0% found this document useful (0 votes)
280 views29 pages

Historical Development of The Magnetic Method in Exploration

The magnetic method has evolved significantly since its earliest uses for navigation. It was a key exploration tool developed during World War II using new fluxgate magnetometers. Since then, instrumentation and data processing have improved dramatically, allowing high-resolution mapping of crustal geology from regional to local scales. The magnetic method is now widely used for mineral, oil/gas, geothermal, and groundwater exploration as well as other applications like hazards assessment and engineering studies.

Uploaded by

Gleizer Ferreira
Copyright
© Attribution Non-Commercial (BY-NC)
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
280 views29 pages

Historical Development of The Magnetic Method in Exploration

The magnetic method has evolved significantly since its earliest uses for navigation. It was a key exploration tool developed during World War II using new fluxgate magnetometers. Since then, instrumentation and data processing have improved dramatically, allowing high-resolution mapping of crustal geology from regional to local scales. The magnetic method is now widely used for mineral, oil/gas, geothermal, and groundwater exploration as well as other applications like hazards assessment and engineering studies.

Uploaded by

Gleizer Ferreira
Copyright
© Attribution Non-Commercial (BY-NC)
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 29

GEOPHYSICS, VOL. 70, NO. 6 (NOVEMBER-DECEMBER 2005); P. 33ND61ND, 6 FIGS. 10.1190/1.

2133784

75th Anniversary The historical development of the magnetic method in exploration

M. N. Nabighian1 , V. J. S. Grauch2 , R. O. Hansen3 , T. R. LaFehr4 , Y. Li1 , J. W. Peirce5 , J. D. Phillips2 , and M. E. Ruder6

ABSTRACT
The magnetic method, perhaps the oldest of geophysical exploration techniques, blossomed after the advent of airborne surveys in World War II. With improvements in instrumentation, navigation, and platform compensation, it is now possible to map the entire crustal section at a variety of scales, from strongly magnetic basement at regional scale to weakly magnetic sedimentary contacts at local scale. Methods of data ltering, display, and interpretation have also advanced, especially with the availability of low-cost, high-performance personal computers and color raster graphics. The magnetic method is the primary explo-

ration tool in the search for minerals. In other arenas, the magnetic method has evolved from its sole use for mapping basement structure to include a wide range of new applications, such as locating intrasedimentary faults, dening subtle lithologic contacts, mapping salt domes in weakly magnetic sediments, and better dening targets through 3D inversion. These new applications have increased the methods utility in all realms of exploration in the search for minerals, oil and gas, geothermal resources, and groundwater, and for a variety of other purposes such as natural hazards assessment, mapping impact structures, and engineering and environmental studies.

HISTORY OF MAGNETIC EXPLORATION


The earliest observations on magnets are supposedly traced back to the Greek philosopher Thales in the sixth century B . C . E . (Appendix A). The Chinese were using the magnetic compass around A . D . 1100, western Europeans by 1187, Arabs by 1220, and Scandinavians by 1300. Some speculate that the Chinese had discovered the orientating effect of magnetite, or

lodestone as early as the fourth century B . C . E . and that Chinese ships had reached the east coast of India for the rst time in 101 BCE using a navigational compass. Sir William Gilbert (15401603) made the rst investigation of terrestrial magnetism. In De Magnete (abbreviated title) he showed that the earths magnetic eld can be approximated by the eld of a permanent magnet lying in a general north-south direction near the earths rotational axis (Telford et al., 1990).

Manuscript received by the Editor May 19, 2005; revised manuscript received July 11, 2005; published online November 3, 2005. 1 Colorado School of Mines, 1500 Illinois Street, Golden, Colorado 80401-1887. E-mail: [email protected]; [email protected]. 2 U.S. Geological Survey, Box 25046, Federal Center MS 964, Denver, Colorado 80225. E-mail: [email protected]; [email protected]. 3 PRJ Inc., 12640 West Cedar Drive, Suite 100, Lakewood, Colorado 80228. E-mail: [email protected]. 4 Colorado School of Mines (retired), 1500 Illinois Street, Golden, Colorado 80401-1887. E-mail: [email protected]. 5 GEDCO, 1200, 8158th Avenue SW, Calgary, Alberta, T2P 3E2, Canada. E-mail: [email protected]. 6 Wintermoon Geotechnologies, Inc., 280 Columbine Street, Suite 301, Denver, Colorado 80206. E-mail: [email protected]. c 2005 Society of Exploration Geophysicists. All rights reserved. 33ND

34ND

Nabighian et al.

The attraction of compass needles to natural iron formations eventually led to their use as a prospecting tool by the 19th century. 1 As the association between magnetite and base metal deposits became better understood, demand for more sensitive instruments grew. Until World War II, these instruments were mostly specialized adaptations of the vertical compass (dipping needle), although instruments based on rotating coil inductors were also developed and used both for ground and airborne surveys. Victor Vacquier and his associates at Gulf Research and Development Company were key players in developing the rst uxgate magnetometer for use in airborne submarine detection during World War II (Reford and Sumner, 1964; Hanna, 1990). This instrument offered an order-of-magnitude improvement in sensitivity over previous designs. After the war, this improvement initiated a new era in the use of airborne magnetic surveys, both for the exploration industry and for government efforts to map regional geology at national scales (Hanna, 1990; Hood, 1990). Oceanographers quickly adapted early airborne magnetometers to marine use. In 1948, Lamont Geological Observatory borrowed a gimbal-mounted uxgate magnetometer from the U. S. Geological Survey and towed it across the Atlantic (Heezen et al., 1953). Scripps Institution of Oceanography began towing a similar instrument in late 1952 and in 1955 conducted the rst 2D marine magnetic survey off the coast of southern California (Mason, 1958). This now-famous marine magnetic survey showed a pattern of magnetic stripes offset by a fracture zone: the stripes were later attributed to seaoor spreading during periods of geomagnetic reversals (Vine and Mathews, 1963; Morley and Larochelle, 1964). As new instruments continued to be developed from the 1950s to 1970s, sensitivity was increased from around 1 nT for the proton precession magnetometer to 0.01 nT for alkalivapor magnetometers. With higher sensitivities, the error budget for aeromagnetic surveys became dominated by location accuracy, heading errors, temporal variations of the magnetic eld, and other external factors (e.g., Jensen, 1965). The development of magnetic gradiometer systems in the 1980s highlighted the problem of maneuver noise caused by both the ambient magnetic eld of the platform and by currents induced in the platform while moving in the earths magnetic eld (Hardwick, 1984). The availability of the Global Positioning System (GPS) by the early 1990s tremendously improved the location accuracy and thus the error budget of airborne surveys. At the same time, explorationists began to design airborne surveys so they could resolve subtle magnetic-eld variations such as those caused by intrasedimentary sources (see papers in Peirce et al., 1998). The higher resolution was achieved primarily by tightening line spacing and lowering the ight altitude. Today, high-resolution aeromagnetic (HRAM) surveys are considered industry standard, although exactly what ight specications constitute a high-resolution survey is ill dened. Typical exploration HRAM surveys have ight
1 There is a story that a Cretan shepherd named Magnes, while tending sheep on the slopes of Mount Ida, found that the nails of his boots were attracted to the ground. To nd the source of the attraction he dug up the ground and found stones that we now refer to as lodestones.

heights of 80150 m and line spacings of 250500 m (Millegan, 1998). Exploration surveys are generally own lower in Australia, at 6080 m above ground (e.g., Robson and Spencer, 1997), and even lower if acquired by the Geological Survey of Finland (3040 m ight height with 200m line spacing; http://www.gsf./aerogeo/eng0.htm). (Airspace regulations, urban development, or rugged terrain prevent such low-altitude ying in many places.) In contrast to these typical exploration specications, aeromagnetic studies that require high resolution of anomalies in plan view, such as those geared toward mapping complicated geology, usually entail uniform line spacings and ight heights, following the guidelines established by Reid (1980). Unmanned aerial systems are also becoming available and should be a cost-effective tool for acquiring low-altitude magnetic data in relatively unpopulated areas, although their eventual role in exploration is difcult to predict.

THE EARTHS MAGNETIC FIELD


The largest component (8090%) of the earths eld is believed to originate from convection of liquid iron in the earths outer core (Campbell, 1997), which is monitored and studied using a global network of magnetic observatories and various satellite magnetic surveys (Langel and Hinze, 1998). To a rst approximation, this eld is dipolar and has a strength of approximately 50 000 nT, but there are signicant additional spherical harmonic components up to about order 13. Furthermore, this eld changes slowly with time and is believed to undergo collapse, often followed by reversal, on a time scale of 100 000 years or so. Understanding the history of reversals, both as a pattern over time and in terms of decay and rebuilding of the primary dipole eld, is the focus of paleomagnetic studies [see Cox (1973) and McElhinny (1973) for excellent historical perspectives]. Although the crustal eld is the focus of exploration, magnetic elds external to the earth have a large effect on magnetic measurements and must be removed during data processing. These effects are the product of interaction between the global eld and magnetic elds associated with solar wind (Campbell, 1997). First, the earths eld is compressed on the sunward side, giving rise to a daily (diurnal) variation; at midlatitudes, diurnal variations are roughly 60 nT. Second, the interaction generates electrically charged particles that maintain a persistent ring current along the equator, called the equatorial electrojet. Instabilities in the ring current give rise to unpredictable magnetic-eld uctuations of tens of nT near the earths surface. Finally, near the poles, entrainment of charged particles along eld lines creates strong magnetic-eld uctuations during magnetic storms on time scales of a few hours and with amplitudes in excess of 200 nT. The remaining component of the earths eld originates in iron-bearing rocks near the earths surface where temperatures are sufciently low, i.e., less than about 580 C (the Curie temperature of magnetite). This region is conned to the upper 2030 km of the crust. The crustal eld, its relation to the distribution of magnetic minerals within the crust, and the information this relation provides about exploration targets are the primary subjects of the magnetic method in exploration.

Historical Development of Magnetic Method

35ND

APPLICATIONS OF MAGNETIC MEASUREMENTS


Magnetic measurements for exploration are acquired from the ground, in the air, on the ocean, in space, and down boreholes, covering a large range of scales and for a wide variety of purposes. Measurements acquired from all but the borehole platform focus on variations in the magnetic eld produced by lateral variations in the magnetization of the crust. Borehole measurements focus on vertical variations in the vicinity of the borehole.

Ground and airborne magnetic surveys


Ground and airborne magnetic surveys are used at just about every conceivable scale and for a wide range of purposes. In exploration, they historically have been employed chiey in the search for minerals. Regional and detailed magnetic surveys continue to be a primary mineral exploration tool in the search for diverse commodities, such as iron, base and precious metals, diamonds, molybdenum, and titanium. Historically, ground surveys and today primarily airborne surveys are used for the direct detection of mineralization such as iron oxidecoppergold (FeO-Cu-Au) deposits, skarns, massive suldes, and heavy mineral sands; for locating favorable host rocks or environments such as carbonatites, kimberlites, porphyritic intrusions, faulting, and hydrothermal alteration; and for general geologic mapping of prospective areas. Aeromagnetic surveys coupled with geologic insights were the primary tools in discovering the Far West Rand Goldelds gold system, one of the most productive systems in history (Roux, 1970). Kimberlites (the host rock for diamonds) are explored successfully using high-resolution aeromagnetic surveys (positive or negative anomalies, depending on magnetization contrasts) (Macnae, 1979; Keating, 1995; Power et al., 2004). Another economically important use of the magnetic method is the mapping of buried igneous bodies. These generally have higher susceptibilities than the rocks that they intrude, so it is often easy to map them in plan view. Commonly, the approximate 3D geometry of the body can also be determined. Because igneous bodies are frequently associated with mineralization, a magnetic interpretation can be the rst step in nding areas favorable for the existence of a mineral deposit. In sedimentary basins, buried igneous bodies may have destroyed hydrocarbon deposits in their immediate vicinity, their seismic signature can be mistaken for a sedimentary structure (Chapin et al., 1998), or their orientation is important in understanding structural traps in an area (e.g., the Eocene Lethbridge dikes in southern Alberta, Canada). Igneous bodies can also form structural traps for subsequent hydrocarbon generation. For example, brecciated igneous rocks [e.g., Eagle Springs Field, Nevada; Fabero Field, Mexico; Badejo and Linguado Fields, Brazil; Jatibarang Field, Indonesia; and reported potential in the Taranaki Basin, New Zealand, all cited in Batchelor and Gutmanis (2002)] are known to be reservoirs. For regional exploration, magnetic measurements are important for understanding the tectonic setting. For example, continental terrane boundaries are commonly recognized by the contrast in magnetic fabric across the line of contact (e.g., Ross et al., 1994; examples in Finn, 2002). Such regional interpretations require continent-scale magnetic databases. Development of these databases commonly involves merging

numerous individual aeromagnetic surveys with highly variable specications and quality. Such efforts have been ongoing for decades. For example, two major compilations have been completed for North America (Committee for the Magnetic Anomaly Map of North America, 1987; North American Magnetic Anomaly Group, 2002), which updated earlier efforts for the United States (Zietz, 1982) and Canada (Teskey et al., 1993). A comprehensive, near-global compilation of magnetic data outside the United States, Canada, Australia, and the Arctic regions was undertaken in a series of projects by the University of Leeds, the International Institute for Geo-Information Science and Earth Observation (ITC), and commercial partners (Fairhead et al., 1997). Several countries (e.g., Australia, Canada, Finland, Sweden, and Norway) have vigorous government programs to develop countrywide, modern, high-resolution aeromagnetic databases, which include data acquisition and merging of data from individual surveys. These efforts have been successful in promoting mineral exploration and facilitating ore deposit discoveries. The study of basin structure is an important economic application of magnetic surveys, especially in oil and gas exploration. For the most part, basin ll typically has a much lower susceptibility than the crystalline basement. Thus, it is commonly possible to estimate the depth to basement and, under favorable circumstances, quantitatively map basement structures, such as faults and horst blocks (e.g., Prieto and Morton, 2003). Since structure in shallower sections often lies conformably over the basement, at least to some depth, and faulting in shallower sections is often controlled by reactivation of basement faults, it is often possible to identify structures favorable to hydrocarbon accumulation from basement interpretation. With the advent of HRAM surveys and the subnanotesla resolution they offer, it is now possible to map intrasedimentary faults by identifying their small and complex magnetic anomalies that occur where there are marker beds containing greater than average quantities of magnetite. Displacement of these marker beds generates modest (a few tenths to about 10 nT at 150 m elevation) anomalies that can be used to trace corresponding fault systems. The complex nature of these anomalies is illustrated in a case history in the Albuquerque Basin (see Case Histories section) where the magnetizations are high enough to clearly understand the relationships of bed thickness, offset, fault dip, etc. (Grauch et al., 2001). In hydrocarbon exploration, such techniques can be used to help correlate complex fault systems for exploration (Spaid-Reitz and Eick, 1998; Peirce et al., 1999) or for reservoir development (see Case Histories section; Goussev et al., 2004). In areas where beds carrying a magnetic signature dip at a signicant angle, a good magnetic survey can be used to map surface geology very precisely (e.g., Abaco and Lawton, 2003). The magnetic method has thus expanded from its initial use solely as a tool for nding iron ore to a common tool used in exploration for minerals, hydrocarbons, ground water, and geothermal resources. The method is also widely used in additional applications such as studies focused on water-resource assessment (Smith and Pratt, 2003; Blakely et al., 2000a), environmental contamination issues (Smith et al., 2000), seismic hazards (Blakely et al., 2000b; Saltus et al., 2001; Langenheim et al., 2004), park stewardship (Finn and Morgan,

36ND

Nabighian et al.

2002), geothermal resources (Smith et al., 2002), volcanorelated landslide hazards (Finn et al., 2001), regional and local geologic mapping (Finn, 2002), mapping of unexploded ordnances (Butler, 2001), locating buried pipelines (McConnell et al., 1999), archaeological mapping (Tsokas and Papazachos, 1992), and delineating impact structures (Campos-Enriquez et al., 1996; Goussev et al., 2003), which can sometimes be of economic importance (Mazur et al., 2000).

hypothesis. The two sets of ages matched very well, and the Vine-Matthews-Morley model was generally accepted; plate tectonics became a new paradigm in earth sciences. Marine magnetic measurements also are routinely used for normal exploration applications, although not in the volume of aeromagnetic work.

Satellite magnetic measurements


Magnetic surveying entered the space age in 1964 with the launch of a scalar magnetometer on the Cosmos 49 mission. Subsequently, the POGO suite of polar-orbiting satellites, OGO-2, OGO-4, and OGO-6, conducted scalar measurements over a seven-year period. Magsat, own in 1979 1980 in polar orbit, carried the rst vector magnetometer. Satellite DE-1 collected vector information as well, in spite of its highly elliptical orbit (500 km to 22,000 km perigee and apogee, respectively). Recent launches in 1999 and 2000 of the Oersted (Olsen et al., 2000), CHAMP (Reigber et al., 2002), and Oersted-2/SAC-C missions were equipped with more sensitive scalar and vector magnetometers and have furthered our understanding of the core, crustal, and external elds. Since 1970, satellite measurements of the geomagnetic eld have been used to better model the dynamics of the core eld and its secular variation. These models have been incorporated into the International Geomagnetic Reference Field (IGRF) (see Magnetic Data Processing section) to provide more accurate information for processing exploration-quality magnetic surveys. Satellite magnetometers have provided new insights into the external magnetic eld as well. Although explorationists are not using satellite magnetic measurements for prospect generation and for mapping of the crustal eld, we are reaping great benets from their impact on the core eld model and its regular ve-year updates. In recent years, the magnetic method has formed an important component of extraterrestrial exploration (see the special section in the August 2003 issue of T HE L EADING E DGE ). From planetary scales to areas of a few meters, the magnetic method has had a role to play, in some cases a decisive one. It could be argued that no other geophysical method has such a broad range of applicability or offers such economical information.

Borehole magnetic surveys


Borehole measurements of magnetic susceptibility and of the three orthogonal components of the magnetic eld began in the early 1950s (Broding et al., 1952). Both types of measurements can be used to determine rock magnetic properties, which aids in geologic correlation between wells. However, magnetic-eld measurements in boreholes also can be used to determine both location and orientation of magnetic bodies missed by previous drilling. Levanto (1959) describes the use of three-component uxgate magnetometers to determine the extension of magnetic ore bodies. Interpretation of borehole magnetic surveys was originally accomplished graphically by plotting the eld lines along the borehole, extrapolating them outside the borehole, and looking for areas of eld-line convergence. Least-square techniques were also employed to determine the parameters of the magnetic body (Silva and Hohmann, 1981). Today, acquisition of borehole magnetic surveys is not common practice, perhaps owing to the expense required in accurately determining borehole azimuth and dip.

Marine magnetic surveys


Marine magnetic measurements began at Lamont in the late 1940s (Oreskes, 2001) and led to the development of the Vine-Matthews-Morley model of seaoor spreading (Dietz, 1961; Hess, 1962; Vine and Matthews, 1963; and Morley and Larochelle, 1964). The name of this model has been updated by consensus (Vine, 2001) to recognize Larry Morley of the Geological Survey of Canada as the independent developer of the theory of seaoor spreading (Morleys original paper, which was rejected in early 1963, is reproduced in Morley, 2001). In fact, these marine magnetic measurements were a major factor in the acceptance of both the plate-tectonic theory and of the dynamo theory of generation of the earths core eld. The seaoor spreading model is based on the concept that the seaoor is magnetized either positively or negatively, depending on the polarity epoch of the earths magnetic eld. New seaoor is created at mid-ocean ridges and becomes part of oceanic plates moving away from the spreading center. Thus, magnetic anomalies along a section transverse to the spreading center show a regular pattern of highs and lows (stripes) often symmetric about the spreading center that can be calibrated in age to the geomagnetic timescale (this timescale was new and unproven in 1963; a more recent compilation accompanies the Geological Society of Americas 1999 geologic timescale). Leg 3 of the original Deep Sea Drilling Project (Maxwell and von Herzen et al., 1970) was designed to test the theory of sea-oor spreading by comparing the paleontological ages of the oldest sediments in the South Atlantic Ocean to the ages predicted by the seaoor spreading

MAGNETIC INSTRUMENTATION
Historical instruments
The Swedish mining compass was one of the earliest magnetic prospecting instruments. Developed in the midnineteenth century, it consisted of a light needle suspended in such a way as to allow it to move in both horizontal and vertical directions. An improved version, the American mining compass, was developed around 1860. These were the rst in a class of so-called dipping-needle instruments with automatic meridian adjustment. Although still in use these instruments were soon replaced by earth inductors, which could measure both the inclination and the various components of the earths magnetic eld from the voltage induced in a rotating coil. In 1936, Logachev (1946) used such a device with a sensitivity of about 1000 nT over the Kursk iron-ore deposit (Reford and Sumner, 1964). Soon after, the Schmidt vertical magnetometer was developed, which could measure the vertical component of the

Historical Development of Magnetic Method

37ND

earths magnetic eld using a magnetic system (rhomb-shaped needle) oriented at a right angle to the magnetic meridian; it measured the system dip through a mirror attached to the needle and an autocollimation telescope system. The vertical magnetometer was followed by the Schmidt horizontal magnetometer, which measured the horizontal component of the earths eld. Both instruments had an accuracy of 1020 nT. The Schmidt magnetometers came to be known as AskaniaSchmidt magnetometers. In 1910, Edelman designed a vertical balance to be used in a balloon (Heiland, 1935). In 1946, a vertical-intensity magnetometer of the earth-inductor type was introduced by Lundberg (1947) for helicopter surveys, and a vibrating coil variety of the earth-inductor magnetometer was developed for both airborne and shipborne use (Frowe, 1948). A complete description of early magnetic prospecting instruments and their uses can be found in Heiland (1940), Jakosky (1950), and Reford and Sumner (1964).

Proton precession magnetometer


Proton precession magnetometers were introduced in the mid-1950s, and by the mid-1960s had supplanted uxgate magnetometers for almost all exploration applications. Proton precession magnetometers do not require orientation, a great advantage over earlier devices. The proton precession magnetometer is based on the splitting of nuclear spin states into substates in the presence of an ambient magnetic eld by an amount proportional to the intensity of the eld and a proportionality factor (the nuclear gyromagnetic ratio), which depends only on fundamental physical constants. The sensor consists of a quantity of material with odd nuclear spin, almost always hydrogen. The actual sensor lling is usually charcoal lighter uid, decane, benzene, or, if necessary, even water. The sensor is surrounded by a coil through which a dc current is applied. This induces transitions to the higher energy of two nuclear spin substates. The current is then turned off and used to detect the elds associated with the transition back to the lower of the spin substates. This transition emits an electromagnetic eld whose frequency is proportional to the earths eld intensity, around 2 kHz. A frequency counter is then used to measure the eld strength. The full treatment of the physics behind proton precession magnetometers (Hall, 1962) is usually explained intuitively in textbooks by envisioning the transition between nuclear substates as a precession of the nuclear magnetic moments around the earths eld direction at an angular frequency proportional to the intensity of the eld. The proton precession magnetometer has a number of advantages: it is rugged, simple, has essentially no intrinsic heading error, and does not require an orienting platform. However, to obtain reasonable signal strength, a fairly large quantity of sensor liquid and a large coil are required, making the instrument somewhat heavy, bulky, and power hungry. Furthermore, because a signicant polarizing time is required and because the output signal is only around 2 kHz, the sample rate is somewhat limited if reasonable sensitivity is required. The best airborne units (now out of production) had a sensitivity of 0.05 nT at 2 Hz. More typical values would be 0.1 nT at 0.2 Hz for portable instruments still in use. A variant (the Overhauser magnetometer) uses radiofrequency excitation and effectively displays continuous oscillation, which can be sampled at 5 Hz with resolution of 0.01 nT. The Overhauser variant also offers the lowest power drain of any modern magnetometer, a small sensor head, and minimal heading error. It is widely used in subsea magnetometers and is also used for airborne and ground survey work, often in gradient arrays.

Fluxgate magnetometer
The uxgate magnetometer was developed during World War II for airborne antisubmarine warfare applications; after the war, it was immediately adopted for exploration geophysics and remained the primary airborne instrument until the proton precession magnetometer was introduced in the 1960s. Fluxgate magnetometers today have two major applications. In airborne systems they are used in a strap-down (nonoriented) conguration to perform heading corrections by measuring the altitude of the aircraft in the earths eld. They are also the dominant instrument in downhole applications because of their small size, ruggedness, and ability to tolerate high temperatures. The basic elements of a uxgate magnetometer are two matched cores of highly permeable material, typically ferrite, with primary and secondary windings around each core. The primary windings are connected in series but with opposite orientations and are driven by a 501000-Hz current which saturates the cores in opposite directions, twice per cycle. The secondary coils are connected to a differential amplier to measure the difference between the magnetic eld produced in the two cores. This signal is asymmetrical because of the ambient magnetic eld along the core axis, producing a spike at twice the input frequency whose amplitude is proportional (for small imbalances) to the eld along the core axis. A detailed discussion of the uxgate magnetometer can be found in Telford et al. (1990). Typically, uxgate elements are packaged into sets of three core pairs with orthogonal axes, so all three components of the earths eld can be measured. The resolution of a uxgate system is dependent on the accuracy with which the cores and windings can be matched, hysteresis in the cores, and related effects; nevertheless, uxgate units with better than 1 nT sensitivity are widely available. They are rugged, lightweight, and can be operated at relatively high measurement rates. Their major disadvantage for airborne applications is that because they are component instruments, they must be oriented. At least until recently, the accuracy of uxgate measurements was limited by the stability of the gyro tables on which they were mounted.

Alkali vapor magnetometer


Alkali vapor magnetometers, with sensitivities around 0.01 nT and sample rates of 10 Hz, appeared in laboratories about the same time that proton precession magnetometers became popular eld instruments. Because they were more fragile than proton precession magnetometers, and because the increased sensitivity was of marginal value, their use as eld instruments was mostly restricted to gradiometers until the late 1970s. Today, alkali vapor magnetometers are the dominant instrument used for magnetic

38ND

Nabighian et al.

surveys, although some proton precession instruments are still in use for ground surveys, and uxgates are used for borehole surveys. The operating principle and the actual construction of alkali vapor magnetometers are somewhat complex. However, since they have become the dominant type in current airborne, shipborne, and ground exploration, a summary explanation of their operation is appropriate. The sensing medium is an alkali vapor consisting of atoms randomly distributed between two different atomic-energy levels, separated by energy equivalent to a visible frequency. In the presence of a magnetic eld, the most stable energy level is split (Zeeman splitting) by an amount proportional to the magnitude of the eld. For ambient elds of around 50 000 nT, the splitting energy will correspond to a frequency in the range of a few hundred kHz, i.e., the AM radio band. By shining light of the correct frequency through a vapor of a single-valence atom such as cesium or potassium, all of the electrons are forced into the higher-energy component of the split state (optical pumping). When this absorption is complete, the glass cell in which the vapor is contained becomes transparent because there are no further electrons to absorb the pumping radiation. Now, a radio-frequency eld is applied to this cell. If the eld is of exactly the right frequency, the electrons are redistributed back to the lower level, and the cell becomes opaque. The correct frequency depends on the ambient magnetic-eld strength, so a swept-frequency eld is applied, and the precise frequency at which opacity occurs is used to derive the ambient-eld intensity. Alkali vapor instruments have excellent sensitivity, better than 0.01 nT. Because the frequency can be swept rapidly, 10-Hz sample rates are typical, and considerably higher ones are possible. These features account for the overwhelming popularity of this design. In addition, alkali vapor magnetometers are built to be lightweight and compact. The less desirable features are the fragility of the glass envelope and an intrinsic heading error. A good discussion of alkali vapor magnetometers can be found in Telford et al. (1990).

pearance of high-temperature superconductors, it has been anticipated that liquid nitrogen could be used as the cooling uid, which is much easier to manufacture and handle than liquid helium. However, high-temperature SQUIDs have only recently appeared on the market. They have somewhat lower sensitivities than the liquid-helium instruments, primarily because the 1/f noise in the amplier electronics is higher, but this should not be a major concern for most applications. It seems likely that the use of SQUIDs may increase in the near future, as shown by recent surveys own in gradient mode for mineral surveys. Stuart (1972) gives a comprehensive review of magnetic instruments used in geophysical applications. Grivet and Malnar (1967) give detailed discussions of instruments based on Zeeman splitting, including proton precession magnetometers, alkali vapor magnetometers, and related designs, not limited to geophysical instruments.

ROCK-MAGNETIC PROPERTIES
In geologic interpretation of magnetic data, knowledge of rock-magnetic properties for a particular study area requires an understanding of both magnetic susceptibility and remanent magnetization. Seventy-ve years ago, studies were already underway to explain the geologic factors inuencing rock-magnetic properties that produce magnetic anomalies (Slichter, 1929; Stearn, 1929a). Factors inuencing rockmagnetic properties for various rock types are summarized by Haggerty (1979), McIntyre (1980), Clark (1983, 1997), Bath and Jahren (1984), Grant (1985), Reynolds et al. (1990a), and Clark and Emerson (1991). The Norwegian, Swedish, and Finnish surveys have been amassing large amounts of rockproperty information in conjunction with their national geophysical programs. Several studies have focused on developing classication schemes based on the statistical correlations between rock types and these petrophysical measurements (Korhonen et al., 2003). Less progress has been made in understanding how information on magnetic properties measured from hand samples can be transferred to scales more appropriate for aeromagnetic interpretation. Reford and Sumner (1964) and Clark (1983) discussed how the high variability of properties measured in hand samples contradicts the apparent homogeneity in the bulk effects of large bodies at the scale of aeromagnetic studies. Understanding this contradiction remains elusive, especially in understanding sedimentary sources. Improved understanding may result from case studies that directly investigate the relationship between magnetic anomalies, rock properties, and geology (Abaco and Lawton, 2003; Davies et al., 2004). The importance of sedimentary sources of magnetic anomalies was the subject of considerable discussion before the end of World War II (e.g., Jenny, 1936; Wantland, 1944). Magnetic anomalies produced by glacial till were also widely known (summarized in Gay, 2004). However, experience with the relatively low resolution of the early aeromagnetic data allowed workers to effectively ignore their effects (Steenland, 1965; Nettleton, 1971), giving rise to the misconception that sediments are nonmagnetic. As data resolution increased, magnetic anomalies arising from sedimentary sources were again recognized (Grant, 1972). This recognition gained prominence in the 1980s, when studies were initiated to test

Superconducting quantum interference device (SQUID) magnetometer


The fact that a persistent current can exist in a superconducting loop has been known since the 1930s. This current is inherently insensitive to the ambient magnetic eld; one of the main features of superconductors is that they expel external elds. Josephson (1962) showed that persistent currents could be maintained across small gaps in the superconducting loop and that the currents across the gap are sensitive to the magnetic ux passing through the loop. Flux is the product of the component of the magnetic eld perpendicular to the loop and the loop area. Current changes can be monitored by a normal resonant circuit and used to obtain component eld values. Instruments of this type are called superconducting quantum interference devices (SQUIDs) (Weinstock and Overton, 1981). SQUIDs have not been widely used in magnetic eld applications, although they have been used extensively in magnetotelluric and paleomagnetic studies and, recently, in both ground and airborne EM surveys as magnetic component sensors. The main reason is the need for cryogenic supplies, which reduces the mobility of SQUID magnetometers. Since the ap-

Historical Development of Magnetic Method

39ND

for magnetic effects related to hydrocarbon seepage. These and subsequent studies demonstrated that magnetization capable of producing aeromagnetic anomalies in clastic sedimentary rocks and sediments arise from the abundance of detrital magnetite (Reynolds, Rosenbaum et al., 1990, 1991; Gay and Hawley, 1991; Gunn, 1997, 1998; Wilson et al., 1997; Grauch et al., 2001; Abaco and Lawton, 2003), remanence residing in iron suldes that replaced the original detrital material (Reynolds, Rosenbaum et al., 1990, 1991), or possibly some other kind of remanence (Phillips et al., 1998). A recent study of the Edwards aquifer in central Texas has revealed, in a low-level helicopter survey, that carbonates may also contain enough detrital magnetite to produce magnetic anomalies at faults (Smith and Pratt, 2003). At local scales, magnetite can be produced by microbial activity (Machel and Burton, 1991) or destroyed by suldization (Goldhaber and Reynolds, 1991) in processes related to hydrocarbon migration, although it is still debated whether this effect can be detected from airborne surveys (Gay, 1992; Reynolds et al., 1990; Millegan, 1998; Stone et al., 2004). Morgan (1998) postulates that weak aeromagnetic lows in oil elds of the Irish Sea are caused by complex migration and mixing of uids with hydrocarbons that reduced the magnetization of the host sandstones. The importance of remanent magnetization in magnetic interpretation has been recognized by many previous workers (see references in Zietz and Andreasen, 1967). To simplify analytical methods, remanent magnetization has been commonly neglected or assumed to be collinear with the induced component. Bath (1968) considered remanent and induced components within 25 of each other to be collinear for practical purposes. Although valid in many geologic situations, a common misconception is that only mac igneous rocks have high remanence. Several rock-magnetic and aeromagnetic studies have shown that remanence can be very high in felsic ash-ow tuffs (Bath, 1968; Rosenbaum and Snyder, 1985; Reynolds, Rosenbaum et al., 1990a; Grauch et al., 1999; Finn and Morgan, 2002).

cable, thereby reducing the errors caused by the magnetic eld of the vehicle. Fixed-wing airborne operations, on the other hand, usually use rigid magnetometer installations, such as a stinger protruding from the rear of the aircraft. Fixed installations offer better control of the sensor location and have been the preferred conguration since the early days of airborne data acquisition. The eld of the aircraft is signicant unless the aircraft has been extensively rebuilt. This problem is overcome by compensating for the platform eld. Error models were developed during World War II but not published until much later (Leliak, 1961). Early compensation methods consisted of attaching bar magnets and strips of Permalloy near the sensor to approximately cancel the aircraft eld (EG&G Geometrics, 1970s). Later, feedback compensators were developed for military use (CAE, 1968, Study guide for the 9-term compensator: Tech. Doc. TD-2501, as cited in Hardwick, 1984). However, Hardwick (1984) pointed out in a landmark paper that these were unsuitable for geophysical use because they were limited to the frequency band appropriate for submarine detection. Hardwick (1984) also noted that good compensation was crucial to the usefulness of the magnetic gradiometer systems then being built. He described a software compensation system that was eventually commercialized and is now in widespread use, even in single-sensor systems. Alternatives, such as postprocessing compensation, are also available. It is fair to say that, along with the introduction of GPS, the use of these more sophisticated compensation models has produced the largest improvement in data quality over the past 20 years.

Global eld models


The main component of the measured magnetic eld originates from the magnetic dynamo in the earths outer core (Campbell, 1997). This eld is primarily dipolar, with amplitude of around 50 000 nT, but spherical harmonic terms up to about order 13 are signicant. Since the core eld is almost always much larger than that of the crustal geology, and since it has a signicant gradient in many parts of the world, it is desirable to remove a model of the global eld from the data before further processing; this can be done as soon as all positioning errors are corrected. The model most widely used today is the International Geomagnetic Reference Field (IGRF, Maus and Macmillan, 2005). It was established in 1968 and became widely used with the availability of digital data in the mid-1970s (Reford, 1980). In 1981 the IGRF was modied in order to be continuous for all dates after 1944 (Peddie, 1982, 1983; Paterson and Reeves, 1985; Langel, 1992). Today, the IGRF is updated every ve years and includes coefcients for predicting the core eld into the near future. Coefcients are available for the time period 1900 through 2005 (Barton, 1997; Macmillan et al., 2003). In practice today, the IGRF is calculated for every data point before any further processing. Prior to GPS navigation, however, it was common practice to level surveys rst, then remove a trend based on the best t, either to the data or to a few IGRF values. For many of the earliest analog surveys, an arbitrary (sometimes unspecied) constant was subtracted from the measured data solely as a matter of convenience before contouring.

MAGNETIC DATA PROCESSING


Data processing includes everything done to the data between acquisition and the creation of an interpretable prole, map, or digital data set. Standard steps in the reduction of aeromagnetic data, some of which also apply to marine and ground data, include removal of heading error and lag, compensation for errors caused by the magnetic eld of the platform, the removal of the effects of time-varying external elds, removal of the International Geomagnetic Reference Field, leveling using tie-lines, microleveling, and gridding. One comprehensive reference that summarizes most aspects of magnetic data processing is Blakely (1995).

Compensation
All moving-platform magnetic measurements are subject to errors caused by the magnetic eld of the platform, whether from in situ magnetic properties of the platform or from currents induced in the platform while moving in the earths magnetic eld. In shipborne and helicopter surveys, these effects are typically minimized by mounting the sensor on a long tow

40ND

Nabighian et al.

In the future, the IGRF is likely to be supplanted by the Comprehensive Model (CM), which does a much better job of modeling time-varying elds from a variety of sources (Sabaka et al., 2002, 2004; Ravat et al., 2003).

External (time-varying) eld removal


Ground-based and airborne surveys generally include a stationary magnetometer that simultaneously measures the stationary, time-varying magnetic eld for later subtraction from the survey data (Hoylman, 1961; Whitham and Niblett, 1961; Morley, 1963; Reford and Sumner, 1964; Paterson and Reeves, 1985). There is still considerable debate on how many base stations are needed to adequately sample the spatial variations of the external eld for larger surveys or when the survey area is at a considerable distance from the base of operations. At sea, it is generally not possible to have a base-station magnetometer in the survey area, and the problem is either ignored or measurements are made in a gradient mode. The measurement of multisensor gradiometer data can reduce the need for a base station because the common external signal at the two sensors is removed by the differencing process, but recovery of the total eld data from the gradiometer data can be difcult (Breiner, 1981; Hansen, 1984; Paterson and Reeves, 1985). A method to t distant base signals to the eld signal in order to remove time-varying effects was proposed by OConnel (2001) using a variable time-shift cross-correlator. The leveling of surveys using tie-lines was originally developed as an alternative to the use of base-station data (Whitham and Niblett, 1961; Reford and Sumner, 1964; Mittal, 1984; Paterson and Reeves, 1985) but is now a standard step after base-station correction. The purpose of leveling today is to minimize residual differences in level between adjacent lines and long-wavelength errors along lines that inevitably remain after compensation and correction for external eld variations by base station subtraction. These residual longwavelength effects, even if small, can be visually distracting, particularly on image displays. A set of tie-lines perpendicular to the main survey lines is normally acquired for leveling. The tie-line spacing is generally considerably greater than that of the main survey lines, although 1:1 ratios have been used where geologic features lack a dominant strike. The differences in eld values at the intersections of the survey and tie-lines are calculated and corrections are applied to minimize these differences. A number of different strategies for computing these corrections are in use. Perhaps the most common is to calculate a constant correction for all lines by least-squares methods, sometimes augmented to a low-order polynomial. Other algorithms regard the tie-lines as xed and adjust only the survey lines. All of these strategies are empirical, and no one method performs best under all circumstances.

is that we know exactly where our data are located horizontally at the time of measurement, but we can only guess at the nal observation surface after leveling and microleveling have been applied. As in tie-line leveling, a number of microleveling algorithms are in use that differ in detail but all rely on the principle of removing the corrugation effects from a grid and using the decorrugated grid to correct the long-wavelength errors on the prole data. Because microleveling uses the grid in an essential way, it effectively erases the small corrugations. However, it also largely obliterates any features that actually trend along the survey lines.

Deculturing
Cultural anomalies are a serious problem in the geologic interpretation of airborne magnetic data, especially modern HRAM surveys that typically y low above cultural sources. Many man-made structures (e.g., wells, pipelines, railroads, bridges, steel towers, and commercial buildings) are ferrous and so create sharp anomalies of tens to hundreds of nanoteslas. Cultural anomalies are often much larger in magnitude than the geologic anomalies of interest. Moreover, their shapes are effectively spikes with broadband frequency responses, making them difcult if not impossible to remove with linear lters. Several approaches have been developed for cultural editing. The utility of each approach depends on the magnitude and type of cultural anomalies present. One approach is to avoid ying low-level surveys to suppress the cultural signal (Balsley, 1952), but this may diminish useful geological signals from shallow sources and does not eliminate noise spikes. The Naudy lter (Naudy and Dreyer, 1968) uses nonlinear lters to solve the problem. Hassan et al. (1998) discuss the relative merits and limitations of manual editing on a proleby-prole basis, of semiautomatic ltering using a Naudytype lter, and of fully automatic ltering using neural nets. Hassan and Peirce (2005) present an improved approach to manual editing for situations where good digital databases of existing culture are available. Wavelet ltering is another method that offers promise in terms of developing more effective automated techniques, but there will always be a need to manually oversee the results to prevent the removal of an important shallow geological signal. For special cases where the source structure can be modeled, such as for well casings (Frischknecht et al., 1983; Boardman, 1985), it is possible to design effective automatic removal techniques (Dickinson, 1986; Pearson, 1996). However, all of these methods depend on recognition of a known anomaly signature. In general, it is not possible to construct such models; for example, the anomaly of a town is an aggregate of anomalies from many man-made sources, clearly beyond reasonable modeling capabilities. In such cases, it is necessary to resort to deleting the cultural anomaly from the data using ight-path video or digital cultural data as a guide and replacing it with interpolated values (Hassan and Peirce, 2005).

Microleveling
Leveling, as described in the previous subsection, generally produces acceptable results for contour map displays, but small corrugations generally can still be seen on images. To suppress these, microleveling or decorrugation is applied (Hogg, 1979; Paterson and Reeves, 1985; Urquhart, 1988; Minty, 1991). One of the ironies of modern GPS navigation

Gridding
Gridding of ight-line data is another area of continuing research. Because the density of data is so much greater along

Historical Development of Magnetic Method

41ND

the ight-line direction than across ight lines, early efforts concentrated on bidirectional interpolation (Bhattacharyya, 1969). Minimum curvature (Briggs, 1974) has proved to be a popular gridding algorithm for unaliased data. Present surveys are planned to minimize the amount of cross-track aliasing, following the guidelines of Reid (1980). However, for reconnaissance surveys and for xed-wing surveys in areas of rough terrain, there will always be residual cross-track aliasing. An extension of the minimum-curvature-gridding algorithm designed to address this problem has recently been developed by OConnell et al. (2005). Another way to address the issue of ight-line data density is to use kriging with an anisotropic covariance function (Hansen, 1993). More exotic approaches use equivalent sources to produce a grid with the characteristics of a potential eld (Cordell, 1992; Mendonc a and Silva, 1994, 1995); others employ fractals (Keating, 1993) and wavelets (Ridsdill-Smith, 2000). The evolution of magnetic data processing can be characterized more by the things we no longer need to discuss than those mentioned above. Included in the dustbin of earlier concerns are contouring algorithms, display hardware, cameraand map-based navigation, radio navigation systems, and base stations versus control lines.

served data. The specic goals of these lters vary, depending on the situation. The general purpose is to enhance anomalies of interest and/or to gain some preliminary information on source location or magnetization. Most of these methods have a long history, preceding the computer age. Modern computing power has increased their efciency and applicability tremendously, especially in the face of the ever-increasing quantity of digital data associated with modern airborne surveys. Most lter and interpretation techniques are applicable to both gravity and magnetic data. As such, it is common, when applicable, to reference a paper describing a technique for ltering magnetic data when processing gravity data and vice versa.

Regional-residual separation
Regional-residual separation is a crucial step in the interpretation of magnetic data for mining or unexplored ordnance (UXO) applications but less so for petroleum applications because the depth range of hydrocarbon exploration extends throughout the sedimentary section. Historically, this problem was approached either by using a simple graphical approach (manually selecting data points to represent a smooth regional eld) or by using various mathematical tools to obtain the regional eld. This problem has been extensively treated for gravity data (Nabighian et al., 2005), and the proposed techniques apply equally well to magnetic investigations. The graphical approach was initially limited to analyzing prole data and, to a lesser extent, gridded data. The earliest nongraphical approach considered the regional eld at a point to be the average of observed values around a circle centered on the point; the residual eld was simply the difference between this average value and the value observed at the central point (Grifn, 1949). Henderson and Zietz (1949) and Roy (1958) showed that this method was equivalent to calculating the second vertical derivative except for a constant factor. Agocs (1951) proposed using a least-squares polynomial t to data to determine the regional eld, an approach criticized by Skeels (1967) since the anomalies themselves will affect somewhat the determined regional. Zurueh (1967) proposed using two-dimensional linear wavelength lters of different cutoff wavelengths. This method was further expanded by Agarwal and Kanasewich (1971), who also used a crosscorrelation function to obtain trend directions from magnetic data. A comprehensive discussion of application of Fourier transforms to potential eld data can be found in Gunn (1975). Syberg (1972a) described a matched-lter method for separating the residual eld from the regional eld. A method based on frequency-domain Wiener ltering for gravity data was proposed by Pawlowski and Hansen (1990) that is equally applicable to magnetic data. Matched lters and Wiener lters have much in common with other linear bandpass lters but have the distinct advantage of being optimal for a class of geologic models. Based on experience, however, it seems that signicantly better results can be obtained using appropriate statistical geologic models than by attempting to adjust band the parameters of band-pass lter manually. Li and Oldenburg (1998a) use a 3D magnetic inversion algorithm to invert the data over a large area in order to construct a regional susceptibility distribution from which a

ACCOUNTING FOR MAGNETIC TERRAIN


Rugged terrain poses a number of difculties for data acquisition, processing, analysis, and interpretation (Hinze, 1985a). Data processing and acquisition errors caused by difculties in ying over rugged terrain have been largely overcome with the advent of preplanned draped surfaces and GPS navigation, but these steps do not account for the effects of magnetic sources in the terrain itself. Magnetic anomalies produced by the magnetic effects of rocks that form topography are called topographic anomalies or magnetic terrain effects (Allingham, 1964; Grauch, 1987) and should not be confused with the effects produced by irregular terrain clearance. They are easily recognized by the strong correlation of the anomaly shapes to topography (Blakely and Grauch, 1983). Magnetic terrain effects can severely mask the signatures of underlying sources, as demonstrated by Grauch and Cordell (1987). Many workers have attempted to remove or minimize magnetic terrain effects by using some form of ltering or modeling scheme (summarized in Grauch, 1987). Unlike gravity terrain corrections, however, these attempts have been successful only in favorable conditions. In more recent studies, workers have used rugged terrain to their advantage. In a basaltic volcanic eld within the Rio Grande rift, for example, high-amplitude negative and positive anomalies correlate with topography and helped geologists distinguish similarlooking basalts with different ages (Thompson et al., 2002). At Mt. Rainier, the lack of magnetic anomalies correlating with terrain helped estimate the volume of hydrothermally altered material available for potential landslides (Finn et al., 2001).

MAGNETIC DATA FILTERING


The beginning stages of magnetic data interpretation generally involve the application of mathematical lters to ob-

42ND

Nabighian et al.

regional eld can then be calculated. In certain aspects, this method is a magnetic application of a gravity interpretation technique known as stripping (Hammer, 1963). Spector and Grant (1970) analyzed the shape of power spectra calculated from observed data. Clear breaks between low- and highfrequency components of the spectrum were used to design either band-pass or matched lters. In hydrocarbon exploration, this is the most common approach to separating different depth ranges of interest based on their frequency content. This approach is discussed in a modern context by Guspi and Introcaso (2000). The existence of so many techniques for regional-residual separation proves that there are still some unresolved problems in this area. There is no single right answer for how to highlight ones target of interest.

solving an inverse problem in which a global objective function is minimized subject to tting the observed data. All of these techniques assume that the directions of magnetization and ambient eld are invariant over the entire survey area. While this is appropriate for many studies, it is not appropriate for continent-scale studies, over which the earths magnetic-eld direction varies signicantly, or in geologic environments, where remanent magnetization is important and variable. Arkani-Hamed (1988) addressed the former problem with an equivalent-layer scheme, in which variations in magnetization and ambient-eld directions were treated as perturbations on uniform directions.

Pseudogravity transformation
Poissons relation shows that gravity and magnetic anomalies caused by a uniformly dense, uniformly magnetized body are related by a rst derivative. Baranov (1957) used this principle to transform an observed magnetic anomaly into the gravity anomaly that would be observed if the distribution of magnetization were replaced with a proportional density distribution. Baranov called the transformed data pseudogravity, although the pseudogravity anomaly is equivalent to the magnetic potential. The pseudogravity transformation is most commonly used as an interim step to several other edge-detection or depth-estimation techniques or in comparing with observed gravity anomalies. Since calculation of the pseudogravity anomaly involves a reduction to the pole followed by a vertical integration, it is affected by the same instabilities that were present in calculating the RTP eld. In addition, the pseudogravity transformation amplies long wavelengths, and so grids must be expanded carefully before processing to minimize amplication of long-wavelength noise.

Reduction to pole (RTP)


Like a gravity anomaly, the shape of a magnetic anomaly depends on the shape of the causative body. But unlike a gravity anomaly, a magnetic anomaly also depends on the inclination and declination of the bodys magnetization, the inclination and declination of the local earths magnetic eld, and the orientation of the body with respect to magnetic north. To simplify anomaly shape, Baranov (1957) and Baranov and Naudy (1964) proposed a mathematical approach known as reduction to the pole. This method transforms the observed magnetic anomaly into the anomaly that would have been measured if the magnetization and ambient eld were both vertical as if the measurements were made at the magnetic pole. This method requires knowledge of the direction of magnetization, often assumed to be parallel to the ambient eld, as would be the case if remanent magnetization is either negligible or aligned parallel to the ambient eld. If such is not the case, the reduced-to-the-pole operation will yield unsatisfactory results. Reduction to the pole is now routinely applied to all data except for data collected at high magnetic latitudes. The RTP operator becomes unstable at lower magnetic latitudes because of a singularity that appears when the azimuth of the body and the magnetic inclination both approach zero. Numerous approaches have been proposed to overcome this problem. Leu (1982) suggested reducing anomalies measured at low magnetic latitudes to the equator rather than the pole; this approach overcomes the instability, but anomaly shapes are difcult to interpret. Pearson and Skinner (1982) proposed a whitening approach that strongly reduced the peak amplitude of the RTP lter, thus reducing noise. Silva (1986) used equivalent sources, which gave good results but could become unwieldy for large-scale problems. Hansen and Pawlowski (1989) designed an approximately regulated lter using Wiener techniques that accounted well for noise. Mendonc a and Silva (1993) used a truncated series approximation of the RTP operator. Gunn (1972, 1995) designed Wiener lters in the space domain by determining lter coefcients that transform a known input model at the survey location to a desired output at the pole. Keating and Zerbo (1996) also used Wiener ltering by introducing a deterministic noise model, allowing the method to be fully automated. Li and Oldenburg (1998b, 2000a) proposed a technique that attempts to nd the RTP eld under the general framework of an inverse formulation, with the RTP eld constructed by

Upward-downward continuation
Magnetic data measured on a given plane can be transformed to data measured at a higher or lower elevation, thus either attenuating or emphasizing shorter wavelength anomalies (Kellogg, 1953). These analytic continuations lead to convolution integrals which can be solved either in the space or frequency domain. The earliest attempts were done in the space domain by deriving a set of weights which, when convolved with eld data, yielded approximately the desired transform (Peters, 1949; Henderson, 1960; Byerly, 1965). Fuller (1967) developed a rigorous approach to determining the required weights and analyzing their performance. The space-domain operators were soon replaced by frequency-domain operators. Dean (1958) was the rst to recognize the utility of using Fourier transform techniques in performing analytic continuations. Bhattacharyya (1965), Byerly (1965), Mesko (1965), and Clarke (1969) contributed to the understanding of such transforms, which now are carried out on a routine basis. It is worth mentioning that while upward continuation is a very stable process, the opposite is true for downward continuation where special techniques, including lter response tapering and regularization, have to be applied in order to control noise. Analytic continuations are usually performed from one level surface to another. To overcome this limitation, Syberg (1972b) and Hansen and Miyazaki (1984) extended the

Historical Development of Magnetic Method

43ND

potential-eld theory to continuation between arbitrary surfaces, and Parker and Klitgord (1972) used a SchwarzChristoffel transformation to upward continue uneven prole data. Methods using equivalent sources were proposed by Bhattacharyya and Chan (1977a) and Li and Oldenburg (1999). Techniques designed to approximate the continuation between arbitrary surfaces include the popular chessboard technique (Cordell, 1985a), which calculates the eld at successively higher elevations, followed by a vertical interpolation between various strata and a Taylor series expansion (Cordell and Grauch, 1985).

Derivative-based lters
First and second vertical derivatives emphasize shallower anomalies and can be calculated either in the space or frequency domains. These operators also amplify high-frequency noise, and special tapering of the frequency response is usually applied to control this problem. A stable calculation of the rst vertical derivative was proposed by Nabighian (1984) using 3D Hilbert transforms in the X and Y directions. Before the digital age, use of the second vertical derivative for delineating and estimating depths to the basement formed the basis of aeromagnetic interpretation (Vacquier et al., 1951; Andreasen and Zietz, 1969). Many modern methods for edge detection and depth-tosource estimation rely on horizontal and vertical derivatives. Gunn et al. (1996) proposed using vertical gradients of order 1.5 and also showed the rst use of complex analytic signal attributes in interpretation. Use of the horizontal gradient for locating the edges of magnetic sources developed as an extension of Cordells (1979) technique to locate edges of tabular bodies from the steepest gradients of gravity data. Like gravity anomalies, a pseudogravity anomaly has its steepest gradients located approximately over the edges of a tabular body. Thus, Cordell and Grauch (1982, 1985) used crests of the magnitude of the horizontal gradient of the pseudogravity eld as an approximate tool for locating the edges of magnetic bodies. In practice, this approach can also be applied to the reduced-to-the-pole magnetic eld. This results in improved edge resolution, but some caution is required to avoid misinterpreting low-amplitude gradients attributable to side lobes (Phillips, 2000; Grauch et al., 2001). Blakely and Simpson (1986) presented a useful method for automatically locating and characterizing the crests of the horizontal gradient magnitude. A method by Pearson (2001) nds breaks in the direction of the horizontal gradient by application of a moving-window articial-intelligence operator. Another, similar technique is skeletonization (Eaton and Vasudevan, 2004), which produces not only an image but also a database of each lineament element, which can be sorted and decimated by length or azimuth criteria. Thurston and Brown (1994) developed convolution operators for controlling the frequency content of the horizontal derivatives and, thus, of the resulting edges. Cooper and Cowan (2003) introduced the combination of visualization techniques and fractional horizontal gradients to more precisely highlight subtle features of interest. The main advantages of the horizontal gradient method are its ease of use and stability in the presence of noise (Phillips, 2000; Pilkington and Keating, 2004). Its disadvan-

tages arise when edges are dipping or close together (Grauch and Cordell, 1987; Phillips, 2000) or when assumptions regarding magnetization direction are incorrect during the initial RTP or pseudogravity transformation. The method can also give misleading results when gradients from short-wavelength anomalies are superposed on those from long-wavelength anomalies. To address this problem, Grauch and Johnston (2002) developed a windowed approach to help separate local from regional gradients. The total gradient (analytic signal) is another popular method for locating the edges of magnetic bodies. For magnetic prole data, the horizontal and vertical derivatives t naturally into the real and imaginary parts of a complex analytic signal (Nabighian, 1972, 1974, 1984; Craig, 1996). In 2D (Nabighian, 1972), the amplitude of the analytic signal is the same as the total gradient, is independent of the direction of magnetization, and represents the envelope of both the vertical and horizontal derivatives over all possible directions of the earths eld and source magnetization. In 3D, Roest et al. (1992) introduced the total gradient of magnetic data as an extension to the 2D case. Unlike the 2D case, the total gradient in 3D is not independent of the direction of magnetization (Haney et al., 2003), nor does it represent the envelope of both the vertical and horizontal derivatives over all possible directions of the earths eld and source magnetization. Thus, despite its popularity, the total gradient is not the correct amplitude of the analytic signal in 3D. It is worth noting that what is now commonly called analytic signal should correctly be called the total gradient. The main advantage of the total gradient over the maximum horizontal gradient is its lack of dependence on dip and magnetization direction, at least in 2D. The approaches used to locate magnetic edges using the crests of the horizontal gradient can also be applied to the crests of the total gradient. This difference in the two methods can be used to advantage differences in edge locations determined by the two techniques can be used to identify the dip direction of contacts (Phillips, 2000) or to identify remanent magnetization (Roest and Pilkington, 1993). If the total gradient of the magnetic eld is somewhat analogous to the instantaneous amplitude used in seismic data analysis, then the local phase, dened as the arctangent of the ratio of the vertical derivative of the magnetic eld to the horizontal derivative of the eld, is analogous to the instantaneous phase. The local wavenumber, analogous to the instantaneous frequency, is dened as the horizontal derivative in the direction of maximum curvature of the local phase. Thurston and Smith (1997) and Thurston et al. (1999, 2002) showed that the local wavenumber is another function that has maxima over the edges of magnetic sources. Like the total gradient, the local wavenumber places maxima over the edges of isolated sources, regardless of dip, geomagnetic latitude, magnetization direction, or source geometry (see Magnetic Inverse Modeling section, Source parameter imaging subsection, which follows). The full expression for calculating the 3D local wavenumber is complicated (Huang and Versnel, 2000) and tends to produce noisy results. A better result is achieved by considering only the effects of 2D sources (Phillips, 2000; Pilkington and Keating, 2004). An alternate function that is easy to compute and approximates the absolute value of the full 3D local wavenumber

44ND

Nabighian et al.

is the horizontal gradient magnitude of the tilt angle (Miller and Singh, 1994; Pilkington and Keating, 2004; Verduzco et al., 2004). The tilt angle, rst introduced by Miller and Singh (1994), is the ratio of the rst vertical derivative to the horizontal gradient and is designed to enhance subtle and prominent features evenly. Finally, a form of lter that can be used to highlight faults is the Goussev lter, which is the scalar difference between the total gradient and the horizontal gradient (Goussev et al., 2003). This lter, in combination with a depth separation lter (Jacobsen, 1987), provides a different perspective from other lters and helps discriminate between contacts and simple offset faults. Wrench faults show up particularly well as breaks in the linear patterns of a Goussev lter.

Matched ltering
Spector (1968) and Spector and Grant (1970) showed that logarithmic radial-power spectra of gridded magnetic data contain constant-slope segments that can be interpreted as arising from statistical ensembles of sources, or equivalent source layers, at different depths. Spector (1968) designed the rst Fourier and convolution lters designed to separate the magnetic anomalies produced at two different source depths. The convolution lter was published by Spector (1971), while the Fourier lter was published, in simplied form, by Spector and Parker (1979). Syberg (1972a) rst applied the term matched lter to this process of matching the lter parameters to the power spectrum and developed Fourier-domain lters for separating the magnetic eld of a thin, shallow layer with azimuthally dependent power from the magnetic eld of a deeper magnetic half-space having different azimuthally dependent power. A Fourier lter for extracting the anomaly of the deepest source ensemble, without requiring any estimated parameters for shallow sources, was presented by Cordell (1985b). Ridsdill-Smith (1998a, b) developed waveletbased matched lters, while Phillips (2001) generalized the Fourier approach of Syberg (1972a) to sources at more than two depths and explained how matched Wiener lters could be used as an alternative to the more common amplitude lters. An alternative to matched lters, based on differencing of upward continued elds, was developed by Jacobsen (1987). Cowan and Cowan (1993) reviewed separation ltering and compared results of Spectors matched lter, the Cordell lter, the Jacobsen lter, and a second vertical derivative lter on an aeromagnetic data set from Western Australia.

The use of wavelets has been approached in three principle ways. The rst approach uses continuous wavelet transforms based on physical wavelets, such as those developed by Moreau et al. (1997) and Hornby et al. (1999). The former analyzes potential-eld data using various wavelets derived from a solution of Poissons equation, while the latter group takes a more intuitive approach and recasts commonly used processing methods in potential elds in terms of continuous wavelet transforms. These wavelets are essentially second-order derivatives of the potential produced by a point monopole source taken in different directions. Methods based on the continuous wavelet transform identify locations and boundaries of causative bodies by tracking the extrema of the transforms. Sailhac et al. (2000) applied a continuous wavelet transform to aeromagnetic proles to identify source location and boundaries. Haney and Li (2002) developed a method for estimating dip and the magnetization direction of two-dimensional sources by examining the behavior of extrema of continuous wavelet transforms. A second class of wavelet methodologies utilizes discrete wavelet transforms based on compactly supported orthonormal wavelets. Chapin (1997) applied wavelet transforms to the interpretation of gravity and magnetic proles. Ridsdill-Smith and Dentith (1999) used wavelet transforms to enhance aeromagnetic data. LeBlanc and Morris (2001) applied discrete wavelet transforms to remove noise from aeromagnetic data. Finally, Vallee et al. (2004) used this method to perform depth estimation and identify source types. In a third approach, discrete wavelet transforms are used to improve the numerical efciency of inversion-based techniques. Li and Oldenburg (2003) used discrete wavelet transforms to compress the dense sensitivity matrix in 3D magnetic inversion and thus reduce both memory requirement and CPU time in large-scale 3D inverse problem. A similar approach is also applied to the problem of upward continuation from uneven surfaces (Li and Oldenburg, 1999) and reduction to the pole using equivalent sources (Li and Oldenburg, 2000a).

MAGNETIC FORWARD MODELING


Before the use of electronic computers, magnetic anomalies were interpreted using characteristic curves calculated from simple models (Nettleton, 1942) or by comparison with calculated anomalies over tabular bodies (Vacquier et al., 1951). In the 1960s, computer algorithms became available for calculating magnetic anomalies across two-dimensional bodies of polygonal cross sections (Talwani and Heirtzler, 1964) and over three-dimensional bodies represented by right rectangular prisms (Bott, 1963; Bhattacharyya, 1964), by polygonal faces (Bott, 1963), or by stacked, thin, horizontal sheets of polygonal shape (Talwani, 1965). The 2D magnetic forward-modeling algorithm of Talwani and Heirtzler (1964) was later modied to include bodies of nite strike length (Shuey and Pasquale, 1973; Rasmussen and Pedersen, 1979; Cady, 1980), referred to as 2 1/2D. Computer programs to calculate magnetic and gravity proles across these 2 1/2D bodies, and also perform inversions, began to appear in the 1980s (Saltus and Blakely, 1983, 1993; Webring, 1985). The 3D magnetic forward-modeling algorithm of Talwani (1965) was modied by Plouff (1975, 1976), who replaced the

Wavelet transform
The wavelet transform is emerging as an important processing technique in potential-eld methods and has contributed signicantly to the processing and inversion of both gravity and magnetic data. The concept of continuous wavelet transform was initially introduced in seismic data processing (Goupillaud et al., 1984), while a form of discrete wavelet transform has long been used in communication theory. These were unied through an explosion of theoretical developments in applied mathematics. Potential-eld analysis and magnetic methods in particular, have beneted greatly from these developments.

Historical Development of Magnetic Method

45ND

thin horizontal sheets with nite-thickness prisms. The approach of Bott (1963) to modeling 3D bodies using polygonal facets was also used by Barnett (1976), who used triangular facets, and by Okabe (1979). A subroutine based on Botts approach appears in Blakely (1995). A complete treatment of gravity and magnetic anomalies of polyhedral bodies can be found in Holstein (2002a, b). Much attention has been paid to expressions for the Fourier transforms of magnetic elds of simple sources, both as a means of forward modeling and as an aid to inversion (Bhattacharyya, 1966; Spector and Bhattacharyya, 1966; Pedersen, 1978; Blakely, 1995). Parker (1972) presented a practical Fourier method for modeling complex topography, in which the observations are on a at plane or other surface that is above all the sources. Blakely (1981) published a computer program based on Parkers method, and Blakely and Grauch (1983) used the method to investigate terrain effects in aeromagnetic data own on a barometric surface over the Cascade Mountains of Oregon. The venerable right-rectangular prism has remained popular for voxel-based magnetic forward modeling and inversion. Hjelt (1972) presented the equations for the magnetic eld of a dipping prism having two opposite vertical sides that are parallelograms. This particular form of the voxel is useful for modeling magnetic anomalies caused by layered strata distorted by geologic processes, such as faulting and folding (Jessell et al., 1993; Jessell and Valenta, 1996; Jessell, 2001; Jessell and Fractal Geophysics, 2002).

exploration. Smith (1959) gave various rules for estimating the maximum possible depth to various magnetic sources. Trialand-error methods were also developed (Talwani, 1965), in which magnetic anomalies were calculated iteratively until a good t with observed data was obtained. In the 1970s, automated depth analysis began to supplant the graphical and trial-and-error techniques. These new methods took advantage of the digital aeromagnetic data that began to appear at that time, and they typically generated large numbers of depth estimates along magnetic proles based on simple but geologically reasonable 2D models such as sheets, contacts, or polygonal corners. Because validity of the models could not be assumed, the depth estimates still needed to be tested for reasonableness by appropriate forward modeling. In the 1990s, 3D automated depth-estimation methods began to appear. These were largely extensions of 2D methods designed for application to gridded magnetic data. Most of the methods mentioned below still exist in commercial or publicdomain software. There is no best method, and it is wise to use a variety of methods to identify consistent results: forward modeling is still a good idea.

Werner deconvolution
Automated depth-determination techniques have been limited mostly to prole data by assuming that targets are twodimensional. Werner (1955) proposed a method for interpreting overlapping effects of nearby anomalies if they can be interpreted as attributable to thin sheets. Assuming the causative bodies are two-dimensional and have a polygonal cross section, this can be achieved by taking the horizontal derivative of the observed prole. The method, now known as Werner deconvolution, was rst implemented by Hartman et al. (1971) and further rened by Jain (1976), Kilty (1983), Ku and Sharp (1983) and Tsokas and Hansen (1996). The method was rst extended to multiple 2D sources by Hansen and Simmonds (1993) and later extended to 3D multiple sources by Hansen (2002). The extension to multiple sources was achieved using deconvolution on the complex form of the analytic signal.

MAGNETIC INVERSE MODELING


From a purely mathematical point of view, there is always more than one model that will reproduce the observed data to the same degree of accuracy (the so-called nonuniqueness problem). However, geologic units producing the magnetic data that we acquire in real-world problems do not have an arbitrary variability. Imposing simple restrictions on admissible solutions based on geologic knowledge and integration with other independent data sets and constraints leads usually to distinct and robust results.

Depth-to-source estimation techniques


With the rst aeromagnetic surveys came the recognition that the largest magnetic anomalies were produced by sources near the top of the crystalline basement, and that the wavelengths of these anomalies increased as the basement rocks became deeper. Techniques were devised to estimate the depths to the magnetic sources and, thus, the thickness of the overlying sedimentary basins. Mapping basement structure became an important application of the new aeromagnetic method. Early depth-to-source techniques were mostly of graphical nature and applicable only to single-source anomalies (Henderson and Zietz, 1948; Peters, 1949; Vacquier et al., 1951; Smellie, 1956; Hutchison, 1958; Grant and Martin, 1966; Koulomzine et al., 1970; Barongo, 1985). These techniques estimated target parameters by looking at various attributes of an anomaly (curve matching, straight-slope, half-width, amplitude, horizontal extent between various characteristic points, etc). The straight-slope method in particular enjoyed immense popularity with interpreters working in petroleum

CompuDepth
OBrien (1972) introduced CompuDepth, a frequencydomain technique that determines location and depth to 2D magnetic sources based on successive frequency shifting of the Fourier spectrum, linear phase ltering, and solving a system of equations for the various target parameters. Wang and Hansen (1990) extended the method of OBrien to invert for the corners of 3D homogeneous polyhedral bodies.

Naudy method
Naudy (1971) proposed a method that uses a matched lter based on the calculated prole over a vertical dike or thin plate. The lter is applied twice, rst to the symmetrical component of the aeromagnetic prole and then to the symmetrical component of the same prole reduced to the pole. Shi (1991) improved convergence of the Naudy method by using horizontal and vertical components of the magnetic prole instead of observed and reduced-to-the-pole components

46ND

Nabighian et al.

and also extended the analysis to include dip estimates for the dikes.

Analytic signal
Nabighian (1972, 1974) introduced the concept of the analytic signal for magnetic interpretation and showed that its amplitude yields a bell-shaped function over each corner of a 2D body with polygonal cross section. For an isolated corner, the maximum of the bell-shaped curve is located exactly over the corner, and the width of the curve at half its maximum amplitude equals twice the depth to the corner. The determination of these parameters is not affected by the presence of remanent magnetization. Horizontal locations are usually well determined by this method, but depth determinations are only reliable for polyhedral bodies. Roest et al. (1992) used the total magnetic gradient, which they called the 3D analytic signal to approximately estimate positions of magnetic contacts and obtain some depth estimates from gridded data. Their results, however, are strongly dependent on the direction of total magnetization, in sharp contrast with the 2D case.

methodology for 3D magnetic source locations and structural indices using extended Euler or analytic signal methods. Finally, Mushayandebvu et al. (2004) showed that eigenvalues generated in the grid Euler solution could be exploited to decide automatically whether an individual anomaly was 2D or 3D and, in the former case, could be exploited to deduce strike and dip.

Source parameter imaging (SPITM )


Thurston and Smith (1997) and Thurston et al. (1999, 2002) developed the source parameter imaging (SPI) technique, based on the complex analytic signal, which computes source parameters from gridded magnetic data. The technique is sometimes referred to as the local wavenumber method. The local wavenumber has maxima located over isolated contacts, and depths can be estimated without assumptions about the thickness of the source bodies (Smith et al., 1998). Solution grids using the SPI technique show the edge locations, depths, dips and susceptibility contrasts. The local wavenumber map more closely resembles geology than either the magnetic map or its derivatives. The technique works best for isolated 2D sources such as contacts, thin sheet edges, or horizontal cylinders. The SPI method requires rst- and second-order derivatives and is thus susceptible to both noise in the data and to interference effects. Phillips (2000) compared the SPI method with the total-horizontal-gradient and analytic-signal methods and showed how the methods differ in their assumptions, accuracy, and sensitivity to noise and anomaly interference.

Euler deconvolution
Thompson (1982) proposed a technique for analyzing magnetic proles based on Eulers relation for homogeneous functions. The Euler deconvolution technique uses rst-order x, yand z derivatives to determine location and depth for various idealized targets (sphere, cylinder, thin dike, contact), each characterized by a specic structural index. Although theoretically the technique is applicable only to a few body types which have a known constant structural index, the method is applicable in principle to all body types. Reid et al. (1990) extended the technique to 3D data by applying the Euler operator to windows of gridded data sets. Mushayandebvu et al. (2000) and Silva and Barbosa (2003), among others, helped in understanding the applicability of the technique. Mushayandebvu et al. (2001) introduced a second equation derived from Eulers homogeneity equation which, when used in conjunction with the standard Euler equation, led to more stable solutions. This technique is now known as extended Euler deconvolution. The extended Euler deconvolution technique was generalized to 3D by Nabighian and Hansen (2001) using generalized Hilbert transforms (Nabighian, 1984). In the same paper, the authors showed that their proposed technique is also a 3D generalization of the Werner deconvolution technique, and thus both can be presented under a single unied theory. Although Barbosa et al. (1999) showed that attempting joint estimation of depth and structural index leads to unstable results using the traditional least-squares approach, others have claimed success by using alternative approaches such as differential similarity transformations (Stavrev, 1997; Gerovska and Arauzo-Bravo, 2003) and generalized Hilbert transform (Nabighian and Hansen, 2001; Fitzgerald et al., 2003). Hansen and Suciu (2002) extended the single-source Euler deconvolution technique to multiple sources to better account for the overlapping effects of nearby anomalies. Keating and Pilkington (2000) and Salem and Ravat (2003) proposed applying Euler deconvolution to the amplitude of the analytic signal, while Zhang et al. (2000) showed how the technique could be applied to tensor data. Phillips (2002) proposed a two-step

Statistical methods
All of the above techniques attempt to determine the location, shape, and depth of specic isolated targets. An entirely different approach considers the anomaly to be caused by an ensemble of magnetic sources in order to determine their average depth. The method was rst proposed by Spector and Grant (1970) and further rened by Treitel et al. (1971). Their method assumes that parameters of individual sources (length, width, depth, etc.) are governed by probabilities. Spector and Grant (1970) showed that the spectral properties of an ensemble of sources is equivalent to the spectral properties of an average member of the ensemble. For a single ensemble, the natural log of the radial power density spectrum as a function of wavenumber will have a linear slope approximately twice the maximum depth of the ensemble. For multiple ensembles, one obtains linear slopes approximately twice the maximum depths to the various magnetic ensembles. More accurate depths can be estimated by progressively stripping off effects of the shallowest ensembles (or equivalent layers), and by correcting the power spectrum for source body width (Spector and Grant, 1970, 1974) or for fractal magnetization models (Pilkington et al., 1994). The last approach was further expanded by Maus (1999) and Maus et al. (1999) as a robust method for depth-to-basement calculations. Computer programs for source depth estimation from magnetic proles using windowed statistical approaches were published by Phillips (1979) and Blakely and Hassanzadeh (1981).

Historical Development of Magnetic Method

47ND

Physical Property Mapping

Terracing
Terracing (Cordell and McCafferty, 1989) is an iterative ltering method applied to gridded magnetic (or gravity) data that gradually increases the slopes of the rst horizontal derivatives while simultaneously attening the eld between gradients. The resulting map is similar to a terraced landscape, hence the name applied to this technique. When imaged as a color map and illuminated from above, a terraced map resembles a geologic map in which the color scheme approximates the relative magnetizations of the geologic units. The method can be further rened by assigning susceptibility values to each unit by least-squares approximations until the calculated eld agrees reasonably well with the measured data.

Susceptibility mapping
Grant (1973) introduced a special form of inversion in which gridded magnetic data are inverted in the frequency domain to provide the apparent magnetic susceptibility of a basement represented by a large number of innite vertical prisms. The maps thus obtained reect the geology of the area, insofar as susceptibility is related to rock type. A similar approach was applied to the inversion of marine magnetic anomalies in 2D (Parker and Huestis, 1974) and 3D (Macdonald et al., 1980). Analogous space-domain methods were developed by Bhattacharyya and Chan (1977b), Silva and Hohmann (1984), and Misener et al. (1984) by reducing the problem to solving a large system of equations relating the observed data to magnetic sources in the ground.

Inversion
Inversion refers to an automated numerical procedure that constructs a model of subsurface geology from measured magnetic data and any prior information, with the additional condition that the input data are reproduced within a given error tolerance. Quantitative interpretation is then carried out by drawing geologic conclusions from the inverted models. As is typical for geophysical inverse problems, a purely mathematical solution of magnetic inversion is nonunique. The nonuniqueness arises mainly for two reasons. First, there are only a nite number of inaccurate measurements. Consequently, there is always more than one model that will reproduce the observed data to the same degree of accuracy. Second, Greens theorems dictate that many subsurface distributions of magnetization can produce exactly the same surface response. It is therefore important to recognize that even though magnetic inversion is nonunique from a purely mathematical point of view, it is equally important to understand that the often overemphasized nonuniqueness stems mainly from the mathematical properties of potential elds and has little to do with realistic geologic scenarios. In reality, geologic units producing the magnetic data that we acquire in real-world problems do not have an arbitrary variability. Imposing simple restrictions on admissible solutions based on geologic knowledge and integration with other independent data sets and constraints usually leads to distinct and robust results. In inversion methodology, a model is parameterized to describe either source geometry or the distribution of a physi-

cal property such as magnetic susceptibility. These lead to two major approaches to magnetic inversion. The rst approach inverts for the geometry of the source distribution. For example, following Botts (1960) work on inverting gravity data for basin depth by iteratively adjusting the depth of vertical prisms, several authors have formulated depth-to-basement inversion in a similar manner (e.g., Pedersen, 1977). Pilkington and Crossley (1986) inverted magnetic anomalies to estimate basement relief by applying linear inverse theory and Parkers (1972) forward-calculation technique. Pustisek (1990) developed a noniterative procedure to invert for magnetic basement. For isolated anomalies, the rst attempts parameterized the causative body with a single dike in 2D and rectangular prism in 3D. A parametric inversion was then carried out to recover the target parameters through the use of nonlinear least squares (e.g., Whitehill, 1973; Ballantyne, 1980; Bhattacharyya, 1980; and Silva and Homann, 1983). Alternatively, causative bodies are represented as polygonal bodies in 2D or polyhydronal bodies in 3D (Pedersen, 1979; Wang and Hansen, 1990), and the vertices of the objects are recovered as the unknowns. The second approach inverts for either magnetic susceptibility or magnetization. Parker and Huestis (1974) in their crustal studies inverted for the distribution of magnetization in a layer. Cribb (1976) represented the magnetic source by a set of dipoles and attempted to recover the dipole strengths by applying linear inverse theory. Guillen and Menichetti (1984) used a prismatic representation and performed a regularized inversion by minimizing the moment of inertia of the causative body. Li and Oldenburg (1996) formulated a generalized 3D inversion of magnetic data by using the Tikhonov regularization and a model objective function that measures the structural complexity of the model and incorporates a depthweighting function. A positivity constraint was also imposed on the recovered density contrast to further stabilize the solution. Pilkington (1997) introduced acceleration to this method by using the conjugate gradient. Li and Oldenburg (2000b, 2003) extended the method to include borehole data and applied wavelet-compression-based acceleration and a logarithmic method for imposing positivity. Most of these methods assume that the magnetization direction is known. As a result, their application is limited when strong remanent magnetization alters the total magnetization direction. To overcome this difculty, Shearer and Li (2004) developed a 3D nonlinear inversion to recover the magnetization magnitude by inverting the amplitude of the anomalous magnetic vector or the total gradient of the magnetic anomaly. These two quantities exhibit weak dependence on the direction of magnetization; therefore, precise knowledge of the latter is not required.

GEOLOGIC INTERPRETATION OF MAGNETIC DATA


Magnetic data, processing, and analysis give information and constraints about the distribution of magnetic materials at the surface and below. The goal of geologic interpretation is to render this information into a model of salient geologic features where they are not exposed. Modern geologic interpretation involves a complex synthesis of multiple aspects: the results of magnetic analysis, geologic knowledge of the study

48ND

Nabighian et al.

area, an understanding of rock-magnetic properties, integration with other independent data sets and constraints, geologic characterization of anomaly shapes and patterns, and identication of the contributions of topography and cultural sources. Over the past 75 years, the most signicant improvements in the quality and reliability of geologic interpretation have followed from increased data resolution, concurrent advancements in our understanding of all the aspects of the synthesis, and improvements to the processes used to synthesize them (e.g., modeling/inversion, data presentation, and data integration). These advances have continually improved our ability to constrain nonunique results predicted by theory, allowing for solutions that are distinct and robust in practice (see discussions in Gibson and Millegan, 1998, p. 68).

had submitted interpretations of an aeromagnetic survey in Venezuela: Local basement relief shown by the two magnetic interpretations is in poor agreement with basement depth information from seismograph and well data. Moreover, the two magnetic pictures bear little resemblance one to the other. However, in a discussion of this paper [R. J. Bean of Shell (Bean et al., 1961, p. 317)] commented that One of the contour maps (Interpretation A) is excellent and that any magnetic interpreter would be extremely gratied if all his basement contour maps checked as well as this one with data obtained subsequently by seismic surveys or by drilling. Interestingly, Steenland (who did not comment at the time because he was the Editor of (G EOPHYSICS) later (1963b) revealed that he was the author of Interpretation A.

Magnetic basement mapping


Shortly after World War II, when the antisubmarine warfare magnetometer was converted to use in geophysical airborne surveys, a new era in oil and gas exploration was born [along with a number of new aeromagnetic erale companies: AeroService, Fairchild, Compagnie Gen de Geophysique (CGG), Airmag, and Hunting]. The principal product of this ourishing activity was the magnetic basement map interpretation, as described by Steenland (1998). Steenlands interpretations, when coupled with an understanding of the local and regional geology, were intended as a basis for determining economic basement and (where it was deemed warranted) sedimentary structures controlled by local basement structures. The rst step in generating magnetic basement maps was to determine the depths to magnetic sources. He and others favored the straight-slope method, which grew out of the model anomalies comprehensively produced by Vacquier et al. (1951). Because they relied on human computers and were required to interpret very large volumes of aeromagnetic data, many interpreters of that era used this method to such an extent that it dominated the magnetic basement mapping industry. Based on simplicity and speed of calculation, the horizontal distance between points of departure of the magnetic trace from a coincident straight edge (corrected for azimuth) formed the basis for determining depth to basement. Steenland (1963a) gave a description of this technique and a comparison between the magnetic depth estimates and drilling results in the Peace River Arch in western Canada, and his statistical estimates of depth accuracy are still valid. At that time, the only model used was the so-called intrabasement model, believed to be zones of higher magnetic susceptibilities occurring within the basement rocks having great depth extent. At about the same time, but too late to incorporate in the Peace River publication, the suprabasement (or thin-plate) model was introduced, together with new indices for adjusting the derived depths. This model is included in a study of the Paradox Basin, Utah, together with a discussion of several aeromagnetic products that were generated in the 1960s and later (Steenland, 1962). Steenland (1963a) did not argue that the straight-slope method is the most accurate method; he and others favored it for economic reasons. But they did insist (Nettleton, 1971, p. 98) that this method particularly, and the aeromagnetic method generally, produced consistently accurate depth-tobasement maps. Not all agreed. Jacobsen (1961, p. 316) published an interesting blind test, in which two contractors

Geologic characterization of magnetic anomalies


Recognition of characteristic patterns and shapes of anomalies in relation to particular rock units or geologic structures is one of the rst steps in qualitative interpretation of a magnetic map. Correlation of magnetic maps to exposed geologic units was well established over 75 years ago (Stearn, 1929a). The correlation involves recognition of anomaly patterns typical of certain rock types or units particular to a study area, identication of breaks in anomaly patterns that may indicate structures, and delineating linear gradients. Seeing changes in anomaly patterns can be subjective and interpreted as differences in terranes, lithologies, or alteration. With the advent of HRAM surveys, many near-surface geologic features are so clearly expressed that their geologic origin is obvious in color shaded-relief images. For example, dendritic patterns of modern channels and paleochannels or glacial till are mimicked in the aeromagnetic data (Figure 1; Gunn, 1997; Davies et al., 2004; Gay, 2004). Folds look like folds (Figure 1); fault expressions can exhibit en echelon and anastomosing behavior (Figure 2; Grauch et al., 2001; Langenheim et al., 2004); and individual dikes within swarms are clearly resolved (Hildenbrand and Raines, 1990; Modisi et al., 2000). Within sedimentary basins, HRAM surveys allow a clear distinction between basement anomalies and near-surface volcanic rocks and between basement faults and near-surface faults (southwest corner of Figure 2). Volcanic rocks are typically associated with characteristic high-frequency patterns, which can be differentiated from circular and sometimes broader anomalies associated with intrusions (Figure 2). Magnetic anomalies produced by rocks with strong, reverse-polarity remanence display characteristic, high-amplitude negative anomalies (southcentral border of Figure 2; Books, 1962; Grauch et al., 1999) that, without high-resolution data, might be confused with magnetic lows caused by a lack of magnetization, which are also negative but generally featureless (Airo, 2002). The featureless character is well demonstrated in the HRAM image from the Murray Basin (Figure 1), where the underlying interpreted granite was conrmed by drilling (Bush et al., 1995).

DATA INTEGRATION / PRESENTATION


Some of the most profound improvements in geologic interpretation over the past 75 years have occurred in the realm of data presentation and integration with other data sets.

Historical Development of Magnetic Method

49ND

Visualization is key to understanding the patterns in data and how they interact with independent data sets. Although magnetic data were acquired along grid-like traverses or parallel lines as early as the late 19th century, the data were usually displayed in prole form (Smock, 1876; Smyth, 1896). By the 1920s, contour maps were a common way to display magnetic measurements (Stearn, 1929b; Heiland and Courtier, 1929). Contour maps of magnetic eld intensity became the primary display for magnetic maps for decades afterward, surviving the analog to digital transition of the 1970s through the development of automated contouring programs. Automated color-lled contour maps made their debut in the early 1980s, which facilitated the assessment of regional trends and magnitude variations (Paterson and Reeves, 1985). The volume, The Utility of Regional Gravity and Magnetic Maps, edited by Hinze (1985b), was one of the rst SEG publications to rely extensively on color contour maps. By the late 1980s, magnetic interpreters were borrowing from remote-sensing imaging technologies in the form of gray and color gradational images and shaded-relief images (Cordell and Knepper, 1987). By the mid-1990s, the color shaded-relief display was in common use. Today, many of these algorithms allow real-time variation of sun angle and 3D perspective. The shaded-relief display highlights fault zones, dikes, and other semilinear features that are difcult to see in contour-type displays. Complex geology with overlapping anomalies arising from different depths can limit the effectiveness of automated methods, such as fault detection. Sometimes subtle contiguous faults show up better using shaded-relief imaging. However, because a given sun direction highlights features that strike perpendicular to it, it is important to generate enough images with varying sun angles to illuminate all azimuths of lineament/fault trends. Pearson (2001) developed a way to use color to display 24 different sun angles at once. The Geographic Information System (GIS) revolution in the last decade allows unprecedented digital blending of magnetic anomalies with independent vector and raster information, such as remote sensing images, digital elevation models, electromagnetic data, gamma-ray data, and gravity data. 3D displays can combine any imaginable type of data, including seismic sections, drillhole data, and interpretive results. For example, the magnetic fault identication cube (MaFIC, Rhodes and Peirce, 1999) allows magnetic depth solutions to be integrated with seismic, well, topographic, and ltered magnetic data on any seismic work station.

Figure 1. Aeromagnetic data from a portion of the heavymineral sand province in the Murray Basin, the region draining into the Murray River in southeastern Australia. The area is almost entirely covered by Quaternary-Tertiary alluvium. Many deposits of the mineral sands produce subtle magnetic anomalies (labeled as magnetic strand lines) but are difcult to see at this regional scale. More obvious are the expressions of a wide variety of other geologic features, as labeled (from interpretations by Bush et al., 1995 and Moore, 2005). The image is derived from data that are c State of Victoria, Australia, 1999.

CASE HISTORIES
Murray Basin, Australia
The Murray Basin, in southeastern Australia, has become a major exploration target for deposits of heavy mineral sands, in large part due to a program of the Victorian government that began to provide high-resolution airborne geophysical data Figure 2. Color shaded-relief image of HRAM data extracted from Sweeney et al. (2002) for a strip crossing the Albuquerque basin just south of the metropolitan area. Geologic contacts (white lines) outline bedrock areas. Intrabasin faults and buried volcanic rocks, which are important for understanding the hydrogeology, are clearly imaged in the HRAM data. The magnetic expressions of the faults commonly connect isolated exposures, which signicantly increase the knowledge of their linear extents, patterns, and density. Interstate 25 (brown) and the Rio Grande (dark blue) are labeled.

50ND

Nabighian et al.

to the exploration industry in 1994 (www.dpi.vic.gov.au). Heavy mineral sands, which provide titanium and other industrial minerals, are associated with strand lines of Pliocene beach deposits (Roy et al., 2000). They have a weak but discernable signature in airborne magnetic data (Bush et al., 1995). An example of the Victorian aeromagnetic data, which were acquired at 80 m above ground along lines spaced 250 m apart, is shown in Figure 1. The area is almost entirely devoid of bedrock exposures and the sands are commonly buried under several tens of meters of alluvium as well. Despite this extensive cover, the aeromagnetic image shows not only the subtle features related to heavy mineral sands, but also reveals an amazing variety of geologic features and rock types that reect the underlying pre-Tertiary rocks (Bush et al., 1995; Moore, 2005). The wide range of geologic sources include magnetic intrusive, volcanic, and sedimentary rocks; nonmagnetic granites surrounded by magnetic, contact metamorphic aureoles; heavy mineral sands in beach deposits; and magnetite and maghemite concentrated in paleochannels. Prominent fold patterns in the western half of the map are accentuated by interlayered magnetic and nonmagnetic granites. Although heavy mineral sands are the focus of exploration in this area, the incredible view into the subsurface that the survey provides demonstrates the utility of aeromagnetic methods for cost effectively mapping the regional geology under cover.

Albuquerque Basin, New Mexico


The Albuquerque basin, part of the Rio Grande rift in north-central New Mexico, not only is a target for oil and gas exploration (e.g., Johnson et al., 2001), but also hosts basin aquifers that are the primary source of water for nonagricultural uses (Bartolino and Cole, 2002). Driven primarily by the increased water demands of a burgeoning population, HRAM surveys (with line spacings and terrain clearances of

100150 m) were own over the area in the late 1990s to map buried hydrogeologic features (Grauch et al., 2001). The resulting maps showed intrabasin faults and buried igneous rocks in incredible detail. An example is shown in Figure 2. Both these geologic features can signicantly inuence groundwater ow paths, rates, and storage. Figure 2 is extracted from the HRAM data for a strip crossing the basin just south of the Albuquerque metropolitan area. Intrabasin faults appear as widespread, semilinear, shaded anomalies. Exposed basalt elds (labeled as Qb and Tb) correspond to characteristic high-frequency anomaly patterns. The anomaly patterns outside the exposed elds show where volcanic rocks are concealed. Volcanic centers in the southern part of the area produce high-amplitude negative anomalies (area labeled strong R-polarity in Figure 2), a signature that is typical of volcanic rocks with strong, reversed-polarity remanent magnetization. The results from the Albuquerque HRAM survey have implications not only for groundwater exploration but for petroleum exploration as well. First, the magnetic anomalies at the faults can be entirely explained by the tectonic juxtaposition of sediments with differing magnetic properties, despite apparent magnetic lows over the fault zone (Grauch et al., 2001). This result revised a commonly held belief that all such anomalies were caused by alteration or mineralization along the fault plane related to the introduction of hydrocarbonrelated uids. Second, the magnetic images show that the density and linear extent of intrabasin faults are much greater than previously known (Figure 2). Thus, ideas need to be revised concerning the structural style and amount of extension for this and perhaps other basins. Finally, locations of shallow intrabasin faults can be compared to those of deeper basement faults (such as the intrabasin fault located next to the basement high in the southwest corner of Figure 2). The comparison can aid in correlating faults between seismic lines or in developing an understanding of how basement faults propagate to the surface.

Imperial Anticline, Northwest Territories, Canada


The Imperial Anticline is part of a regional thrust system in the northern part of the Canadian Cordillera, Northwest Territories. Although hydrocarbon exploration in the area has waned since the 1980s, the structural setting is similar to those of many other thrust structure plays in the western Cordillera. The regional thrust packages are composed of Phanerozoic and upper Proterozoic sedimentary rocks that were folded and thrust during Laramide time over a regionally thick (up to 14 km) sequence of Proterozoic sedimentary rocks (Cook and MacLean, 2004). Seismic data indicate that the Imperial Anticline formed above a structurally complex, beddingparallel thrust ramp (Cook and MacLean, 1999). Aeromagnetic data together with seismic reection data provide a comprehensive 3D view of the complex structure within the Imperial Anticline (Figure 3). Phanerozoic strata

Figure 3. Block diagram illustrating the relationship between thrust structures and residual magnetic anomalies over part of the Imperial Anticline, Northwestern Territories, Canada. The surface panel shows an image from aeromagnetic data collected by the Geological Survey of Canada (Geological Survey of Canada, 2005). Magnetic anomalies clearly correlate with steeply dipping stratigraphy in the hanging wall of the thrust fault, as identied by seismic-reection data (MacLean and Cook, 2002), shown in the cross-sectional view. Example compiled by Jim Davies, Image Interpretation Technologies, Inc.

Historical Development of Magnetic Method

51ND

involved in thrusting, shown in seismic reection data (MacLean and Cook, 2002), can be correlated to subtle magnetic anomalies in an aeromagnetic survey own at 200-m mean terrain clearance at 800-m line spacing (Geological Survey of Canada, 2005). After removing a smoothed version of the gridded data from the observed grid, the residual magnetic anomalies (3 nT amplitude) correlate to two separate magnetic horizons within the Phanerozoic stratigraphy. The anomalies may be produced by magnetic contrasts between shale and carbonate lithologies juxtaposed on one another through deposition. The seismic data show the structural complexities in cross section, whereas the aeromagnetic map displays the lateral extent and orientation of the strata that compose the larger structure.

Raglan deposit, northern Quebec, Canada


The Raglan deposit is located in northern Quebec, Canada, and its nickel mineralization is hosted in ultramac ow units. Little surface geologic expression is available, and exploration has relied primarily on geophysical studies, in particular magnetic surveys. Total-eld magnetic data in the area is typi-

ed by seemingly isolated magnetic highs interconnected by arc-like low intensity anomalies. Total-eld-intensity magnetic data covering an area of 4 km by 4 km is shown in Figure 4a. Two regions of high magnetic-eld intensity are observed, and they correspond to highly magnetic ultramac outcrops, which contain economic-grade ores. The geologic question was whether the outcrops were associated with a single ow unit or whether they were isolated bodies. The answer had important implications; in the former scenario, it meant that there is great potential for extending the ore reserves beyond that known from the shallower, isolated deposits. To answer this question the data was interpreted using generalized 3D inversion (Watts, 1997; Oldenburg et al., 1998). For the 3D inversion, the earth below the survey area was represented with 16 000 cubic cells (40 by 40 by 10 cells), each cell having dimensions of 100 m. The inversion was formulated to construct a susceptibility distribution that is smooth in all three spatial directions and close to a zero background,

a)

b)
MAG3D volume = 4 km x 4 km x 1 km Mag3D representation of UM flow Intersected 350 m of ultramafics, incl. 5 m of 1% Ni

Drillhole 781381

Location of 58 Orebody at surface

Manitobs
650 m 1050 m

Montans
MAG3D ISOMETRIC IMAGE (Looking NE) Susceptiblity range 0.02 to >0.08 SI

North Dakota

Figure 4. (a) Total eld aeromagnetic data. The parameters of the inducing eld are I = 83 and D = 32 . Data are contoured in nT. (b) The 3D susceptibility model recovered from magnetic inversion at Raglan deposit. This is a volumerendered image of the inverted susceptibility model, and the displayed surface provides a representation of the ultramac ow. Indicated in the volume-rendered representation is the intersecting drill hole that was spotted based upon the inversion results. Darker red colors indicate higher susceptibilities.

Figure 5. Index map for project showing the use of enhanced HRAM anomalies to correlate faults on 2D seismic data. This example was completed as part of the IEA Weyburn CO2 Sequestration Project. The study area is shown in red in the inset. The inner inset shows the distribution of wells in the area, including the Weyburn Field. Red lines show the HRAM data (500 1500-m line spacing) and the light green lines show the 2D seismic data being correlated. The drainage is shown in blue and the interpreted Souris River Fault is shown as a dark green dashed line.

52ND

Nabighian et al.

since the host rocks are, in general, nonmagnetic. The inversion indicated the presence of a continuous zone of highly magnetic material that extended between the two outcrops. The zone is shown as a volume-rendered image (Figure 4b) generated by displaying only susceptibility values greater than 0.04 SI. Ultramac ows are the only magnetic rocks in the area, so it seemed likely that the highly magnetic region found at depth has the same lithology as that of known deposits. A deep 1100-m hole, sited on the basis of this image, intersected magnetic rocks at a depth of 650 m. Moreover, a 10-m mineralized section (sub-ore grade, approximately 1% nickel) was intersected within the 350-m-thick intersection of magnetic ultramac rocks. Subsequently, 3D inversion has been used extensively in this region, and new geologic horizons and economic reserves have been found as a direct consequence (A. Watts, personal communication, 1966).

IEA Weyburn CO2 Sequestration Project, Saskatchewan, Canada

In the IEA Weyburn CO2 Sequestration Project (Figure 5; Goussev et al., 2004; Wilson and Monea, 2004) a large amount of 2D seismic data was made available to the project for mapping regional-scale faults in the area. The purpose of the mapping was to assess the security of the earth as a container for injected CO2 gas. The concern about leakage relates to some impurities in the injected gas that would be detrimental to the environment if they leaked to the surface. Because the faulting patterns were somewhat complicated and the seismic data were relatively widely spaced (Figure 5), Goussev et al. (2004) used GEDCOs proprietary HRAM data as an additional constraint to resolve the spatial aliasing of the fault correlations. Figure 6 shows three seismic lines and one ltered version of the magnetic data. At least six faults are imaged on these three seismic lines, and there is no straightforward correlation of the faults between the lines. The situation is further complicated because the seismic expression of the faults varies from line to line. Using the HRAM data, as enhanced by the Goussev lter, the preferred correlation is shown on the righthand map of Figure 6, with Fault A being the same on all three seismic images and following the distinct magnetic signature of the fault. This previously unknown fault is now called the Souris River Fault because it offsets the ow of the Souris River from its southeasterly regional ow into a short southerly leg for about 10 km. The fault is clearly present at the basement level on depth-migrated seismic processing, and it penetrates through the entire section to the surface, as evidenced by the course of the Souris River. In addition to demonstrating the utility of using HRAM data to constrain ambiguous seismic interpretations, this project also demonstrates clearly that some basement faults penetrate throughout the section in southeastern Saskatchewan. This is an important nding for the IEA CO2 Sequestration Project. Although there is no evidence that this fault is a leakage path from the reservoir to the surface, the possibility of other basement-to-surface faults exists, and each must be tested for gas leakage to ensure the integrity of the reservoir Figure 6. The map on the right shows a Goussev lter of the HRAM data (reds are highs and magenta colors are lows) from the IEA Weyburn CO2 Sequestration as a long-term storage container. Project. The four seismic lines are shown in dark brown and the fault locations as picked on a seismic workstation are indicated. Panels of seismic lines SOU-1, 2, and LOOKING FORWARD 3 are shown on the left, with interpreted faults labeled by letters A-D and PFS (positive ower structure) and NFS (negative ower structure). The positions of those It seems likely that in the near term, we faults on the map are connected to the seismic images of the faults by the yellow arrows. Because the seismic character of the faults is so variable, it is unlikely that will see continuing improvements in optianyone would correlate fault A across all three lines with the HRAM data as an cally pumped magnetometers. Resolution additional constraint. Goussev et al. (2004) name this fault the Souris River Fault because it offsets the course of the Souris River into a north-south direction for in the picotesla range and sample rates of about 10 km. This fault offsets basement and penetrates to the surface, so it is an around 1 kHz both seem achievable, and important consideration in the IEA Weyburn CO2 Sequestration Project. various applications could benet from that

Historical Development of Magnetic Method

53ND

instrument performance. The longer term is as always more difcult to predict. The design of magnetometers has seemed to be a mature science for many decades now, yet there have been order of magnitude improvements in performance over the past 20 years. Perhaps the next 20 will be just as interesting. We can expect to see a big boom in boutique data acquisition systems, such as autonomous and tethered remotely operated vehicles in the deep sea, helicopters carrying multiple sensors, and unmanned drones. Continentwide aeromagnetic data collection could be done almost entirely with drones using a long-lasting power source and cruise missile technology, collecting data 24/7 at very tight line spacing and low terrain clearance over unpopulated areas. Improvements in GPS will allow exact x, y, and z knowledge of sensor location. Tie-lines will no longer be required because equivalent source technology (or similar methods) will be able to process data directly where collected. Magnetic gradiometer measurements will be used more commonly as we turn our exploration focus from hydrocarbons and minerals to groundwater. The use of low-ying drones and very accurate GPS will allow us to measure gradient signatures in the uppermost 1000 m of the sedimentary section very accurately. We will need to improve our understanding of what makes sedimentary rocks and unconsolidated sediments magnetic in order to make better interpretations for groundwater exploration and fracture identication. Aeromagnetic surveys with tighter line spacing will yield information about intrasedimentary lithology, mineralized faults, and geochemical alteration in reconnaissance areas. Aeromagnetic modeling of 3D volumes will proliferate into the hands of more explorationists thanks to competition between software companies. More prolic modeling will yield a better understanding of cultural noise, deep regional noise, and, therefore, exploration targets. Laboratory measurements of remanent magnetization will discover why 2D and 3D models typically underestimate the amplitude of the observed magnetic eld. Magnetic Curie isotherm and upper-crustal intrusion studies, bolstered by geothermal measurements, will point to prospective areas of basin-centered gas generation. Spectral decomposition ltering, possibly in the form of matched Wiener models, will assist the Werner and Euler depth techniques at mapping edges within the vertical geologic column. Filters will be developed that will improve our ability to discriminate sources at different depths. Loading all magnetic data and depth solutions onto a seismic workstation will become standard practice.

cal development of the magnetic method. In a paper of this magnitude it is inevitable that we might have missed some. This was not intentional and we apologize for any unintended omissions.

APPENDIX A TIMELINE FOR MAGNETIC METHODS OF EXPLORATION


Date 600 B . C . E . 1600 1640 1840 1850 18501890 1880 18901900 1896 19001930 1910 1915 1925-1930 1929 1931 1941 1942 1943 1944 19451955 1945 1946 1946 1947 1947 1948 1949 19501960 Event Thales Magnetic forces of lodestones Sir William Gilbert publishes De Magnete Iron ore prospecting begins in Sweden Carl Friedrich Gauss Spherical harmonic analysis of geomagnetic data The rst dip needle (Swedish Mining Compass) is developed Dip needles used primarily as divining rods to prospect for iron ore Development of earth inductors for magnetic eld component measurements Use of dip needle expands to map iron formations under cover H. L. Smyth develops magnetic calculations for tabular bodies Applications expand to include exploration for base metals, oil, and gold Edelmann First airborne measurement in balloon Adolph Schmidt develops vertical eld balance Rock magnetic properties are rst studied for interpreting magnetic surveys Matayama proposes that the earths magnetic eld has experienced reversals Development of uxgate magnetometer Vacquier et al. develop airborne detection of submarines during World War II Nettleton Gravity and magnetic calculations for simple bodies George P. Woollard First transcontinental magnetic prole of U. S. A. Balsley First aeromagnetic (biplane) survey, Boyertown, Pennsylvania Graphical depth-to-source techniques are developed Balsley First ore deposit discovered from aeromagnetic data, Adirondacks, New York Balsley First offshore aeromagnetic survey, coastal Gulf of Mexico First aeromagnetic maps published Canada begins systematic national aeromagnetic coverage First aeromagnetic survey in Australia First shipborne marine magnetic survey for study of the ocean oor Convolution methods are developed for derivatives and analytical continuation Joint aeromagnetic-EM surveys become common for base-metal exploration Magnetic basement mapping becomes popular in oil exploration Vacquier et al. Interpretation of aeromagnetic maps published by Geological Society of America Australia begins systematic national aeromagnetic coverage Finland begins systematic national aeromagnetic coverage Development of proton precession magnetometer

ACKNOWLEDGMENTS
The authors wish to sincerely thank Jim Davies for supplying images and information on the Imperial Valley Anticline case history and to David Moore for providing updates and advice on the Murray Basin case history. The manuscript beneted greatly from thoughtful reviews by Rick Blakely, Carol Finn, Peter Gunn, and Alan Reid. Valuable insight on data presentation/integration was provided by William Pearson. The authors made a serious effort to include in this paper all pertinent developments and references related to the histori-

1951 1951 1951 1955

54ND

Nabighian et al.

Date 1955 1957 1958 1960 1962 1963 1964 19651975 1965 1965 1965 1965 1967 1967 1968 1969 1969 1970 1971 1971 1972 1972 1972 1972 1972 1972 1972 1972 1973 1973 19741980 1974 1975 1976 1979 1979 1980 1980

Event Werner originates depth-estimation method now known as Werner deconvolution Baranov Reduction-to-the-pole and pseudogravity transformation Dean Use of Fourier methods for derivatives and analytical continuation Bott Iterative inversion of gravity and magnetic data Development of optically pumped magnetometers Vine-Matthews-Morley model for seaoor spreading Cosmos 49 scalar satellite magnetometer is launched Digital recording of aeromagnetic surveys becomes routine Talwani Digital computation of magnetic anomalies Cooley and Tukey Development of fast Fourier transform (FFT) Hood Gradient measurements for airborne surveying Hood introduces concept of Euler deconvolution Fuller Comprehensive analysis of space-domain lters First national aeromagnetic map of Canada published International Geomagnetic Reference Field (IGRF) established Australia begins collecting joint spectral gamma-ray and aeromagnetic data Dampney Equivalent source technique Spector and Grant Statistical methods for magnetic interpretation Naudy Automatic magnetic depth determination Hartmann et al. Implementation of Werner deconvolution OBrien CompuDepth Nabighian Analytic Signal Syberg Potential eld continuation and matched lters Wiggins Generalized linear inverse theory Parker Fourier modeling of complex topography Gunn Wiener lters for transformations of gravity and magnetic elds Finland completes national high-altitude (150-m) aeromagnetic coverage Finland begins national low-altitude (40-m) aeromagnetic coverage Grant Susceptibility mapping McGrath and Hood Multimodel least-squares interpretation Joint aeromagnetic and gamma-ray surveys of U. S. for uranium exploration Briggs Minimum curvature gridding Zimmerman and Campbell SQUID magnetometers First aeromagnetic anomaly map of Australia published Launch of MAGSAT, the rst vector magnetometer satellite mission First aeromagnetic anomaly map of Soviet Union published Reid Aeromagnetic survey design First aeromagnetic anomaly map of Finland published

Date 1982 1982 1984 1984 1984 1984 19801990 1985 1985 1985 1987 1989 1990 19901995 19902000 1992 1996 1997 1997 1998 19902000 2001 2001 2002 2002 2002 2004

Event Composite magnetic anomaly map of the conterminous U. S. published Thompson Euler deconvolution method for 2D depth estimation Lines and Treitel Least-squares inversion techniques Hardwick Compensation of aircraft magnetic eld Nabighian 3D Hilbert transforms Goupillaud et al. Wavelet transforms for geophysical applications HRAM surveys own to test direct detection of hydrocarbons Cordell and Grauch introduce horizontal gradient method for magnetic data Cordell introduces chessboard method for continuation to irregular surfaces Shaded-relief image display of aeromagnetic data becomes popular Grauch Magnetic terrain effects Cordell and McCafferty Terracing Reid et al. 3D Euler deconvolution on gridded data sets GPS navigation increases location accuracy for airborne surveys HRAM surveys become standard Roest et al. Total gradient Li and Oldenburg 3D inversion of magnetic data Thurston and Smith SPI (local wavenumber) technique Moreau et al. Wavelet analysis of potential elds Archibald et al. Multiscale edge analysis of potential eld data GIS and 3D visualization greatly improve magnetic interpretation Mushayandebvu et al. Extended Euler deconvolution Nabighian and Hansen Unication of Euler and Werner deconvolutions Hansen Multisource Werner deconvolution Hansen and Suciu Multisource Euler deconvolution Comprehensive model (CM) proposed as replacement for IGRF Mushayandebvu et al. Eigenvalue analysis for the 3D Euler equation

Note: In the space allotted, it is impossible to include all the important stages of development of the magnetic method, and omissions are inevitable.

REFERENCES
Abaco, C. I., and D. C. Lawton, 2003, Magnetic anomalies in the Alberta foothills, Canada: SEG Expanded Abstracts, 22, 612 615. Agarwal, R. G., and E. R. Kanasewich, 1971, Automatic trend analysis and interpretation of potential eld data: Geophysics, 36, 339348. Agocs, W. B., 1951, Least-squares residual anomaly determination: Geophysics, 16, 686696. Airo, M. -L., 2002, Aeromagnetic and aeroradiometric response to hydrothermal alteration: Surveys in Geophysics, 23, 273302. Allingham, J. W., 1964, Low-amplitude aeromagnetic anomalies in southeastern Missouri: Geophysics, 29, 537552. Andreasen, G. E., and I. Zietz, 1969, Magnetic elds for a 4 6 prismatic model: U. S. Geological Survey Professional Paper 666. Arkani-Hamed, J., 1988, Differential reduction-to-the-pole of regional magnetic anomalies: Geophysics, 53, 15921600. Ballantyne Jr., E. J., 1980, Magnetic curve t for a thin dike Calculator program (TI-59): Geophysics, 45, 447455.

Historical Development of Magnetic Method Balsley, J. R., 1952, Aeromagnetic surveying, in H. E. Landsberg, ed., Advances in Geophysics, 1: Academic Press, 313350. Baranov, V., 1957, A new method for interpretation of aeromagnetic maps pseudo-gravimetric anomalies: Geophysics, 22, 359383. Baranov, V., and H. Naudy, 1964, Numerical calculation of the formula of reduction to the magnetic pole: Geophysics, 29, 6779. Barbosa, V. C. F., J. B. C. Silva, and W. E. Medeiros, 1999, Stability analysis and improvement of structural index estimation in Euler deconvolution: Geophysics, 64, 4860. Barnett, C. T., 1976, Theoretical modeling of the magnetic and gravitational elds of an arbitrarily shaped three-dimensional body: Geophysics, 41, 13531364. Barongo, J. O., 1985, Method for depth estimation and aeromagnetic vertical gradient anomalies: Geophysics, 50, 963968. Bartolino, J. R., and J. C. Cole, 2002, Ground-water resources of the Middle Rio Grande Basin, New Mexico: U. S. Geological Survey Circular 1222. Barton, C. E., 1997, International Geomagnetic Reference Field: The seventh generation: Journal of Geomagnetism and Geoelectricity, 49, 123148. Batchelor, A., and J. Gutmanis, 2002, Hydrocarbon production from fractured basement reservoirs version 7: www.geoscience.co.uk/ downloads/fracturedbasementver7.pdf. Bath, G. D., 1968, Aeromagnetic anomalies related to remanent magnetism in volcanic rock, Nevada Test Site: Geological Society of America Memoir 110, 135146. Bath, G. D. and C. E. Jahren, 1984, Interpretation of magnetic anomalies at a potential repository site located in the Yucca Mountain area, Nevada Test Site: U. S. Geological Survey Open File Report 84120. Bean, R. J., W. F. Fillippone, N. R. Paterson, and I. Zietz, 1961, Discussion of An evaluation of basement depth determination from airborne magnetometer data, by Peter Jacobsen Jr.: Geophysics, 26, 317319. Bhattacharyya, B. K., 1964, Magnetic anomalies due to prism-shaped bodies with arbitrary polarization: Geophysics, 29, 517531. , 1965, Two-dimensional harmonic analysis as a tool for magnetic interpretation: Geophysics, 30, 829857. , 1966, Continuous spectrum of the total-magnetic-eld anomaly due to a rectangular prismatic body: Geophysics, 31, 97 121. , 1969, Bicubic spline interpolation as a method for treatment of potential eld data: Geophysics, 34, 402423. , 1980, A generalized multibody model for inversion of magnetic anomalies: Geophysics, 45, 255270. Bhattacharyya, B. K., and K. C. Chan, 1977a, Reduction of magnetic and gravity data on an arbitrary surface acquired in a region of high topographic relief: Geophysics, 42, 14111430. , 1977b, Computation of gravity and magnetic anomalies due to inhomogeneous distribution of magnetization and density in a localized region: Geophysics, 42, 602609. Blakely, R. J., 1981, A program for rapidly computing the magnetic anomaly over digital topography: U. S. Geological Survey Open File Report 81298. , 1995, Potential theory in gravity and magnetic applications: Cambridge University Press. Blakely, R. J., and V. J. S. Grauch, 1983, Magnetic models of crystalline terrane; accounting for the effect of topography: Geophysics, 48, 15511557. Blakely, R. J., and S. Hassanzadeh, 1981, Estimation of depth to magnetic source using maximum entropy power spectra, with application to the Peru-Chile Trench, Nazca Plate; Crustal formation and Andean convergence: Geological Society of America Memoir 154, 667682. Blakely, R. J., and R. W. Simpson, 1986, Approximating edges of source bodies from magnetic or gravity anomalies: Geophysics, 51, 14941498. Blakely, R. J., V. E. Langenheim, D. A. Ponce, and G. L. Dixon, 2000a, Aeromagnetic survey of the Amargosa Desert, Nevada and California; a tool for understanding near-surface geology and hydrology: U. S. Geological Survey Open File Report 00-0188, http://pubs.usgs.gov/open-le/of00-188/. Blakely, R. J., R. E. Wells, T. L. Tolan, M. H. Beeson, A. M. Trehu, and L. M. Liberty, 2000b, New aeromagnetic data reveal large strike-slip faults in the northern Willamette Valley, Oregon: Geological Society of America Bulletin, 112, 12251233. Boardman, J. W., 1985, Magnetic anomalies over oil elds: M.S. thesis, Colorado School of Mines. Books, K. G., 1962, Remanent magnetism as a contributor to some aeromagnetic anomalies: Geophysics, 27, 359375. Bott, M. H. P., 1960, The use of rapid digital computing methods for direct gravity interpretation of sedimentary basins:

55ND

Geophysical Journal of the Royal Astronomical Society, 3, 6367. , 1963, Two methods applicable to computers for evaluating magnetic anomalies due to nite three dimensional bodies: Geophysical Prospecting, 11, 292299. Breiner, S., 1981, Horizontal gradient methods for airborne and marine geophysical exploration: 51st Annual International Meeting, SEG, Expanded Abstracts, Geophysics, 441442. Briggs, I. C., 1974, Machine contouring using minimum curvature: Geophysics, 39, 3948. Broding, R. A., C. W. Zimmerman, E. V. Somers, E. S. Wilhelm, and A. A. Stripling, 1952, Magnetic well-logging: Geophysics, 17, 126. Bush, M. D., R. A. Cayley, S. Rooney, K. Slater, and M. L. Whitehead, 1995, The geology and prospectivity of the southern margin of the Murray Basin: Geological Survey of Victoria VIMP Report 4. Butler, D. K., 2001, Potential elds methods for location of unexploded ordnance: The Leading Edge, 20, 890895. Byerly, P. E., 1965, Convolution ltering of gravity and magnetic maps: Geophysics, 30, 281283. Cady, J. W., 1980, Calculation of gravity and magnetic anomalies of nite-length right polygonal prisms: Geophysics, 45, 15071512. Campbell, W. C., 1997, Introduction to geomagnetic elds: Cambridge University Press. Campos-Enriquez, J. O., R. Diaz-Navarro, J. M. Espindola, and M. Mena, 1996, Chicxulub Subsurface structure of impact crater inferred from gravity and magnetic data: The Leading Edge, 15, 357359. Chapin, D. A., 1997, Wavelet transforms: A new paradigm for interpreting gravity and magnetics data?: 67th Annual International Meeting, SEG, Expanded Abstracts, 486489. Chapin, D. A., S. V. Yalamanchili, and P. H. Daggett, 1998, The St. George Basin, Alaska, COST #1 well: An example of the need for integrated interpretation, in R. I. Gibson, and P. S. Millegan, eds., Geologic applications of gravity and magnetics: Case histories: SEG and AAPG. Clark, D. A., 1983, Comments on magnetic petrophysics: Bulletin of Australian Society of Exploration Geophysicists, 14, 4962. , 1997, Magnetic petrophysics and magnetic petrology; aids to geological interpretation of magnetic surveys: AGSO Journal of Australian Geology and Geophysics, 17, 83103. Clark, D. A., and D. W. Emerson, 1991, Notes on rock magnetization characteristics in applied geophysical studies: Exploration Geophysics, 22, 547555. Clarke, G. K. C., 1969, Optimum second derivative and downward continuation lters: Geophysics, 34, 424437. Committee for the Magnetic Anomaly Map of North America, 1987, Magnetic anomaly map of North America: Geological Society of America Continent Scale Map 003. Cook, D. G., and B. C. MacLean, 1999: The Imperial Anticline, a fault-bend fold above a bedding-parallel thrust ramp, Northwest Territories, Canada: Journal of Structural Geology, 21, 215228. , 2004: Subsurface Proterozoic stratigraphy and tectonics of the western plains of the Northwest Territories. Geological Survey of Canada, Bulletin 575. Cooper, G. R. J., and D. R. Cowan, 2003, Sunshading geophysical data using fractional order horizontal gradients: The Leading Edge, 22, 204. Cordell, L., 1979, Gravimetric expression of graben faulting in Santa Fe County and the Espanola Basin, New Mexico, in R. V. Ingersoll, ed., Guidebook to Santa Fe County: 30th Field Conference, New Mexico Geological Society, 5964. , 1985a, Applications and problems of analytical continuation of New Mexico aeromagnetic data between arbitrary surfaces of very high relief: Proceedings of International Meeting on Potential Fields in Rugged Topography, Institut de Geophysique de Univer de Lausanne, Bulletin 7, 96101 site , 1985b, A stripping lter for potential eld data: SEG, Expanded Abstracts, 55th Annual International Meeting, SEG, Expanded Abstracts, 217218. , 1992, A scattered equivalent-source method for the interpolation and gridding of potential-eld data in three dimensions: Geophysics, 57, 629636. Cordell, L., and V. J. S. Grauch, 1982, Mapping basement magnetization zones from aeromagnetic data in the San Juan Basin, New Mexico: 52nd Annual International Meeting, SEG, Expanded Abstracts, 246247. , 1985, Mapping basement magnetization zones from aeromagnetic data in the San Juan Basin New Mexico, in W. J. Hinze, ed., Utility of regional gravity and magnetic maps: SEG, 181197. Cordell, L., and D. H. Knepper, 1987, Aeromagnetic images: Fresh in-

56ND

Nabighian et al. tion methods in gravity and magnetics, 196471: Geophysics, 37, 647661. Grant, F. S., 1973, The magnetic susceptibility mapping method for interpreting aeromagnetic surveys: 43rd Annual International Meeting, SEG, Expanded Abstracts, 1201. , 1985, Aeromagnetics, geology and ore environments, in Magnetite in igneous, sedimentary and metamorphic rocks: An overview: Geoexploration, 23, 303333. Grant, F. S., and L. Martin, 1966, Interpretation of aeromagnetic anomalies by the use of characteristic curves: Geophysics, 31, 135148. Grauch, V. J. S., 1987, A new variable-magnetization terrain correction method for aeromagnetic data: Geophysics, 52, 94107. Grauch, V. J. S., and L. Cordell, 1987, Limitations on determining density or magnetic boundaries from the horizontal gradient of gravity or pseudogravity data: Geophysics, 52, 118121. Grauch, V. J. S., and C. S. Johnston, 2002, Gradient window method: A simple way to separate regional from local horizontal gradients in gridded potential-eld data: 72nd Annual International Meeting, SEG, Expanded Abstracts, 762765. Grauch, V. J. S., D. A. Sawyer, C. J. Fridrich, and M. R. Hudson, 1999, Geophysical framework of the southwestern Nevada volcanic eld and hydrogeologic implications: U. S. Geological Survey Professional Paper1608. Grauch, V. J. S., M. R. Hudson, and S. A. Minor, 2001, Aeromagnetic expression of faults that offset basin ll, Albuquerque basin, New Mexico: Geophysics, 66, 707720. Grifn, W. R., 1949, Residual gravity in theory and practice: Geophysics, 14, 3956. Grivet, P. A., and L. Malnar, 1967, Measurement of weak magnetic elds by magnetic resonance: Advances in Electronics and Electron Physics, 23, 39151. Guillen, A., and V. Menichetti, 1984, Gravity and magnetic inversion with minimization of a specic functional: Geophysics, 49, 1354 1360. Gunn, P. J., 1972, Application of Wiener lters to transformations of gravity and magnetic elds: Geophysical Prospecting, 20, 860871. , 1975, Linear transformations of gravity and magnetic elds: Geophysical Prospecting, 23, 300312. , 1995, An algorithm for reduction to the pole that works at all magnetic latitudes: Exploration Geophysics, 26, 247254. , 1997, Application of aeromagnetic surveys to sedimentary basin studies: AGSO Journal of Australian Geology and Geophysics, 17, 133144. , 1998, Aeromagnetics locates prospective areas and prospects: The Leading Edge, 17, 6769. Gunn, P. J., D. FitzGerald, and N. Yassi, 1996, Complex attributes: New tools for enhancing aeromagnetic data: Australian Geological Survey Organization (AGSO) Research Newsletter, no. 25, 1617. Guspi, F., and B. Introcaso, 2000, A sparse spectrum technique for gridding and separating potential eld anomalies: Geophysics, 65, 11541161. Haggerty, S. E., 1979, The aeromagnetic mineralogy of igneous rocks: Canadian Journal of Earth Sciences, 16, 12811293. Hall, S. H., 1962, The modulation of a proton magnetometer signal due to rotation: Geophysical Journal, 7, 131142. Hammer, S., 1963, Deep gravity interpretation by stripping: Geophysics, 28, 369378. Haney, M., and Y. Li, 2002, Total magnetization direction and dip from multiscale edge: 72nd Annual International Meeting, SEG, Expanded Abstracts, 735738. Haney, M., C. Johnston, Y. Li, and M. Nabighian, 2003, Envelopes of 2D and 3D magnetic data and their relationship to the analytic signal: Preliminary results: 73rd Annual International Meeting, SEG Expanded Abstracts, 596599. Hanna, W. F., 1990, Some historical notes on early magnetic surveying in the U. S. Geological Survey, in W. F. Hanna, ed., Geologic applications of modern aeromagnetic surveys: U. S. Geological Survey Bulletin 1924, 6373. Hansen, R. O., 1984, Two approaches to total eld reconstruction from gradiometer data: 54th Annual International Meeting, SEG Expanded Abstracts, 245. , 1993, Interpretative gridding by anisotropic kriging: Geophysics, 58, 14911497. , 2002, 3D multiple-source Werner deconvolution: 72nd Annual International Meeting, SEG, Expanded Abstracts, 802805. Hansen, R. O., and Y. Miyazaki, 1984, Continuation of potential elds between arbitrary surfaces: Geophysics, 49, 787795. Hansen, R. O., and R. S. Pawlowski, 1989, Reduction-to-thepole at low latitudes by Wiener ltering: Geophysics, 54, 1607 1613.

sight to the buried basement, Rolla quadrangle, southeast Missouri: Geophysics, 52, 218231. Cordell, L., and A. E. McCafferty, 1989, A terracing operator for physical property mapping with potential eld data: Geophysics, 54, 621634. Cowan, D. R., and S. Cowan, 1993, Separation ltering applied to aeromagnetic data: Exploration Geophysics, 24, 429436. Cox, A., 1973, Plate tectonics and geomagnetic reversals: W. H. Freeman and Company. Craig, M., 1996, Analytic signals for multivariate data: Mathematical Geology, 28, 315329. Cribb, J., 1976, Application of the generalized linear inverse to the inversion of static potential data: Geophysics, 41, 13651369. Davies, J., M. F. Mushayandebvu, and R. Smith, 2004, Magnetic detection and characterization of Tertiary and Quaternary buried channels: 74th Annual International Meeting, SEG, Expanded Abstracts, 734. Dean, W. C., 1958, Frequency analysis for gravity and magnetic interpretation: Geophysics, 23, 97127. Dickinson, M., 1986, A search for intra-sedimentary aeromagnetic anomalies over a known oil eld in the Denver Basin: M.S. thesis, Colorado School of Mines. Dietz, R. S., 1961, Continent and ocean basin evolution by spreading of the sea oor: Nature, 190, 854857. Eaton, D., and K. Vasudevan, 2004, Skeletonization of aeromagnetic data: Geophysics, 69, 478488. EG&G Geometrics (ca.1970s), A guide to passive magnetic compensation of aircrafts: EG&G Geometrics Technical Report Number 15. Fairhead, J. D., J. D. Misener, C. M. Green, G. Bainbridge, and S. W. Reford, 1997, Large scale compilations of magnetic, gravity, radiometric and electromagnetic data: The new exploration strategy for the 90s, in A. G. Gubins, ed., Proceedings of Exploration 97: Fourth Decennial International Conference on Mineral Exploration, 805816. Finn, C. A., ed., 2002, Examples of the utility of magnetic anomaly data for geologic mapping: U. S. Geological Survey Open-le Report 02-0400, available at http://pubs.usgs.gov/of/2002/ofr-02-0400/ Finn, C. A., T. W. Sisson, and M. Deszcz-Pan, 2001, Aerogeophysical measurements of collapse-prone hydrothermally altered zones at Mount Rainier Volcano: Nature, 409, 600603. Finn, C. A., and L. A. Morgan, 2002, High-resolution aeromagnetic mapping of volcanic terrain, Yellowstone National Park: Journal of Volcanology and Geothermal Research, 115, 207231. Fitzgerald, D., A. Reid, and P. McInerney, 2003, New discrimination techniques for Euler deconvolution: 8th South African Geophysical Association (SAGA) Biennial Technical Meeting and Exhibition. Frischknecht, F. C., L. Muth, R. Grette, T. Buckley, and B. Kornegay, 1983, Geophysical methods for locating abandoned wells: U. S. Geological Survey Open File Report 83702. Frowe, E., 1948, A total eld magnetometer for mobile operation: Geophysics, 13, 209214. Fuller, B. D., 1967, Two-dimensional frequency analysis and design of grid operators, in Mining Geophysics, II: Society of Exploration Geophysicists, 658708. Gay Jr., S. P., 1992, Epigenetic versus syngenetic magnetite as a cause of magnetic anomalies: Geophysics, 57, 6068. , 2004, Glacial till: A troublesome source of near-surface magnetic anomalies: The Leading Edge, 23, 542547. Gay, S. P., and B. W. Hawley, 1991, Syngenetic magnetic anomaly sources: Three examples: Geophysics, 56, 902913. Geological Survey of Canada, 2005, Canadian Aeromagnetic Data Base, Central Canada Division, Earth Sciences Sector, Natural Resources Canada. Gerovska, D., and M. J. Arauzo-Bravo, 2003, Automatic interpretation of magnetic data based on Euler deconvolution with unprescribed structural index: Computers & Geosciences, 29, 949960. Gibson, R. I., and P. S. Millegan, eds., 1998, Geologic applications of gravity and magnetics: Case histories: SEG and AAPG. Goldhaber, M. B., and R. L. Reynolds, 1991, Relations among hydrocarbon reservoirs, epigenetic suldization, and rock magnetization: Examples from the south Texas coastal plain: Geophysics, 56, 748 757. Goupillaud, P. L., A. Grossmann, and J. Morlet, 1984, Cycle-octave and related transforms in seismic signal analysis: Geoexploration, 23, 85102. Goussev, S. A., R. A. Charters, J. W. Peirce, and W. E. Glenn, 2003, Jackpine magnetic anomaly: Identication of a buried meteorite impact structure: The Leading Edge, 22, 740741. Goussev, S. A., L. Grifth, J. Peirce, and A. Cordsen, 2004, Enhanced HRAM anomalies correlate faults between 2D seismic lines: 74th Annual International Meeting, SEG, Extended Abstracts, 730. Grant, F. S., 1972, Review of data processing and interpreta-

Historical Development of Magnetic Method Hansen, R. O., and M. Simmonds, 1993, Multiple-source Werner deconvolution: Geophysics, 58, 17921800. Hansen, R. O., and L. Suciu, 2002, Multiple-source Euler deconvolution: Geophysics, 67, 525535. Hardwick, C. D., 1984, Important design considerations for inboard airborne magnetic gradiometers: Geophysics, 49, 20042018. Hartman, R. R., D. J. Teskey, and J. L. Friedberg, 1971, A system for rapid digital aeromagnetic interpretation: Geophysics, 36, 891918. Hassan, H. H., and J. W. Peirce, 2005, SAUCE: A new technique to remove cultural noise from HRAM: The Leading Edge, 24, 246 250. Hassan, H. H., J. W. Peirce, W. C. Pearson, and M. J. Pearson, 1998, Cultural editing of HRAM data: Comparison of techniques: Canadian Journal of Exploration Geophysics, 34, www.cseg.ca. Heezen, B. C., M. Ewing, and E. T. Miller, 1953, Trans-Atlantic prole of total magnetic intensity and topography: Dakar to Barbados: Deep Sea Research, 1, 2523. Heiland, C. A., 1935, Geophysical mapping from the air: Its possibilities and advantages: Engineering and Mining Journal, 136, 609610. , 1940, Geophysical Exploration: Prentice Hall. Heiland, C. A., and W. H. Courtier, 1929, Magnetometric investigation and gold placer deposits near Golden (Jefferson County), Colorado: Transactions of the American Institute of Mining, Metallurgical and Petroleum Engineers (AIME), 81, 364384. Henderson, R., 1960, A comprehensive system of automatic computation in magnetic and gravity interpretation: Geophysics, 25, 569 585. Henderson, R. G., and I. Zietz, 1948, Analysis of total magnetic intensity anomalies produced by point and line sources: Geophysics, 13, 428436. , 1949, The computation of second vertical derivatives of geomagnetic elds: Geophysics, 14, 508516. Hess, H. H., 1962, History of ocean basins, in A. E. J. Engel, H. L. James, and B. F. Leonard, eds., Petrologic studies: A volume to honor A. F. Buddington: Geological Society of America, 599620. Hildenbrand, T. G., and G. . Raines, 1990, Need for aeromagnetic data and a national airborne geophysics program: U. S. Geological Survey Bulletin 1924, 15. Hinze, W. J., 1985a, ed., Proceedings of the International Meeting on Potential Fields in Rugged Topography: Institut de Geophysique de Lausanne, Bulletin no. 7. de Universite , 1985b, ed., The utility of regional gravity and magnetic anomaly maps: SEG, 181197. Hjelt, S. E., 1972, Magnetostatic anomalies of a dipping prism: Geoexploration, 10, 239254. Hogg, R. L. S., 1979, Illustration of a new Northway process for eliminating the herringbone component from problem geophysical data: Annual Meeting of Canadian Society of Exploration Geophysicists. Holstein, H., 2002a, Gravimagnetic similarity in anomaly formulas for uniform polyhedra: Geophysics, 67, 11261133. Holstein, H., 2002b, Invariance in gravimagnetic anomaly formulas for uniform polyhedra: Geophysics, 67, 11341137. Hood, P. J., 1990, Aeromagnetic survey program of Canada, mineral applications, and vertical gradiometry, in W. F. Hanna, ed., Geologic applications of modern aeromagnetic surveys: U. S. Geological Survey Bulletin 1924, 723. Hornby, P., F. Boschetti, and F. G. Horowitz, 1999, Analysis of potential eld data in the wavelet domain: Geophysical Journal International, 137, 175196. Hoylman, H. W., 1961, How to determine and remove diurnal effects precisely: World Oil, December, 107112. Huang, D., and P. A. Versnel, 2000, Depth estimation algorithm applied to FTG data: 70th Annual International Meeting, SEG, Expanded Abstracts, 394397. Hutchison, R. D., 1958, Magnetic analysis by logarithmic curves: Geophysics, 23, 749769. Jacobsen, B., 1987, A case for upward continuation as a standard separation lter for potential-eld maps: Geophysics, 52, 11381148. Jacobson Jr., P., 1961, An evaluation of basement depth determinations from airborne magnetometer data: Geophysics, 26, 309317. Jain, S., 1976, An automatic method of direct interpretation of magnetic proles: Geophysics, 41, 531545. Jakosky, J. J., 1950, Exploration geophysics: Trija Publishing Company. Jenny, W. P., 1936, Micromagnetic surveys: Gulf Coast structures may be outlined by this new method: The Oil Weekly, April 27. Jensen, H., 1965, Instrument details and application of a new airborne magnetometer: Geophysics, 30, 875882. Jessell, M. W., 2001, Three-dimensional modeling of potential-eld data: Computers & Geosciences, 27, 455465.

57ND

Jessell, M. W., and Fractal Geophysics Pty Ltd., 2002, An atlas of structural geophysics II: Journal of the Virtual Explorer, 5, http://www.mssu.edu/seg-vm/exhibits/structuraatlas/index.html. Jessell, M. W., and R. K. Valenta, 1996, Structural geophysics: Integrated structural and geophysical mapping, in D. DePaor, ed., Structural geology and personal computers: Elsevier Science Publishing Co., 303324. Jessell, M. W., R. K. Valenta, G. Jung, J. P. Cull, and A. Gerio, 1993, Structural geophysics: Exploration Geophysics, 24, 599602. Johnson, R. C., T.M. Finn, and V. F. Nuccio, 2001, Potential for a basin-centered gas accumulation in the Albuquerque Basin, New Mexico: U. S. Geological Survey Bulletin 2184-C, available online at http://pubs.usgs.gov/bul/b2184-c/. Josephson, B. D., 1962, Possible new effect in superconductive tunneling: Physics Letters, 1, 251253. Keating, P., 1993, The fractal dimension of gravity data sets and its implication for gridding: Geophysical Prospecting, 41, 983994. , 1995, A simple technique to identify magnetic anomalies due to kimberlite pipes: Exploration Mining Geology, 4, 121125. Keating, P., and M. Pilkington, 2000, Euler deconvolution of the analytic signal: 62nd Annual International Meeting, EAGE, Session P0193. Keating, P., and L. Zerbo, 1996, An improved technique for reduction to the pole at low latitudes: Geophysics, 61, 131137. Kellogg, O. D., 1953, Foundations of potential theory: Dover Publications. Kilty, K. T., 1983, Werner deconvolution of prole potential eld data: Geophysics, 48, 234237. avuori, Korhonen, J. V., H. Sa and T. Koistinen, 2003, Petrophysical correlation of Fennoscandian magnetic and gravity anomalies: European Geophysical Society (EGS) American Geophysical Union (AGU)European Union of Geosciences (EUG) Joint Assembly, abstract #13230. Koulomzine, T., Y. Lamontagne, and A. Nadeau, 1970, New methods for the direct interpretation of magnetic anomalies caused by inclined dikes of innite length: Geophysics, 35, 812830. Ku, C. C., and J. A. Sharp, 1983, Werner deconvolution for automated magnetic interpretation and its renement using Marquardt inverse modeling: Geophysics, 48, 754774. Langel, R. A., 1992, International Geomagnetic Reference Field: The sixth generation: Journal of Geomagnetism and Geoelectricity, 44, 679707. Langel, R. A., and W. J. Hinze, 1998, The magnetic eld of the Earths lithosphere: The satellite perspective: Cambridge University Press. Langenheim, V. E., R. C. Jachens, D. M. Morton, R. W. Kistler, and J. C. Matti, 2004, Geophysical and isotopic mapping of preexisting crustal structures that inuenced the location and development of the San Jacinto fault zone, southern California: Geological Society of America Bulletin, 116, 11431157. Leblanc, G., and W. A. Morris, 2001, Denoising of aeromagnetic data via the wavelet transform: Geophysics, 66, 17931804. Leliak, P., 1961, Identication and evaluation of magnetic-eld sources of magnetic airborne detector equipped aircraft: IRE Transactions on Aerospace and Navigational Electronics, 8, 95106. Leu, L., 1982, Use of reduction-to-the-equator process for magnetic data interpretation: Geophysics, 47, 445 Levanto, A. E., 1959, A three-component magnetometer for small drill holes and its use in ore prospecting: Geophysical Prospecting, 7, 183195 Li, Y., and D. W. Oldenburg, 1996, 3-D inversion of magnetic data: Geophysics, 61, 394408. , 1998a, Separation of regional and residual magnetic eld data: Geophysics, 63, 431439. , 1998b, Stable reduction to the pole at the magnetic equator: 68th Annual International Meeting, SEG, Expanded Abstracts, 533536. , 1999, Rapid construction of equivalent sources using wavelets: 60th Annual International Meeting, SEG, Expanded Abstracts, 374377. , 2000a, Reduction to the pole using equivalent sources: 60th Annual International Meeting, SEG, Expanded Abstracts, 386 389. , 2000b, Joint inversion of surface and three-component borehole magnetic data: Geophysics, 65, 540552. , 2003, Fast inversion of large-scale magnetic data using wavelet transforms and logarithmic barrier method: Geophysical Journal International, 152, 251265. Logachev, A. A., 1946, The development and application of airborne magnetometers in the U.S.S.R.: Geophysics, 11, 135147. Lundberg, H., 1947, Results obtained by a helicopter borne magnetometer: Transactions, Canadian Institute of Mining and Metallurgy, 50, 392400.

58ND

Nabighian et al. Morley, L. W., 1963, The geophysics division of the Geological Survey of Canada: Bulletin of the Canadian Institute of Mining Metallurgy, 5, 358364. , 2001, The zebra pattern, in N. Orestes, ed., Plate tectonics: An insiders history of the modern theory of the Earth: Westview Press, 6785. Morley, L. W., and A. Larochelle, 1964, Paleomagnetism as a means of dating geological events: Royal Society of Canada Special Publication 8, 3950. Mushayandebvu, M., A. Reid, and D. Fairhead, 2000, Grid Euler deconvolution with constraints for 2-D structures: 70th Annual International Meeting, SEG, Expanded Abstracts, 398401. Mushayandebvu, M. F., P. van Driel, A. B. Reid, and J. D. Fairhead, 2001, Magnetic source parameters of two-dimensional structures using extended Euler deconvolution: Geophysics, 66, 814823. Mushayandebvu, M. F., V. Lesur, A. B. Reid, and J. D. Fairhead, 2004, Grid Euler deconvolution with constraints for 2D structures: Geophysics, 69, 489496. Nabighian, M. N., 1972, The analytic signal of two-dimensional magnetic bodies with polygonal cross-section Its properties and use for automated anomaly interpretation: Geophysics, 37, 507517. , 1974, Additional comments on the analytic signal of twodimensional magnetic bodies with polygonal cross-section: Geophysics, 39, 8592. , 1984, Toward a three-dimensional automatic interpretation of potential eld data via generalized Hilbert transforms Fundamental relations: Geophysics, 49, 780786. Nabighian, M. N., and R. O. Hansen, 2001, Unication of Euler and Werner deconvolution in three dimensions via the generalized Hilbert transform: Geophysics, 66, 18051810. Naudy, H., 1971, Automatic determination of depth on aeromagnetic proles: Geophysics, 36, 717722. Naudy, H., and H. Dreyer, 1968, Essai de ltrage non-lineaire ap aux proles aeromagnetiques (Attempt to apply nonlinear plique ltering to aeromagnetic proles): Geophysical Prospecting, 16, 171178. Nettleton, L. L., 1942, Gravity and magnetic calculations: Geophysics, 7, 293310. , 1971, Elementary gravity and magnetics for geologists and seismologists: SEG. North American Magnetic Anomaly Group, 2002, Magnetic anomaly map of North America: U. S. Geological Survey Special Map, available online at http://pubs.usgs.gov/sm/mag map/. OBrien, D. P., 1972, CompuDepth, a new method for depth-tobasement computation: Presented at the 42nd Annual International Meeting, SEG. OConnell, M. D., 2001, A heuristic method of removing micropulsations from airborne magnetic data: The Leading Edge, 20, 1242 1244. 2005, Gridding aeroOConnell, M. D., R. S. Smith, and M. A. Vallee, magnetic data using longitudinal and transverse gradients with the minimum curvature operator: The Leading Edge, 24, 142145. Okabe, M., 1979, Analytical expressions for gravity anomalies due to homogeneous polyhedral bodies and translations into magnetic anomalies: Geophysics, 44, 730741. Oldenburg, D. W., Y. Li, C. G. Farquharson, P. Kowalczyk, T. Aravanis, A. King, P. Zhang, and A. Watts, 1998, Applications of geophysical inversions in mineral exploration: The Leading Edge, 17, 461465. Olsen, N., R. Holm, G. Hulot, T. Sabaka, T. Neubert, L. TffnerClausen, F. Primdahl, et al., 2000, rsted initial eld model: Geophysical Research Letters, 27, 3607. Oreskes, N., 2001, Plate tectonics: An insiders history of the modern theory of the Earth: Westview Press. Parker, R. L., 1972, The rapid calculation of potential anomalies: Geophysical Journal of the Royal Astronomical Society, 31, 447455. Parker, R. L., and S. P. Huestis, 1974, Inversion of magnetic anomalies in the presence of topography: Journal of Geophysical Research, 79, 15871593. Parker, R. L., and K. D. Klitgord, 1972, Magnetic upward continuation from an uneven track: Geophysics, 37, 662668. Paterson, N. R., and C. V. Reeves, 1985, Applications of gravity and magnetic surveys The state of the art in 1985: Geophysics, 50, 25582594. Pawlowski, R. S., and R. O. Hansen, 1990, Gravity anomaly separation by Wiener ltering: Geophysics, 55, 539548. Pearson, W. C., 1996, Removing culture from southern Texas A magnetic clean-up and imaging revolution: 66th Annual International Meeting, SEG, Expanded Abstracts, 14071410. , 2001, Finding faults in a gas play: AAPG Explorer, May, 5255.

Macdonald, K. C., S. P. Miller, S. P. Huestis, and F. N. Spiess, 1980, Three-dimensional modeling of a magnetic reversal boundary from inversion of deep-tow measurements: Journal of Geophysical Research, 85, 36703680. Machel, H. G., and E. A. Burton, 1991, Chemical and microbial processes causing anomalous magnetization in environments affected by hydrocarbon seepage: Geophysics, 56, 598605. MacLean, B. C., and D. G. Cook, 2002(updated), Subsurface and surface distribution of Proterozoic units, northwestern NWT: a Cambrian sub-crop map: Geological Survey of Canada, Open File 3502. Macmillan, S., S. Maus, T. Bondar, A. Chambodut, V. Golovkov, R. Holme, B. Langlais et al., 2003, Ninth generation International Geomagnetic Reference Field released: EOS Transactions of the American Geophysical Union, 84, 503. Macnae, J. C., 1979, Kimberlites and exploration geophysics: Geophysics, 44, 13951416. Mason, R. G., 1958, A magnetic survey off the west coast of the United States between latitudes 30 and 36 N, longitudes 121 and 128 W: Geophysical Journal of the Royal Astronomical Society, 1, 320329. Maus, S., 1999, Variogram analysis of magnetic and gravity data: Geophysics, 64, 776784. Maus, S., K. P. Sengpiel, B. Rottger, and E. A. W. Tordiffe, 1999, Variogram analysis of helicopter magnetic data to identify paleochannels of the Omaruru River, Namibia: Geophysics, 64, 785794. Maus, S., and S. Macmillan, 2005, 10th generation International Geomagnetic Reference Field: EOS Transactions of the American Geophysical Union, 86, 159. Maxwell, A. E., R. P. von Herzen, et al., 1970, Initial Reports of the Deep Sea Drilling Project; covering Leg 3 of the cruises of the drilling vessel Glomar Challenger, Dakar, Senegal to Rio de Janeiro, Brazil, December 1968 to January 1969: Deep Sea Drilling Project: U. S. Government Printing Ofce. Mazur, M. J., R. R. Stewart, and A. R. Hildebrand, 2000, The seismic signature of meteorite impact craters: Canadian Society of Exploration Geophysicists Recorder, 35, June, 1016. McConnell, T. J., B. Lo, A. Ryder-Turner, and J. A. Musser, 1999, Enhanced 3D seismic surveys using a new airborne pipeline mapping system: 69th Annual International Meeting, SEG, Expanded Abstracts, 516519. McElhinny, M. W., 1973, Paleomagnetics and plate tectonics: Cambridge University Press. McIntyre, J. I., 1980, Geological signicance of magnetic patterns related to magnetite in sediments and metasediments A review: Bulletin of the Australian Society of Exploration Geophysicists, 11, 1933. Mendonc a, C. A., and J. B. C. Silva, 1993, A stable truncated series approximation of the reduction-to-the-pole operator: Geophysics, 58, 10841090. , 1994, The equivalent data concept applied to the interpolation of potential-eld data: Geophysics, 59, 722732. , 1995, Interpolation of potential-eld data by equivalent layer and minimum curvature: A comparative analysis: Geophysics, 60, 399407. Mesko, A., 1965, Some notes concerning the frequency analysis for gravity interpretation: Geophysical Prospecting, 13, 475488. Millegan, P. S., 1998, High-resolution aeromagnetic surveying, in R. I. Gibson, and P. S. Millegan, eds., Geologic applications of gravity and magnetics: Case histories: SEG and AAPG. Miller, H. G., and V. Singh, 1994, Potential eld tilt; a new concept for location of potential eld sources: Journal of Applied Geophysics, 32, 213217. Minty, B. R. S., 1991, Simple microlevelling for aeromagnetic data: Exploration Geophysics, 22, 591592. Misener, D. J., F. S. Grant, and P. Walker, 1984, Variable depth, space-domain magnetic susceptibility mapping: 54th Annual International Meeting, SEG, Expanded Abstracts, 237. Mittal, P. K., 1984, Algorithm for error adjustment of potential-eld data along a survey network: Geophysics, 49, 467469. Modisi, M. P., E. A. Atekwana, A. B. Kampunzu, and T. H. Ngwisanyi, 2000, Rift kinematics during the incipient stages of continental extension: Evidence from the nascent Okavango rift basin, northwest Botswana: Geology, 28, 939942. Moore, D. H., 2005, Swan Hill 1:250,000 and parts of Balranald and Deniliquin 1:250,000 map areas: A geological interpretation of the geophysical data: Victorian Initiative for Minerals and Petroleum Report 84, Department of Primary Industries, Victoria, Australia. Moreau, F., D. Gibert, M. Holschneider, and G. Saracco, 1997, Wavelet analysis of potential elds: Inverse Problems, 13, 165178. Morgan, R., 1998, Magnetic anomalies associated with the North and South Morecambe Fields, U. K., in R. I. Gibson, and P. R. Milligan, eds., Geologic applications of gravity and magnetics: Case histories: SEG and AAPG, 8991.

Historical Development of Magnetic Method Pearson, W. C., and C. M. Skinner, 1982, Reduction-to-the-pole of low latitude magnetic anomalies: 52nd Annual International Meeting, SEG, Expanded Abstracts, 356. Peddie, N. W., 1982, International Geomagnetic Reference Field: The third generation: Journal of Geomagnetism and Geoelectricity, 34, 309326. , 1983, International Geomagnetic Reference Field Its evolution and the difference in total eld intensity between new and old models for 19651980: Geophysics, 48, 16911696. Pedersen, L. B., 1977, Interpretation of potential eld data A generalized inverse approach: Geophysical Prospecting, 25, 199230. , 1978, Wavenumber domain expressions for potential elds -, and 3-dimensional bodies: Geophysics, 43, from arbitrary 2-, 2 1 2 626630. , 1979, Constrained inversion of potential eld data: Geophysical Prospecting, 27, 726748. Peirce, J. W., W. E. Glenn, and K. Brown, eds., 1998, High resolution aeromagnetics for hydrocarbon exploration: Canadian Journal of Exploration Geophysics, 34, available online at www.cseg.ca. Peirce, J. W., S. A. Goussev, R. McLean, and M. Marshall, 1999, Aeromagnetic interpretation of the Dianango Trough HRAM survey, onshore Gabon: 69th Annual International Meeting, SEG, Extended Abstracts, 343346. Peters, L. J., 1949, The direct approach to magnetic interpretation and its practical application: Geophysics, 14, 290320. Phillips, J. D., 1979, ADEPT, A program to estimate depth to magnetic basement from sampled magnetic proles: U. S. Geological Survey Open File Report No. 79-367. , 2000, Locating magnetic contacts: A comparison of the horizontal gradient, analytic signal, and local wavenumber methods: 70th Annual International Meeting, SEG, Expanded Abstracts, 402405. , 2001, Designing matched bandpass and azimuthal lters for the separation of potential-eld anomalies by source region and source type: 15th Geophysical Conference and Exhibition, Australian Society of Exploration Geophysicists, Expanded Abstracts, CD-ROM. , 2002, Two-step processing for 3D magnetic source locations and structural indices using extended Euler or analytic signal methods: 72nd Annual International Meeting, SEG, Expanded Abstracts, 727730. Phillips, J. D., R. W. Saltus, and R. L. Reynolds, 1998, Sources of magnetic anomalies over a sedimentary basin: Preliminary results from the Coastal Plain of the Arctic National Wildlife Refuge, Alaska, in R. I. Gibson and P. S. Millegan, eds., Geologic applications of gravity and magnetics: Case histories: SEG and AAPG, 130134. Pilkington, M., 1997, 3-D magnetic imaging using conjugate gradients: Geophysics, 62, 11321142. Pilkington, M., and D. J. Crossley, 1986, Determination of crustal interface topography from potential elds: Geophysics, 51, 1277 1284. Pilkington, M., and P. Keating, 2004, Contact mapping from gridded magnetic data A comparison of techniques: Exploration Geophysics, 35, 306311. Pilkington, M., M. E. Gregotski, and J. P. Todoeschuck, 1994, Using fractal crustal magnetization models in magnetic interpretation: Geophysical Prospecting, 42, 677692. Plouff, D., 1975, Derivation of formulas and FORTRAN programs to compute magnetic anomalies of prisms: National Technical Information Service No. PB-243-525, U. S. Department of Commerce. , 1976, Gravity and magnetic elds of polygonal prisms and application to magnetic terrain corrections: Geophysics, 41, 727741. Power, M., G. Belcourt, and E. Rockel, 2004, Geophysical methods for kimberlite exploration in northern Canada: The Leading Edge, 23, 1124 Prieto, C., and G. Morton, 2003, New insights from a 3D earth model, deepwater Gulf of Mexico: The Leading Edge, 22, 356360. Pustisek, A. M., 1990, Noniterative three-dimensional inversion of magnetic data: Geophysics, 55, 782785. Rasmussen, R., and L. B. Pedersen, 1979, End corrections in potential eld modeling: Geophysical Prospecting, 27, 749760. Ravat, D., T. G. Hildenbrand, and W. Roest, 2003, New way of processing near-surface magnetic data: The utility of the Comprehensive Model of the magnetic eld: The Leading Edge, 22, 784785. Reford, M. S., 1980, History of geophysical exploration Magnetic method: Geophysics, 45, 16401658. Reford, M. S., and J. S. Sumner, 1964, Aeromagnetics: Geophysics, 29, 482516. Reid, A. B., 1980, Aeromagnetic survey design: Geophysics, 45, 973 976. Reid, A. B., J. M. Allsop, H. Granser, A. J. Millett, and I. W. Somerton, 1990, Magnetic interpretation in three dimensions using Euler deconvolution: Geophysics, 55, 8091.

59ND

Reigber, C., H. Luehr, and P. Schwintzer, 2002, CHAMP mission status: Advances in Space Research, 30, 289293. Reynolds, R. L., J. G. Rosenbaum, M. R. Hudson, and N. S. Fishman, 1990, Rock magnetism, the distribution of magnetic minerals in the Earths crust, and aeromagnetic anomalies: U. S. Geological Survey Bulletin1924, 2445. Reynolds, R. L., M. Webring, V. J. S. Grauch, and M. Tuttle, 1990, Magnetic forward models of Cement oil eld, Oklahoma, based on rock magnetic, geochemical and petrologic constraints: Geophysics, 55, 344353. Reynolds, R. L., N. S. Fishman, and M. R. Hudson, 1991, Sources of aeromagnetic anomalies over Cement oil eld (Oklahoma), Simpson oil eld (Alaska), and the Wyoming-Idaho-Utah thrust belt: Geophysics, 56, 606617. Rhodes, J., and J. W. Peirce, 1999, MaFIC Magnetic interpretation in 3-D using a seismic workstation: 69th Annual International Meeting, SEG, Expanded Abstracts, 335338. Ridsdill-Smith, T. A., 1998a, Separating aeromagnetic anomalies using wavelet matched lters: 68th Annual International Meeting, SEG, Expanded Abstracts, 550553. , 1998b, Separation ltering of aeromagnetic data using lterbanks: Exploration Geophysics, 29, 577583. , 2000, Wavelet compression of an equivalent layer: 70th Annual International Meeting, SEG, Expanded Abstracts, 378381. Ridsdill-Smith, T. A., and M. C. Dentith, 1999, The wavelet transform in aeromagnetic processing: Geophysics, 64, 10031013. Robson, D. F., and R. Spencer, 1997, The New South Wales governments Discovery 2000 Geophysical surveys and their effect on exploration: Exploration Geophysics, 28, 296298. Roest, W. R., and M. Pilkington, 1993, Identifying remanent magnetization effects in magnetic data: Geophysics, 58, 653659. Roest, W. R., J. Verhoef, and M. Pilkington, 1992, Magnetic interpretation using the 3D analytic signal: Geophysics, 57, 116125. Rosenbaum, J. G., and D. B. Snyder, 1985, Preliminary interpretation of paleomagnetic and magnetic property data from drill holes USW G-1, G-2, GU-3, G-3, and VH-1 and surface localities in the vicinity of Yucca Mountain, Nye County, Nevada: U. S. Geological Survey Open File Report 85-49. Ross, G. M., J. Broome, and W. Miles., 1994, Potential elds and basement structure: Western Canada Sedimentary Basin, in G. D. Mossop and I. Shetsen, comps., Geological atlas of the Western Canada Sedimentary Basin: Canadian Society of Petroleum Geologists and Alberta Research Council, 4146. Roux, A. T., 1970, The application of geophysics to gold exploration in South Africa, in L. W. Morley, ed., Mining and groundwater geophysics, 1967: Geological Survey of Canada, Economic Geology Report No. 26, 425438. Roy, A., 1958, Letter on residual and second derivative of gravity and magnetic maps: Geophysics, 23, 860861. Roy, P. S., J. Whitehouse, P. J. Cowell, and G. Oakes, 2000, Mineral sands occurrences in the Murray Basin, southeastern Australia: Economic Geology, 95, 11071128. Sabaka, T. J., N. Olsen, and R. A. Langel, 2002, A comprehensive model of the quiet-time, near-the-earth magnetic eld: Phase 3: Geophysical Journal International, 151, 3268. Sabaka, T. J., N. Olsen, and M. E. Purucker, 2004, Extending comprehensive models of the Earths magnetic eld with Oersted and Champ data: Geophysical Journal International, 159, 521547. Sailhac, P., A. Galdeano, D. Gibert, F. Moreau, and C. Delor, 2000, Identication of sources of potential elds with the continuous wavelet transform: Complex wavelets and application to aeromagnetic proles in French Guiana: Journal of Geophysical Research, 105, 1945519475. Salem, A., and D. Ravat, 2003, A combined analytic signal and Euler method (AN-EUL) for automatic interpretation of magnetic data: Geophysics, 68, 19521961. Saltus, R. W., and R. J. Blakely, 1983, HYPERMAG, an interactive, two-dimensional gravity and magnetic modeling program: U. S. Geological Survey Open File Report 83-241. /2 -dimensional , 1993, HYPERMAG, an interactive, 2- and 221 gravity and magnetic modeling program, version 3.5: U. S. Geological Survey Open File Report93287. Saltus, R. W., P. J. Haeussler, and J. D. Phillips, 2001, Geophysical mapping of subsurface structures in the upper Cook Inlet basin, Alaska: Geological Society of America, Abstracts with Programs, 33, 345. Shearer, S., and Y. Li., 2004, 3D inversion of magnetic total gradient data in the presence of remanent magnetization: 74th Annual International Meeting, SEG, Expanded Abstracts, 774777. Shi, Z., 1991, An improved Naudy-based technique for estimating depth from magnetic proles: Exploration Geophysics, 22, 357 362.

60ND

Nabighian et al. , 1963b, Discussion on An evaluation of basement depth determinations from airborne magnetometer data, by Peter Jacobsen, Jr. (GEO-26-03-0309-0319): Geophysics, 28, 491492. , 1965, Oil elds and aeromagnetic anomalies: Geophysics, 30, 706739. , 1998, Reecting on exploration in the North Sea in the 1960s: The Leading Edge, 17, 479482. Stone, V. C. A., J. D. Fairhead, and W. H. Oterdoom, 2004, Micromagnetic seep detection in the Sudan: The Leading Edge, 23, 734 737. Stuart, W. F., 1972, Earths eld magnetometry: Reports of Progress in Physics, 35, 803881. Sweeney, R. E., V. J. S. Grauch, and J. D. Phillips, 2002, Merged digi tal aeromagnetic data for the Albuquerque and southern Espanola Basins, New Mexico: U. S. Geological Survey Open File Report 02205. Syberg, F. J. R., 1972a, A Fourier method for the regional-residual problem of potential elds: Geophysical Prospecting, 20, 4775. , 1972b, Potential eld continuation between general surfaces: Geophysical Prospecting, 20, 267282. Talwani, M., 1965, Computation with the help of a digital computer of magnetic anomalies caused by bodies of arbitrary shape: Geophysics, 30, 797817. Talwani, M., and J. R. Heirtzler, 1964, Computation of magnetic anomalies caused by two-dimensional structures of arbitrary shape: Stanford University Publications of the Geological Sciences, Computers in the Mineral Industries. Telford, W. M., L. P. Geldart, and R. E. Sheriff, 1990, Applied geophysics, 2nd ed.: Cambridge University Press. Teskey, D. J., P. J. Hood, L. W. Morley, R. A. Gibb, P. Sawatzky, M. Bower, and E. E. Ready, 1993, The aeromagnetic survey program of the Geological Survey of Canada; contribution to regional geological mapping and mineral exploration: Canadian Journal of Earth Sciences, 30, 243260. Thompson, D. T., 1982, EULDPH A new technique for making computer-assisted depth estimates from magnetic data: Geophysics, 47, 3137. Thompson, R., V. J. S. Grauch, D. Sawyer, and M. R. Hudson, 2002, Aeromagnetic expression of volcanic rocks of the Cerros Del Rio volcanic eld, Rio Grande rift, north-central New Mexico: Geological Society of America, Abstracts with Programs, 34, 451452. Thurston, J. B., and R. J. Brown, 1994, Automated source-edge location with a new variable-pass horizontal gradient operator: Geophysics, 59, 546554. Thurston, J. B., and R. S. Smith, 1997, Automatic conversion of magnetic data to depth, dip, and susceptibility contrast using the SPITM method: Geophysics, 62, 807813 Thurston, J., J.-C. Guillon, and R. Smith, 1999, Model-independent depth estimation with the SPITM method: 69th Annual International Meeting, SEG, Expanded Abstracts, 403406. Thurston, J. B., R. S. Smith, and J.-C. Guillon, 2002, A multimodel method for depth estimation from magnetic data: Geophysics, 67, 555561. Treitel, S., W. G. Clement, and R. K. Kaul, 1971, The spectral determination of depths to buried magnetic basement rocks: Geophysical Journal of the Royal Astronomical Society, 24, 415428. Tsokas, G. N., and C. B. Papazachos, 1992, Two-dimensional inversion lters in magnetic prospecting: Application to the exploration for buried antiquities: Geophysics, 57, 10041013. Tsokas, G. N., and R. O. Hansen, 1996, A comparison between inverse ltering and multiple-source Werner deconvolution: 66th Annual International Meeting, SEG, Expanded Abstracts, 11531156. Urquhart, T., 1988, Decorrugation of enhanced magnetic eld maps: 58th Annual International Meeting, SEG, Expanded Abstracts, 371372. Vacquier, V., N C. Steenland, R. G. Henderson, and I. Zietz, 1951, Interpretation of aeromagnetic maps: Geological Society of America, Memoir 47. M. A., P. Keating, R. S. Smith, and C. St-Hilaire, 2004, EstiVallee, mating depth and model type using the continuous wavelet transform of magnetic data: Geophysics, 69, 191199. Verduzco, B., J. D. Fairhead, C. M. Green, and C. MacKenzie, 2004, New insights into magnetic derivatives for structural mapping: The Leading Edge, 23, 116119. Vine, F. J., 2001, Reversals of fortune, in N. Orestes, ed., Plate tectonics: An insiders history of the modern theory of the Earth: Westview Press, 4666. Vine, F. J., and D. H. Matthews, 1963, Magnetic anomalies over oceanic ridges: Nature, 199, 947949. Wang, X., and R. O. Hansen, 1990, Inversion for magnetic anomalies of arbitrary three-dimensional bodies: Geophysics, 55, 13211326.

Shuey, R. T., and A. S. Pasquale, 1973, End corrections in magnetic prole interpretation: Geophysics, 38, 507512. Silva, J. B. C., 1986, Reduction to the pole as an inverse problem and its application to low-latitude anomalies: Geophysics, 51, 369382. Silva, J. B. C., and G. W. Hohmann, 1981, Interpretation of threecomponent borehole magnetometer data: Geophysics, 46, 1721 1731. , 1983, Nonlinear magnetic inversion using a random search method: Geophysics, 48, 16451658. , 1984, Airborne magnetic susceptibility mapping: Exploration Geophysics, 15, 113. Silva, J. B. C., and V. C. F. Barbosa, 2003, Euler deconvolution: Theoretical basis for automatically selecting good solutions: Geophysics, 68, 19621968. Skeels, D. C., 1967, What is residual gravity?: Geophysics, 32, 872 876. Slichter, L. B., 1929, Certain aspects of magnetic surveying: American Institute of Mining and Metallurgical Engineers, Transactions, 81, 238260. Smellie, D. W., 1956, Elementary approximations in aeromagnetic interpretation: Geophysics, 21, 10211040. Smith, B. D., A. E. McCafferty, and R. R. McDougal, 2000, Utilization of airborne magnetic, electromagnetic, and radiometric data in abandoned mine land investigations, in S. E. Church, ed., Preliminary release of scientic reports on the acidic drainage in the Animas River watershed, San Juan County, Colorado: U. S. Geological Survey Open File Report 00-0034, 8691. Smith, D. V., and D. Pratt, 2003, Advanced processing and interpretation of the high resolution aeromagnetic survey data over the Central Edwards Aquifer, Texas: Proceedings from the Symposium on the Application of Geophysics to Engineering and Environmental Problems, Environmental and Engineering Society. Smith, R. A., 1959, Some depth formulae for local magnetic and gravity anomalies: Geophysical Prospecting, 7, 5563. Smith, R. P., V. J. S. Grauch, and D. D. Blackwell, 2002, Preliminary results of a high-resolution aeromagnetic survey to identify buried faults at Dixie Valley, Nevada: Geothermal Resources Council Transactions, 26, 543546. Smith, R. S., J. G. Thurston, T.-F. Dai, and I. N. MacLeod, 1998, ISPITM the improved source parameter imaging method: Geophysical Prospecting, 46, 141151. Smock, J. C., 1876, The use of the magnetic needle in searching for magnetic iron ore: Transactions, American Institute of Mining and Metallurgical Engineers (AIME), 4, 353362. Smyth, H. L., 1896, Magnetic observations in geological mapping: American Institute of Mining and Metallurgical Engineers, Transactions, 26, 640709. Spaid-Reitz, M., and P. M. Eick, 1998, HRAM as a tool for petroleum system analysis and trend exploration: A case study of the Mississippi Delta survey, southeast Louisiana: Canadian Journal of Exploration Geophysics, 34, 8396. Spector, A., 1968, Spectral analysis of aeromagnetic data: Ph.D. thesis, University of Toronto. , 1971, Aeromagnetic map interpretation with the aid of the digital computer: Canadian Institute of Mining, Metallurgy and Petroleum Bulletin, 64, 711, 2733. Spector, A., and B. K. Bhattacharyya, 1966, Energy density spectrum and autocorrelation function of anomalies due to simple magnetic models: Geophysical Prospecting, 14, 242272. Spector, A., and F. Grant, 1970, Statistical models for interpreting aeromagnetic data: Geophysics, 35, 2933302. , 1974, Reply to the discussion by G. Gudmunddson on Statistical models for interpreting aeromagnetic data: Geophysics, 39, 112113. Spector, A., and W. Parker, 1979, Computer compilation and interpretation of geophysical data, in P. J. Hood, ed., Geophysics and geochemistry in the search for metallic ores: Geological Survey of Canada, Economic Report 31, 527544. Stavrev, P. Y., 1997, Euler deconvolution using differential similarity transformations of gravity or magnetic anomalies: Geophysical Prospecting, 45, 207246. Stearn, N. H., 1929a, A background for the application of geomagnetics to exploration: Transactions of the American Institute of Mining and Metallurgical Engineers, 81, 315345. , 1929b, The dip needle as a geological instrument: Transactions of the American Institute of Mining and Metallurgical Engineers, 81, 345363. Steenland, N. C., 1962, Gravity and aeromagnetic exploration in the Paradox basin: Geophysics, 27, 7389. , 1963a, An evaluation of the Peace River aeromagnetic interpretation: Geophysics, 28, 745755.

Historical Development of Magnetic Method Wantland, D., 1944, Magnetic interpretation: Geophysics, 9, 4758. Watts, A., 1997, Exploration for nickel in the 90s, or Til depth us do part: in A. G. Gubins, ed., Proceedings of Fourth Decennial International Conference on Mineral Exploration, Prospectors and Developers Association of Canada, 1003. Webring, M., 1985, SAKI, A Fortran program for generalized linear inversion of gravity and magnetic proles: U. S. Geological Survey Open File Report 85-122. Weinstock, H., and W. C. Overton, eds., 1981, SQUID applications to geophysics: SEG. Werner, S., 1955, Interpretation of magnetic anomalies of sheet-like bodies, Sveriges Geologiska Undersokning, Series C, Arsbok 43, No. 6. Whitehill, D. E., 1973, Automated interpretation of magnetic anomalies using the vertical prism model: Geophysics, 38, 10701087. Whitham, K., and E. R. Niblett, 1961, The diurnal problem in aeromagnetic surveying in Canada: Geophysics, 26, 211228. Wilson, M., and M. Monea, 2004, IEA GHG Weyburn CO2 Moni-

61ND

toring & Storage Project Summary Report20002004: Proceedings of the 7th International Conference on Greenhouse Gas Control Technologies: Petroleum Technology Research Centre. Wilson, C. R., G. Tsoias, and M. Bartelmann, 1997, A high-precision aeromagnetic survey near the Glen Hummel Field in Texas; Identication of cultural and sedimentary anomaly sources: The Leading Edge, 16, 3744. Zhang, C., M. F. Mushayandebvu, A. B. Reid, J. D. Fairhead, and M. E. Odegard, 2000, Euler deconvolution of gravity tensor gradient data: Geophysics, 65, 512520. Zietz, I., comp., 1982, Composite magnetic anomaly map of the United States; Part A, Conterminous United States: U. S. Geological Survey Geophysical Investigations Map GP-954-A. Zietz, I., and G. E. Andreasen, 1967, Remanent magnetization and aeromagnetic interpretation, Mining geophysics, vol. II: Theory: SEG, 569590. Zurueh, E. G., 1967, Applications of two-dimensional linear wavelength ltering: Geophysics, 32, 10151035.

You might also like