. on the Application of Geophysics to
Engineering and Environmental Problems (SAGEEP), February 2004, pp. 1675-1690.

1434
Smith, J.T., Morrison, F.H., and Becker, A., 2004, Resolution depths for some transmitter-receiver
configurations: IEEE Transactions on Geoscience and Remote Sensing., vol. 42, no. 6, pp. 1215-
1221.

Snyder, D.D., and George, D.C., 2006, Qualitative and quantitative UXO detection with EMI using
arrays of multi-component receivers, in Proc. of the Symp. on the Application of Geophysics to
Engineering and Environmental Problems, April 2-6, 2006, pp. 1749-1760.

Tantum, S.L., Wang, Y.Q., and Collins, L.M., 2006, Statistical and adaptive signal processing for UXO
discrimination for next-generation sensor data, in Proc. on Progress in Electromagnetic Research
Symposium, March 26-29, 2006, pp. 302-305.

West, G.F., Macnae, J.C., and Lamontagne, Y., 1984, A time-domain electromagnetic system measuring
the step response of the ground: Geophysics, vol. 49, pp. 1010-1026.

Wright, D.L., Moulton, C.W., Asch, T.H., Brown, P.J., Hutton, S.R., Nabighian, M.N., and Li, Y., 2006,
ALLTEM for UXO applications – first field tests, in Proc. of the Symp. on the Application of
Geophysics to Engineering and Environmental Problems, April 2-6, 2006, pp. 1761-1775.

Zhdanov, M.S., 2002, Geophysical inverse theory and regularization problems: Amsterdam, Elsevier,
628 p.

1435

You might also like