14.11.2013 Views

Magnetotelluric data analysis using advances in ... - Geomagnetism

Magnetotelluric data analysis using advances in ... - Geomagnetism

Magnetotelluric data analysis using advances in ... - Geomagnetism

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

.<br />

<strong>Magnetotelluric</strong> <strong>data</strong> <strong>analysis</strong> <strong>us<strong>in</strong>g</strong><br />

<strong>advances</strong> <strong>in</strong> signal process<strong>in</strong>g<br />

techniques<br />

C. MANOJ<br />

National Geophysical Research Institute, Hyderabad – 500 007,<br />

India<br />

.<br />

THESIS SUBMITTED TO THE OSMANIA UNIVERSITY FOR THE<br />

AWARD OF THE DEGREE OF DOCTOR OF PHILOSOPHY IN<br />

GEOPHYSICS 2003


ii<br />

Declaration<br />

I, hereby declare that the thesis submitted for the award of the Degree of Doctor of<br />

Philosophy <strong>in</strong> Geophysics of the Osmania University, Hyderabad, India is orig<strong>in</strong>al <strong>in</strong> its<br />

contents and has not been submitted before, either <strong>in</strong> parts or <strong>in</strong> full to any University<br />

for any research degree.<br />

(C. Manoj)<br />

Candidate<br />

Dr. V. P. Dimri<br />

Dr. Nand<strong>in</strong>i Nagarajan<br />

Director<br />

Research Supervisor<br />

National Geophysical Research Institute Scientist<br />

Hyderabad<br />

NGRI<br />

Hyderabad


iii<br />

Certificate<br />

This is to certify that the thesis, entitled ‘<strong>Magnetotelluric</strong> <strong>data</strong> <strong>analysis</strong> <strong>us<strong>in</strong>g</strong> <strong>advances</strong><br />

<strong>in</strong> signal process<strong>in</strong>g techniques’, which is submitted for the award of the Degree of Doctor<br />

of Philosophy <strong>in</strong> Geophysics to the Osmania University, Hyderabad, India, is the bonafide<br />

research work carried out by Mr. C. Manoj at National Geophysical Research Institute,<br />

Hyderabad, India dur<strong>in</strong>g the years 1998 to 2003 under my supervision. The work is<br />

orig<strong>in</strong>al and has not been submitted for any Degree at this or any other University<br />

20-08-2003<br />

(Nand<strong>in</strong>i Nagarajan)<br />

Research Supervisor<br />

Scientist<br />

NGRI<br />

Hyderabad 500007<br />

India


iv<br />

Acknowledgement<br />

I s<strong>in</strong>cerely thank Dr. Nand<strong>in</strong>i Nagarajan for the support and guidance she extended<br />

to my research program as a research supervisor, over the past five years of my work at<br />

National Geophysical Research Institute. As a mentor and group head of <strong>Magnetotelluric</strong>s<br />

Division, Dr S. V. S. Sarma gave me moral and <strong>in</strong>stitutional support <strong>in</strong> the early phases<br />

of my Ph.D program. The present group head Dr. T. Har<strong>in</strong>arayana also adopted the<br />

same approach. Dr. H.K. Gupta and Dr. V.P. Dimri, the former and present Directors<br />

of NGRI, k<strong>in</strong>dly encouraged me to pursue the research work.<br />

Prof. (late) P.S. Moharir and Mr. G. Virupakshi are profusely thanked for the hours<br />

of discussion they had with me. Understand<strong>in</strong>g my limitations, they fixed many bugs<br />

<strong>in</strong> my computer codes as well as knowledge. The staff of the offices of Dean, Faculty of<br />

Science and Head, Department of Geophysics, Osmania University is thanked for their<br />

co-operation throughout the research program. The Senior colleagues of magnetotellurics<br />

group: Mr. D.N. Murthy, Dr. R.S. Sastry, Dr. M. Someswara Rao, Mr. M.V.C. Sarma,<br />

Dr. Madhusudan Rao, Dr. K. Veeraswamy and Mr. S. Prabhakar E. Rao are thanked for<br />

<strong>in</strong>itiat<strong>in</strong>g me <strong>in</strong> magnetotellurics and their help dur<strong>in</strong>g field campaigns. As immediate<br />

colleagues, I acknowledge the support of Dr. B.P.K. Patro, Dr. K. Naganjaneyulu, Dr.<br />

K. Begum, and Mr. K.K. Abdul Azeez for their cooperation and support. F<strong>in</strong>ancial<br />

support from the follow<strong>in</strong>g agencies is also acknowledged.<br />

The magnetotelluric <strong>data</strong> for the present studies were acquired <strong>in</strong> a project funded by<br />

Department of Science & Technology, Government of India (No. ESS/16/118/1997). I<br />

received the Junior Research Fellowship of Council of Scientific and Industrial Research<br />

(No. 2-31/97(i)-E.U.II) dur<strong>in</strong>g the <strong>in</strong>itial phase of the research program. Suggestions<br />

by Dr. Martyn Unsworth, Dr. John Stodt and Dr. Xavier Garcia greatly improved<br />

a part of the thesis. I take this opportunity to thank my friends Dr. Shyam Chand,<br />

Dr. Jimmy Stephen, Dr. R. S. Rajesh and Mr. Tomson J.K for their comradeship that<br />

helped me <strong>in</strong> different stages of the research program. I thank my wife Mrs Namitha Vaz,<br />

daughter Meenakshi, father Mr. M.K.C. Nair and mother Mrs. M.N. Saroj<strong>in</strong>i Amma for<br />

the patience and support they extended to me dur<strong>in</strong>g my research program.


v<br />

<strong>Magnetotelluric</strong> <strong>data</strong> <strong>analysis</strong> <strong>us<strong>in</strong>g</strong><br />

<strong>advances</strong> <strong>in</strong> signal process<strong>in</strong>g techniques<br />

SYNOPSIS<br />

Electrical conductivity is one of the most important physical parameters directly <strong>in</strong>dicat<strong>in</strong>g<br />

the Earth’s subsurface nature. Rocks exhibit a wide range (∼ 10 6 ) of conductivity<br />

and the electrical conductivity of rocks is sensitive to temperature, presence of fluid,<br />

volatiles, melt as well as its bulk composition. These qualities make electrical conductivity<br />

an appropriate method to del<strong>in</strong>eate Earth’s subsurface features with a suitable<br />

measurement at surface. Over the past century a suite of electrical and electromagnetic<br />

methods was established to probe electrical conductivity structure of the Earth. Among<br />

them, the natural source electromagnetic method - <strong>Magnetotelluric</strong>s has many advantages<br />

over all the other methods: with the sk<strong>in</strong> depth relation of EM waves it can virtually<br />

probe any depth and the natural electromagnetic signals have enough power over a wide<br />

range of frequencies to penetrate the subsurface (Cagniard [1953]). Thus magnetotelluric<br />

(MT) method has become a promis<strong>in</strong>g technique to probe deep earth structure. The<br />

magnetotelluric <strong>in</strong>duction, though diffusive <strong>in</strong> nature like potential fields, is not <strong>in</strong>herently<br />

non- unique like potential fields methods. The uniqueness of this method <strong>in</strong> cases<br />

where conductivity varies vertically (1-Dimensional) is established (Bailey [1970], Weidelt<br />

[1972]) <strong>in</strong> theory. Non-uniqueness <strong>in</strong> magnetotellurics may results from 1) errors <strong>in</strong><br />

measurement 2) f<strong>in</strong>ite frequency range of <strong>data</strong> and 3) sparse distribution of measurement<br />

sites. MT practitioners, therefore, strive to obta<strong>in</strong> more precise <strong>data</strong> <strong>in</strong> wider bandwidth<br />

and with more spatial density. As a result remarkable progress has been made to <strong>in</strong>strumentation<br />

(Clarke et al. [1983], Ritter et al. [1998]) and time series process<strong>in</strong>g (Gamble<br />

et al. [1979], Egbert and Booker [1986], Chave and Thomson [1989]) <strong>in</strong> the past two<br />

decades that resulted <strong>in</strong> better estimation of full tensor (transfer function) elements and<br />

their variance over a wide band of frequency. Together with the progress of decompos<strong>in</strong>g<br />

the impedance tensor <strong>in</strong>to regional and residual (Groom and Bailey [1989], McNeice and<br />

Jones [2001]) and the <strong>advances</strong> <strong>in</strong> <strong>in</strong>version (Constable et al. [1987], Smith and Booker<br />

[1991], Siripunvaraporn and Egbert [2000], Rodi and Mackie [2001]), magnetotellurics<br />

has become a standard tool to determ<strong>in</strong>e Earth’s electrical conductivity structure.<br />

The first step <strong>in</strong> <strong>in</strong>terpretation of magnetotelluric <strong>data</strong> is to estimate the MT<br />

impedance tensor <strong>in</strong> frequency doma<strong>in</strong> from measured magnetotelluric time series. <strong>Magnetotelluric</strong><br />

time series consist of five simultaneously measured components of earth’s<br />

electromagnetic field viz. two orthogonal components of horizontal electric fields (Ex,<br />

Ey <strong>in</strong> mV/ km) and three orthogonal components of magnetic field (Hx, Hy, Hz <strong>in</strong> nT).<br />

Usually the measurements are done <strong>in</strong> wide bands of overlapp<strong>in</strong>g frequency ranges, with<br />

different sampl<strong>in</strong>g <strong>in</strong>tervals. Measured time series are the compounded effect of various<br />

signal and noise processes <strong>in</strong> the frequency range of <strong>in</strong>terest to MT. The objective of MT<br />

process<strong>in</strong>g is to discrim<strong>in</strong>ate signal and m<strong>in</strong>imize the effect of noise <strong>in</strong> the estimation of<br />

MT transfer function tensor Z . The use of traditional spectral <strong>analysis</strong> together with<br />

least squares (LS) estimation is warranted only if the <strong>in</strong>put channels (magnetic fields)<br />

are noise free, the output channel noise has a Gaussian distribution and the MT time<br />

series is stationary (Banks [1998]). In reality most <strong>data</strong> usually show gross departures<br />

from the above idealistic model. The ma<strong>in</strong> causes are geomagnetic phenomena, thunder-


vi<br />

storms, cultural <strong>in</strong>terference and <strong>in</strong>strument problems. This results <strong>in</strong> highly oscillat<strong>in</strong>g<br />

and biased estimates of MT transfer functions. ‘Remote reference’ technique (Gamble<br />

et al. [1979]) deals with the noise <strong>in</strong> magnetic fields. Reference magnetic field is recorded<br />

at a site, which is far outside the coherency range of the noise. Us<strong>in</strong>g the cross spectrum<br />

with the remote site <strong>in</strong>stead of the autopower elim<strong>in</strong>ates the bias <strong>in</strong> the magnetic field<br />

power. Problems due to non-stationarity of time series are addressed by subdivid<strong>in</strong>g the<br />

time series (Egbert and Booker [1986], Banks [1998]) <strong>in</strong>to small segments, estimat<strong>in</strong>g<br />

transfer functions for each one and averag<strong>in</strong>g <strong>in</strong> a way that discrim<strong>in</strong>ates aga<strong>in</strong>st noisy<br />

<strong>data</strong> segments. Robust estimation of MT transfer functions (Egbert and Booker [1986],<br />

Chave et al. [1987], Chave and Thomson [1989]) down weights the <strong>data</strong> sections with<br />

such large non Gaussian noise. But this technique still gives erroneous results, if strongly<br />

correlated noise is present dur<strong>in</strong>g most of the record<strong>in</strong>g time. To deal with such coherent<br />

noise, ‘robust multivariate errors-<strong>in</strong>-variables’ (RMEV) estimate was developed by Egbert<br />

[1997]. Correlated and uncorrelated noise are separated iteratively, <strong>us<strong>in</strong>g</strong> <strong>data</strong> from<br />

multiple stations. Variants of this technique viz. Signal-Noise Separation (SNS) method<br />

and SNS-remote-reference method are discussed by Larsen et al. [1996] and Oett<strong>in</strong>ger<br />

et al. [2001]. In a very recent work, Chave and Thomson [2003] proposed a bounded<br />

<strong>in</strong>fluence function to robustly estimate magnetotelluric <strong>data</strong> contam<strong>in</strong>ated with extreme<br />

noises. Over the past decade, conventional robust methods have revolutionized the application<br />

of magnetotellurics <strong>in</strong> geophysics (Jones et al. [1989], Egbert [1997]) and are now<br />

applied rout<strong>in</strong>ely and automatically produc<strong>in</strong>g reliable magnetotelluric responses <strong>in</strong> most<br />

<strong>in</strong>stances. The success of robust procedures may be attributed to three factors. First,<br />

its superiority to other <strong>data</strong> process<strong>in</strong>g techniques is established (Jones et al. [1989]).<br />

Second, these procedures can be justified rigorously (Egbert and Livelybrooks [1996])<br />

and third it can be easily implemented <strong>us<strong>in</strong>g</strong> iterative-weighted LS procedures and extended<br />

to remote reference process<strong>in</strong>g (Chave and Thomson [1989]). However at sites<br />

located near auroral region (Garcia et al. [1997]), major cultural noise centres, electric<br />

railway l<strong>in</strong>es etc where source field non-stationarity /noise contam<strong>in</strong>ation is severe, robust<br />

methods frequently break down. The possible reasons may be attributed to:-<br />

1. Failure <strong>in</strong> identify<strong>in</strong>g noise source for a survey often prompts the process<strong>in</strong>g of all<br />

sites <strong>in</strong> a similar fashion. Whereas, stations that were affected with noise should<br />

undergo special <strong>data</strong> treatment.<br />

2. Any MT time series process<strong>in</strong>g algorithm performs better, if presented with a <strong>data</strong><br />

set that is as clean as possible. This is often done by manual <strong>in</strong>spection of time<br />

series. A failure/omission at this stage may contribute to break down of robust<br />

process<strong>in</strong>g.<br />

3. Conventional robust process<strong>in</strong>g uses an <strong>in</strong>itial estimate of MT transfer functions<br />

usually derived from an LS estimator. When majority of observations deviate from<br />

the true one, the <strong>in</strong>itial LS estimate becomes too far (biased) for the robust iterative<br />

process to improve upon.<br />

4. The cross and auto spectra between electric and magnetic field elements are usually<br />

smoothed by a comb<strong>in</strong>ation of band and section averag<strong>in</strong>g. While effect of outliers<br />

<strong>in</strong> section averag<strong>in</strong>g is well known, the same about band averag<strong>in</strong>g is overlooked.


vii<br />

The present thesis concentrates on these problems of MT time series <strong>analysis</strong>, with<br />

an aim to improve the estimation of transfer functions. The wide band magnetotelluric<br />

<strong>data</strong> collected over Southern Granulite Terra<strong>in</strong>, offers scope for development of applications<br />

that may resist extreme noise contam<strong>in</strong>ation. There are two factors that make this<br />

region important <strong>in</strong> this context. 1) The majority of upper crustal rocks <strong>in</strong> SGT belong<br />

to Archaean and Proterozoic age (Naqvi and Rogers [1987]) and exhibit high electrical<br />

resistivity as other shield regions <strong>in</strong> the world (Mareschal et al. [1994]). Highly resistive<br />

upper crust offers very little attenuation to EM signals and <strong>in</strong> pr<strong>in</strong>ciple noise can<br />

propagate over larger distance <strong>in</strong> SGT as compared to regions where younger rocks are<br />

exposed. 2) The high population density <strong>in</strong> the southern states of India, especially <strong>in</strong><br />

Tamil Nadu, where most of the MT sites are located give rise to various cultural noises.<br />

The <strong>in</strong>dustrial belt along the two banks of Cauvery River is another noise source for MT.<br />

In this context, the objectives of the present thesis are,<br />

1. To characterize the signal/noise <strong>in</strong> magnetotelluric <strong>data</strong> collected over SGT, <strong>in</strong><br />

spatial, temporal and frequency doma<strong>in</strong> and to locate the major sources of noise <strong>in</strong><br />

the <strong>data</strong>.<br />

2. To evolve an efficient and automated method to discrim<strong>in</strong>ate noisy segments of time<br />

series. The enrichment of s/n ratio <strong>in</strong> the <strong>data</strong> provided by such a process will help<br />

the robust process<strong>in</strong>g to better estimate MT transfer functions.<br />

3. To improve the robust process<strong>in</strong>g methods of MT transfer functions particularly<br />

concentrat<strong>in</strong>g on the weak po<strong>in</strong>ts (reasons stated as (iii) and (iv) earlier) <strong>in</strong> the<br />

method, while deal<strong>in</strong>g with noisy <strong>data</strong>.<br />

4. Establish the efficacy of the process<strong>in</strong>g methods proposed by application to sufficiently<br />

large amount of <strong>data</strong> collected from SGT.<br />

A suite of advanced signal process<strong>in</strong>g algorithms will be used for achiev<strong>in</strong>g the above<br />

objectives. Typically MT time series is a large volume of multi-channel <strong>data</strong> (∼10 6 values)<br />

that are usually stored <strong>in</strong> compressed, b<strong>in</strong>ary format with a header file describ<strong>in</strong>g<br />

location, sensor geometry and filter sett<strong>in</strong>g, and the <strong>data</strong> itself. A variety of MT equipment<br />

is currently <strong>in</strong> use, which delivers MT <strong>data</strong> <strong>in</strong> wide bands of frequencies. Standards<br />

have been evolved <strong>in</strong> <strong>in</strong>dustry for the delivery and exchange of electromagnetic <strong>data</strong>, MT<br />

<strong>in</strong> particular, as Electrical Data Interchange (EDI) format by Society of Exploration Geophysicists<br />

(E. [1988]). Though it can support time series <strong>data</strong> as well, it is not be<strong>in</strong>g used<br />

so due to a various reasons. As time series is one-step closer to actual <strong>data</strong> generation<br />

at the <strong>in</strong>strument than the MT transfer functions, the variation <strong>in</strong> MT hardware also<br />

constra<strong>in</strong>ts the adoption of uni-format for MT time series. The major academic software<br />

for MT time series process<strong>in</strong>g, currently available free of charge are codes from Egbert<br />

[1997], RRMT by Chave A.D, LiMS by Jones A.G. (available at http://mtnet.<strong>in</strong>fo) and<br />

EMERALD by Ritter et al. [1998]. These codes are written specifically for certa<strong>in</strong> types<br />

of MT <strong>in</strong>struments and for use on particular operat<strong>in</strong>g systems. The commercial manufacturers<br />

of MT equipment also give process<strong>in</strong>g software as part of the system (for e.g.<br />

ProcMT r○ and MAPROS r○ from Metronix GmBH) which optimize user requirements,


viii<br />

though not as versatile as the academic software. Moreover these software strictly cater<br />

to the needs of one MT equipment, and mostly their source codes are not open. This<br />

scenario puts pressure on researchers work<strong>in</strong>g on MT time series process<strong>in</strong>g. In order to<br />

modify process<strong>in</strong>g rout<strong>in</strong>es, it is necessary to access the time series, calibrate the <strong>data</strong><br />

etc, which have been specifically evolved for particular equipment. Any new process<strong>in</strong>g<br />

codes, need to be attached to a time series read<strong>in</strong>g, calibrat<strong>in</strong>g and stor<strong>in</strong>g utility, which<br />

also has to be developed. Thus time series process<strong>in</strong>g algorithm should have end to end<br />

utilities, which <strong>in</strong>volves read<strong>in</strong>g compressed time series <strong>data</strong>, read<strong>in</strong>g the sensor geometry<br />

and filter sett<strong>in</strong>gs, calibrat<strong>in</strong>g for system responses, the ma<strong>in</strong> process<strong>in</strong>g and f<strong>in</strong>ally the<br />

export<strong>in</strong>g the results <strong>in</strong> EDI format. In one way this facilitates easy <strong>in</strong>terchange of <strong>data</strong><br />

and flexibility of applications, compared to a framework, where<strong>in</strong> one relies on a set of<br />

imported utilities to perform the peripheral tasks. For the current thesis, the process<strong>in</strong>g<br />

codes were written on end to end basis, where it performs all the peripheral process<strong>in</strong>g<br />

tasks as well.<br />

As remote reference process<strong>in</strong>g has been <strong>in</strong>creas<strong>in</strong>gly used by MT practitioners, <strong>in</strong><br />

tandem with a robust process<strong>in</strong>g algorithm, a question can be asked ’Why should we be<br />

concerned with s<strong>in</strong>gle station process<strong>in</strong>g at all?’. There are several good reasons. First,<br />

because of <strong>in</strong>strumental problem (as happened dur<strong>in</strong>g <strong>data</strong> acquisition <strong>in</strong> SGT), it is<br />

common to have s<strong>in</strong>gle station record<strong>in</strong>g <strong>in</strong> many surveys. If the ’reference’ site is very<br />

noisy and the ’local’ site is not, s<strong>in</strong>gle station estimation could be better than remote<br />

reference estimates (Egbert and Livelybrooks [1996]). These factors often necessitate<br />

s<strong>in</strong>gle station process<strong>in</strong>g of MT <strong>data</strong>. However the procedures developed <strong>in</strong> this thesis<br />

can easily be extended to remote reference process<strong>in</strong>g as well (an example <strong>in</strong> this regard is<br />

discussed <strong>in</strong> Chapter 6). The thesis spreads <strong>in</strong> seven chapters. In the follow<strong>in</strong>g sections,<br />

an overview of the different chapters <strong>in</strong> the thesis is given.<br />

Chapter I<br />

Maxwell’s equations describe the properties of electromagnetic waves. The relationship<br />

between electric and magnetic filed with<strong>in</strong> a conductive Earth can be expressed <strong>in</strong><br />

terms of wave equations by comb<strong>in</strong><strong>in</strong>g Maxwell’s equations. A conductive Earth responds<br />

to the electromagnetic illum<strong>in</strong>ation at the surface by allow<strong>in</strong>g the refracted components<br />

to diffuse accord<strong>in</strong>g to their frequency content. In consequence the magnetotelluric fields<br />

conta<strong>in</strong> <strong>in</strong>formation regard<strong>in</strong>g conductivity distribution as a function of frequency. Instrumentation<br />

and field procedures play an important part <strong>in</strong> the quality of measured<br />

magnetotelluric <strong>data</strong>. A discussion on field procedures, sensors and measur<strong>in</strong>g device<br />

employed to measure magnetotelluric <strong>data</strong> is given towards the end of Chapter 1.<br />

Chapter II<br />

The magnetotelluric transfer functions are obta<strong>in</strong>ed by the simultaneous measurement<br />

of Earth’s time vary<strong>in</strong>g electrical and magnetic field. The sources of the natural electromagnetic<br />

fields <strong>in</strong> the frequency range 10 −4 to 10 4 Hz have ’quasi-planar’ properties at<br />

the surface of the Earth. Natural signal sources for MT measurements <strong>in</strong> the frequency<br />

range 10 0 to 10 4 Hz are worldwide thunderstorm activity. At frequencies below 1 Hz the<br />

source signals are generated due to the <strong>in</strong>teraction of magnetosphere with solar w<strong>in</strong>d.<br />

Observed magnetotelluric fields at the surface of Earth conta<strong>in</strong> additive noises from various<br />

manmade and natural sources, which do not behave as plane waves. A discussion on<br />

different signal and noise sources and their effect <strong>in</strong> MT <strong>data</strong> is <strong>in</strong>cluded <strong>in</strong> Chapter II.


ix<br />

Chapter III<br />

This chapter <strong>in</strong>troduces the concepts of random <strong>data</strong> <strong>analysis</strong>. <strong>Magnetotelluric</strong> time<br />

series like many other natural processes can be considered as wide band random <strong>data</strong>.<br />

And thus the estimation of its properties becomes statistical (Bendat and Piersol [1971]).<br />

A brief review of some of its properties viz. probability distribution functions, autocorrelation<br />

and power spectral density function are discussed. As magnetotelluric signals<br />

constitute a multivariate system, the jo<strong>in</strong>t properties of <strong>in</strong>dividual components like cross<br />

correlation, cross spectra and coherencies are also important. Reliable estimation of<br />

magnetotelluric transfer function depends on the amount of noise <strong>in</strong> the time series measurements<br />

(Orange, 1989). It is often difficult to obta<strong>in</strong> noise free measurements as the<br />

method relies on highly variable natural electromagnetic variation. Severe problems are<br />

caused by civilization, which produce all k<strong>in</strong>ds of electromagnetic noises. These noises<br />

manifest themselves <strong>in</strong> the computed magnetotelluric transfer function as both statistical<br />

and bias shifts. Sims et al. [1971] proposed the classical least square solution for MT<br />

transfer functions from noisy measured <strong>data</strong>. The statistical errors can be m<strong>in</strong>imized by<br />

averag<strong>in</strong>g large amount of observations, if the noise distribution is Gaussian. As least<br />

square solutions envisages a noise free <strong>in</strong>put field, noise <strong>in</strong> <strong>in</strong>put field can severely bias<br />

such estimates. The bias errors can be tackled to an extent by measur<strong>in</strong>g a remote<br />

reference (Gamble et al. [1979]). But as shown by Chave et al. [1987], Chave and Thomson<br />

[1989], Egbert and Booker [1986], violation of the Gaussian assumption for errors<br />

may lead to severe problems <strong>in</strong> least square estimation of MT transfer functions. The<br />

properties of a least square estimator is discussed towards the end of Chapter III.<br />

Chapter IV<br />

The <strong>data</strong> used <strong>in</strong> the present study were measured over South India, as a part of an<br />

<strong>in</strong>tegrated geological and geophysical study on the Southern Granulite Terra<strong>in</strong> by NGRI<br />

under a DST project. The magnetotelluric measurements were carried out <strong>in</strong> two field<br />

campaigns from 1998 to 2000. The sites were located <strong>in</strong> a 300-km long NS corridor. The<br />

profile traverses major metamorphic and tectonic elements of the region. Har<strong>in</strong>arayana<br />

et al. [2003] discusses the result of model<strong>in</strong>g of the <strong>data</strong> <strong>in</strong> terms of subsurface electrical<br />

conductivity distributions and their importance <strong>in</strong> local tectonic set up. Five components<br />

of natural EM variations were measured <strong>in</strong> the frequency range 0.0001 - 4000 s. The <strong>data</strong><br />

were acquired <strong>in</strong> four frequency bands. The spatial distribution and quantity of <strong>data</strong><br />

collected is described <strong>in</strong> detail <strong>in</strong> Chapter IV. Most of the <strong>data</strong> were collected <strong>in</strong> s<strong>in</strong>gle<br />

station reference, as the measurement units had time synchronization problems. This<br />

necessitated use of s<strong>in</strong>gle station process<strong>in</strong>g techniques (Chapter III) for majority of MT<br />

sites measured. The measurement corridor passes through many urban areas and one<br />

<strong>in</strong>dustrial belt. Major electrified rail l<strong>in</strong>es passes through some parts of the corridor. More<br />

over the southern region of India is densely populated. The cultural noises thus generated<br />

can propagate over large distances, as the upper crust is highly resistive. The <strong>in</strong>dices<br />

of geomagnetic activity dur<strong>in</strong>g the measurement time were compared with the averaged<br />

long period coherence of each site. The agreement of the coherence and the geomagnetic<br />

<strong>in</strong>dices <strong>in</strong>dicate the validity of such an approach. However a few disagreements were<br />

also evident. This <strong>in</strong>dicates the possible noise processes contribut<strong>in</strong>g to the measured<br />

<strong>data</strong>. Spatial relation of the known cultural noise centers and the quality of MT <strong>data</strong><br />

was exam<strong>in</strong>ed. The strong correlation of the high noise/(signal +noise) ratio to the


x<br />

major <strong>in</strong>dustrial belt <strong>in</strong>dicate that the MT signals may get consistently degraded due<br />

to presence of an active noise sources, irrespective of the signal activity. However the<br />

conclusions drawn from this chapter should be treated rather cautiously. The signal<br />

and noise are def<strong>in</strong>ed <strong>in</strong> this chapter as <strong>in</strong>l<strong>in</strong>e and outl<strong>in</strong>e components of a least square<br />

solution of MT transfer functions.<br />

Chapter V<br />

The techniques outl<strong>in</strong>ed <strong>in</strong> Chapter III for estimation of magnetotelluric transfer functions<br />

from measured imprecise <strong>data</strong> requires that the time series is presented as clean<br />

as possible. Clean<strong>in</strong>g of MT time series is presently done by manual <strong>in</strong>spection (edit<strong>in</strong>g).<br />

Edit<strong>in</strong>g of magnetotelluric time series is subjective <strong>in</strong> nature and time consum<strong>in</strong>g.<br />

Artificial neural networks (ANN) are widely used to automate processes, which requires<br />

human <strong>in</strong>telligence. In Chapter V Artificial Neural Network is used to discrim<strong>in</strong>ate good<br />

sections of <strong>data</strong> aga<strong>in</strong>st noisy ones. Artificial neural networks (ANN) are emerg<strong>in</strong>g tools<br />

that have been applied <strong>in</strong> many areas of science and eng<strong>in</strong>eer<strong>in</strong>g where pattern recognition<br />

is <strong>in</strong>volved, such as speech and character recognition. The learn<strong>in</strong>g and adaptive<br />

capabilities of these models make them attractive for application to some problems <strong>in</strong><br />

geophysics. As ANN based techniques are computationally <strong>in</strong>tensive, a novel approach<br />

was made to the problem, which <strong>in</strong>volves edit<strong>in</strong>g of five simultaneously measured MT<br />

time series. Neural network tra<strong>in</strong><strong>in</strong>g was done at two levels. Signal and noise patterns<br />

of <strong>in</strong>dividual channels were taught first. Tra<strong>in</strong><strong>in</strong>g was stopped when both errors were<br />

below acceptable level. The neural network’s sensitivity to signal to noise ratio and the<br />

relative significance of it’s <strong>in</strong>puts were tested to ensure the tra<strong>in</strong><strong>in</strong>g was correct. The application<br />

of ANN based edit<strong>in</strong>g to magnetotelluric time series br<strong>in</strong>gs out some <strong>in</strong>terest<strong>in</strong>g<br />

results. In a low noise environment the network edit<strong>in</strong>g produces results almost similar<br />

to bl<strong>in</strong>d edit<strong>in</strong>g (<strong>us<strong>in</strong>g</strong> all stacks). On such a <strong>data</strong>, a simple coherency-based estimator<br />

can do the signal discrim<strong>in</strong>ation to a certa<strong>in</strong> level of satisfaction. However the neural<br />

network’s ability to pick out signal from moderate to high noise environment was evident<br />

on the other <strong>data</strong> collected from SGT. In such cases it approximates human <strong>in</strong>telligence<br />

- established from the fact that the neural network based edit<strong>in</strong>g gives a result similar<br />

to manual edit<strong>in</strong>g. These results satisfy the second objective of the thesis, i.e., to provide<br />

a robust alternative to manual edit<strong>in</strong>g of magnetotelluric time series. The signal<br />

and noise characteristics <strong>in</strong> magnetotelluric <strong>data</strong> are very different <strong>in</strong> different frequency<br />

ranges. This is due to the difference <strong>in</strong> signal and noise source mechanism <strong>in</strong> different<br />

part of Earth’s natural electromagnetic spectrum. Sensor geometry and <strong>in</strong>strumentation<br />

also affect the pattern of signal on which the neural network is tra<strong>in</strong>ed. This necessitates<br />

reformulation of MT signal and noise characteristics for tra<strong>in</strong><strong>in</strong>g of ANN, wherever<br />

necessary. However the extra computational requirement of re-tra<strong>in</strong><strong>in</strong>g of ANN may not<br />

pose a burden on resources, tak<strong>in</strong>g <strong>in</strong>to consideration the ever-<strong>in</strong>creas<strong>in</strong>g computational<br />

power of microprocessors.<br />

Chapter VI<br />

In chapter VI, two new approaches are proposed to improve the performance of robust<br />

statistical procedures on MT time series. Non-parametric estimators such as Jackknife<br />

(Efron [1982]) were used to robustly compute the variance of MT transfer functions<br />

(Chave and Thomson [1989]). Its use as an effective <strong>in</strong>itial guess for robust procedures is<br />

discussed. It is shown that <strong>in</strong> majority of the cases, the use of Jackknife for <strong>in</strong>itial guess


xi<br />

resulted <strong>in</strong> better estimation of MT transfer function as compared to LS estimations. It<br />

is usual <strong>in</strong> MT to sub divide the time series, estimate the spectral density matrices for<br />

each segment <strong>in</strong>dividually and then robustly average the spectra or transfer functions<br />

between the sub segments (section averag<strong>in</strong>g). With<strong>in</strong> a segment, it is common to use<br />

a limited number of target frequencies, and obta<strong>in</strong> smooth spectra by averag<strong>in</strong>g several<br />

adjacent Fourier harmonics (frequency band averag<strong>in</strong>g). The documented researches on<br />

robust estimation of MT spectral densities and transfer functions concentrate on section<br />

averag<strong>in</strong>g. This arises from the assumption that, with<strong>in</strong> a narrow frequency band, the<br />

distribution of Fourier coefficients are of Gaussian nature and a simple average (LS)<br />

gives the best estimate. It is shown that this argument often fails, and the problem of<br />

contam<strong>in</strong>ation is applicable to band averag<strong>in</strong>g as well. Robust weight<strong>in</strong>g approach is<br />

proposed for estimation of cross and auto spectral estimation with<strong>in</strong> a band, without<br />

mak<strong>in</strong>g specific model assumptions concern<strong>in</strong>g signal or noise. Both these proposed<br />

procedures, while applied on a large volume of MT <strong>data</strong> collected over SGT, South India,<br />

met with moderate to good improvement of MT transfer functions.<br />

Chapter VII<br />

Chapter VII gives an overall summary of the results discussed <strong>in</strong> chapter IV to VI.<br />

The ma<strong>in</strong> objective of the thesis, i.e. to obta<strong>in</strong> best estimates of transfer function from<br />

measured magnetotelluric time series, may require a comb<strong>in</strong>ation of one or more techniques<br />

<strong>in</strong>troduced and demonstrated <strong>in</strong> the chapters IV to VI. Furthermore, Chapter VII<br />

will take a closer look at the properties of robust process<strong>in</strong>g and neural networks. On<br />

comparison of <strong>in</strong>dividual performances of neural networks and robust process<strong>in</strong>g, it was<br />

found that their reaction to different types of noise differs <strong>in</strong> some cases. One reason is<br />

that the outliers / or coherent noises which are obvious <strong>in</strong> time doma<strong>in</strong>, failed to produce<br />

larger outliers and thus were not down weighted by robust procedures. In another<br />

<strong>in</strong>stance, <strong>data</strong> sets with noises, which the neural network allowed to pass on, were down<br />

weighted by robust procedures. This clearly shows the need to comb<strong>in</strong>e the two techniques<br />

<strong>in</strong> order to discrim<strong>in</strong>ate / down weight a majority of noisy <strong>data</strong>. Further more,<br />

this po<strong>in</strong>ts to the necessity of transform<strong>in</strong>g the <strong>data</strong> more than one doma<strong>in</strong> to discrim<strong>in</strong>ate<br />

noise from signal. In this regard, possible use of wavelet transform <strong>in</strong> identify<strong>in</strong>g<br />

transient signals that are the common form of MT signals <strong>in</strong> for frequencies > 1Hz is<br />

suggested as a future work.


Contents<br />

1 <strong>Magnetotelluric</strong>s: Basic Theory, Sensors and Field Procedures 1<br />

1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2<br />

1.1.1 Source fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2<br />

1.1.2 Maxwell’s equations . . . . . . . . . . . . . . . . . . . . . . . . . 3<br />

1.1.3 Sk<strong>in</strong> depth. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4<br />

1.1.4 Impedance tensor: . . . . . . . . . . . . . . . . . . . . . . . . . . 5<br />

1.1.5 Amplitude and phase of impedance: . . . . . . . . . . . . . . . . . 8<br />

1.2 Sensors and Field Procedures . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

1.2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

1.2.2 Electric field sensors . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

1.2.3 Magnetic Sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

1.2.4 Record<strong>in</strong>g systems . . . . . . . . . . . . . . . . . . . . . . . . . . 12<br />

1.3 Field Procedures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

1.3.1 Survey design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

1.3.2 Site selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

1.3.3 Sensor deployment . . . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

1.3.4 Calibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

1.3.5 Data acquisition. . . . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

2 Signal and Noise Sources for magnetotellurics 18<br />

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

2.2 Signal sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

2.2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

2.2.2 Geomagnetic pulsations . . . . . . . . . . . . . . . . . . . . . . . 21<br />

2.2.3 Thunder storm activity . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

2.3 Noise Sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

2.3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

2.3.2 Noises from power l<strong>in</strong>e signals . . . . . . . . . . . . . . . . . . . . 25<br />

2.3.3 Noise from electric traction . . . . . . . . . . . . . . . . . . . . . 26<br />

2.3.4 Noises from <strong>in</strong>strument & sensors . . . . . . . . . . . . . . . . . . 27<br />

2.3.5 Noises from the other sources . . . . . . . . . . . . . . . . . . . . 28<br />

2.4 Effect of active electrical noise <strong>in</strong> magnetotelluric <strong>data</strong> . . . . . . . . . . 29<br />

xii


xiii<br />

3 <strong>Magnetotelluric</strong> time series <strong>analysis</strong> 31<br />

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32<br />

3.2 Random <strong>data</strong> <strong>analysis</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

3.2.1 Properties of random <strong>data</strong> . . . . . . . . . . . . . . . . . . . . . . 33<br />

3.2.1.1 Probability density function . . . . . . . . . . . . . . . . 33<br />

3.2.1.2 Mean square and variance . . . . . . . . . . . . . . . . . 34<br />

3.2.1.3 Median and average absolute deviation. . . . . . . . . . 34<br />

3.2.1.4 Autocorrelation function . . . . . . . . . . . . . . . . . . 35<br />

3.2.1.5 Power spectral density function . . . . . . . . . . . . . . 35<br />

3.2.2 Jo<strong>in</strong>t signal properties . . . . . . . . . . . . . . . . . . . . . . . . 36<br />

3.2.2.1 Cross correlation functions . . . . . . . . . . . . . . . . . 36<br />

3.2.2.2 Cross-spectral density functions . . . . . . . . . . . . . . 37<br />

3.2.2.3 Coherence . . . . . . . . . . . . . . . . . . . . . . . . . . 38<br />

3.2.3 Computational aspects . . . . . . . . . . . . . . . . . . . . . . . . 38<br />

3.2.3.1 Trend and bias removal . . . . . . . . . . . . . . . . . . 38<br />

3.2.3.2 Power Spectral Density function . . . . . . . . . . . . . 38<br />

3.2.3.3 W<strong>in</strong>dow<strong>in</strong>g . . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

3.2.3.4 Discrete Fourier Transform . . . . . . . . . . . . . . . . 39<br />

3.2.3.5 Smooth<strong>in</strong>g of spectra by band and section averag<strong>in</strong>g . . 40<br />

3.3 <strong>Magnetotelluric</strong> transfer functions . . . . . . . . . . . . . . . . . . . . . . 41<br />

3.3.1 Least Square Solution . . . . . . . . . . . . . . . . . . . . . . . . 41<br />

3.3.1.1 Multi <strong>in</strong>put, multi output l<strong>in</strong>ear system . . . . . . . . . 41<br />

3.3.1.2 Solution with noise free <strong>data</strong> . . . . . . . . . . . . . . . 42<br />

3.3.1.3 Solution with noise <strong>in</strong> measurement . . . . . . . . . . . . 43<br />

3.3.2 Concept of bias . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44<br />

3.3.2.1 Predicted coherence . . . . . . . . . . . . . . . . . . . . 45<br />

3.3.3 Coherent and <strong>in</strong>coherent noises . . . . . . . . . . . . . . . . . . . 45<br />

3.3.4 Variance & Errors . . . . . . . . . . . . . . . . . . . . . . . . . . . 46<br />

3.3.5 Coherence and bias of transfer functions . . . . . . . . . . . . . . 46<br />

3.3.6 Remote reference . . . . . . . . . . . . . . . . . . . . . . . . . . . 48<br />

4 Signal and Noise Characteristics of MT <strong>data</strong> measured over the Southern<br />

Granulite Terra<strong>in</strong> 49<br />

4.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50<br />

4.2 Geological objectives of MT <strong>in</strong>vestigations <strong>in</strong> the SGT . . . . . . . . . . 50<br />

4.3 The Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52<br />

4.3.1 Acquisition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52<br />

4.3.2 Typical MT Time series . . . . . . . . . . . . . . . . . . . . . . . 53<br />

4.3.3 Examples of MT <strong>data</strong> collected over South India . . . . . . . . . . 55<br />

4.4 The EEJ effect on MT <strong>data</strong> . . . . . . . . . . . . . . . . . . . . . . . . . 59<br />

4.5 Signal Activity dur<strong>in</strong>g the Field Campaign . . . . . . . . . . . . . . . . . 61<br />

4.6 Spatial character of coherence . . . . . . . . . . . . . . . . . . . . . . . . 62<br />

4.7 Geographic relation of noise . . . . . . . . . . . . . . . . . . . . . . . . . 64<br />

4.8 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64


xiv<br />

5 The application of the artificial neural networks to magnetotelluric time<br />

series <strong>analysis</strong> 67<br />

5.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68<br />

5.2 Signal and noise <strong>in</strong> the magnetotelluric time series . . . . . . . . . . . . 68<br />

5.3 Visual Inspection (edit<strong>in</strong>g) of magnetotelluric time series <strong>data</strong>; why automation<br />

? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68<br />

5.4 <strong>Magnetotelluric</strong> noise characterization . . . . . . . . . . . . . . . . . . . . 70<br />

5.4.1 Patterns of signal & noise : . . . . . . . . . . . . . . . . . . . . . 70<br />

5.4.2 Amplitude of signals . . . . . . . . . . . . . . . . . . . . . . . . . 71<br />

5.4.3 Correlation between simultaneously measured channels . . . . . . 71<br />

5.5 Artificial Neural Networks . . . . . . . . . . . . . . . . . . . . . . . . . . 71<br />

5.5.1 Why artificial neural network? . . . . . . . . . . . . . . . . . . . . 71<br />

5.5.2 ANN theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

5.6 Data <strong>analysis</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74<br />

5.6.1 Network eng<strong>in</strong>eer<strong>in</strong>g . . . . . . . . . . . . . . . . . . . . . . . . . 74<br />

5.6.2 Pattern Tra<strong>in</strong><strong>in</strong>g . . . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />

5.6.2.1 Data used: . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />

5.6.2.2 Pre-process<strong>in</strong>g . . . . . . . . . . . . . . . . . . . . . . . 75<br />

5.6.2.3 FANN Tra<strong>in</strong><strong>in</strong>g . . . . . . . . . . . . . . . . . . . . . . . 76<br />

5.6.2.4 Sensitivity <strong>analysis</strong> . . . . . . . . . . . . . . . . . . . . . 77<br />

5.6.3 Inter channel tra<strong>in</strong><strong>in</strong>g . . . . . . . . . . . . . . . . . . . . . . . . 78<br />

5.6.3.1 The <strong>data</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . 78<br />

5.6.3.2 FANN tra<strong>in</strong><strong>in</strong>g . . . . . . . . . . . . . . . . . . . . . . . 80<br />

5.6.3.3 Relative significance of <strong>in</strong>put . . . . . . . . . . . . . . . 81<br />

5.7 Application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82<br />

5.8 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86<br />

6 Estimation of <strong>Magnetotelluric</strong> Transfer Functions: Robust Statistical<br />

Methods 88<br />

6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89<br />

6.2 Robust estimation of MT transfer functions . . . . . . . . . . . . . . . . 90<br />

6.2.1 Why robust methods? . . . . . . . . . . . . . . . . . . . . . . . . 90<br />

6.2.2 Robust M estimators . . . . . . . . . . . . . . . . . . . . . . . . . 91<br />

6.2.3 Choice of <strong>in</strong>fluence functions . . . . . . . . . . . . . . . . . . . . . 93<br />

6.2.4 Implementation for MT . . . . . . . . . . . . . . . . . . . . . . . 95<br />

6.2.4.1 Initial guess of transfer function . . . . . . . . . . . . . . 96<br />

6.2.4.2 Jackknife estimate as <strong>in</strong>itial guess . . . . . . . . . . . . . 96<br />

6.2.4.3 Scale estimate . . . . . . . . . . . . . . . . . . . . . . . . 100<br />

6.2.4.4 Robust transfer function estimation . . . . . . . . . . . . 100<br />

6.2.4.5 Tukey weights . . . . . . . . . . . . . . . . . . . . . . . . 101<br />

6.2.4.6 Comput<strong>in</strong>g the variance . . . . . . . . . . . . . . . . . . 101<br />

6.2.4.7 Quantile Quantile plots . . . . . . . . . . . . . . . . . . 101<br />

6.2.5 Application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102<br />

6.2.5.1 Flowchart . . . . . . . . . . . . . . . . . . . . . . . . . . 102<br />

6.2.5.2 Comparison of robust process<strong>in</strong>g schemes . . . . . . . . 104


xv<br />

6.3 Robust Band averag<strong>in</strong>g . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105<br />

6.3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105<br />

6.3.2 Effect of frequency band width . . . . . . . . . . . . . . . . . . . 109<br />

6.3.3 Robust estimation of spectra . . . . . . . . . . . . . . . . . . . . . 111<br />

6.3.4 Flow chart . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114<br />

6.3.5 Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114<br />

6.4 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116<br />

6.4.1 VP14 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116<br />

6.4.2 TT08 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116<br />

6.4.3 TT04 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117<br />

6.4.4 OK18 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117<br />

6.4.5 JN12 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117<br />

6.4.6 VP16 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118<br />

6.4.7 OK16 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118<br />

6.4.8 VP12 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118<br />

6.4.9 Comparison of results from the Vellar - Palani profile <strong>in</strong> SGT . . . 118<br />

6.4.10 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119<br />

7 Discussion and conclusions 131<br />

7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132<br />

7.2 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132<br />

7.3 Neural Network and Robust process<strong>in</strong>g . . . . . . . . . . . . . . . . . . . 133<br />

7.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136<br />

7.5 Suggestions for future work . . . . . . . . . . . . . . . . . . . . . . . . . 137<br />

Bibliography


List of Figures<br />

1.1 Quasi planar nature of electromagnteic wave front at great distance from<br />

source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3<br />

1.2 Behaviour of plane EM wave at the surface of a conduct<strong>in</strong>g Earth . . . . 3<br />

1.3 Diagrammatic representation of 1D, 2D and 3D situations. . . . . . . . 6<br />

1.4 Planar view. Off diagonal tensor elements become unequal when geology<br />

has a preferred direction. On rotat<strong>in</strong>g the tensor to the strike, diagonal<br />

elements get m<strong>in</strong>imized. See text for discussion . . . . . . . . . . . . . . 7<br />

1.5 Implant<strong>in</strong>g of Electrical Field sensor . . . . . . . . . . . . . . . . . . . . 10<br />

1.6 Installation of <strong>in</strong>duction coil magnetometer . . . . . . . . . . . . . . . . 11<br />

1.7 Simplified equivalent circuit for <strong>in</strong>duction coil magnetometer . . . . . . . 11<br />

1.8 Theoretical response function for MFS05 magnetometer (from Pulz and<br />

Ritter [2001]) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12<br />

1.9 Block diagram depict<strong>in</strong>g different signal process<strong>in</strong>g steps <strong>in</strong> GMS05<br />

(adapted from Metronix [1997]) . . . . . . . . . . . . . . . . . . . . . . . 14<br />

1.10 GPU05 (ma<strong>in</strong> unit) and HDU05 (display unit) dur<strong>in</strong>g <strong>data</strong> acquisition . 14<br />

1.11 Typical lay out for MT sensors and <strong>data</strong> acquisition (after Metronix [1997]) 16<br />

2.1 Earths horizontal magnetic field spectrum (Modified after Macnae et al.<br />

[1984]) The Black and gray bars <strong>in</strong>dicate the frequency range of measurement<br />

bands for GMS05. (see 1.1) . . . . . . . . . . . . . . . . . . . . . . 20<br />

2.2 Plot for MT time series recorded at site VP16 show<strong>in</strong>g geomagnetic pulsations<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22<br />

2.3 Averaged magnetic amplitude spectra (Hx) at station VP16. Peaks related<br />

to geomagnetic pulsations are labeled.Thick (gray) l<strong>in</strong>e represents<br />

smoothed spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22<br />

2.4 Plot of MT time series show<strong>in</strong>g impulse signals related to sferics recorded<br />

at site C12. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

2.5 Electric Field (Ex) spectral amplitude depict<strong>in</strong>g the Schumann resonance 24<br />

2.6 Distribution of electromagnetic field strength of three phase power transmission<br />

l<strong>in</strong>e <strong>in</strong> the immediate vic<strong>in</strong>ity of power l<strong>in</strong>es (adopted from Szarka<br />

[1988]). H total magnetic field, E horizontal electric field. . . . . . . . . 26<br />

2.7 Amplitude spectra of electric field computed over different <strong>data</strong> length for<br />

the site OK14. Th<strong>in</strong> black l<strong>in</strong>e - 1024, Thick gray l<strong>in</strong>e 256 and Gray dots<br />

- 128. See text for discussion . . . . . . . . . . . . . . . . . . . . . . . . 27<br />

xvi


xvii<br />

2.8 Current circuit electric railway systems. I is the current <strong>in</strong> the overhead<br />

power l<strong>in</strong>e I1 is the return current <strong>in</strong> the rails and I2 is the return current<br />

<strong>in</strong> the Earth ( modified after Chaize and Lavergne [1970]). . . . . . . . . 28<br />

2.9 Near and far field effect of a grounded dipole on MT measurements<br />

(adapted from Zonge and Hughes [1991]). The two panels on left and<br />

right describes the apparent resistivity and phase values for far field and<br />

near field measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

2.10 Effect of a near field source on MT <strong>data</strong>, as recorded at the station OK14.<br />

See text for discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30<br />

3.1 Ordered —Zxy— plotted aga<strong>in</strong>st number of occurrences observed at station<br />

C13 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

3.2 Crosscorrelogram of two magnetotelluric time series components. . . . . . 37<br />

3.3 The Parzen spectral w<strong>in</strong>dow over the target frequency plotted on background<br />

of a sample spectra . . . . . . . . . . . . . . . . . . . . . . . . . 40<br />

3.4 <strong>Magnetotelluric</strong> l<strong>in</strong>ear system with two horizontal magnetic components<br />

as <strong>in</strong>puts and horizontal electric components as outputs. Adapted from<br />

Jones et al. [1989]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42<br />

3.5 MT Transfer function (Z xy ) computed <strong>us<strong>in</strong>g</strong> upward and downward biased<br />

estimators compared with the coherence functions for station VP20. The<br />

upper part of the diagram shows real and imag<strong>in</strong>ary components of upward<br />

and downward biased estimates. The variance of Z xy is shown as solid l<strong>in</strong>e.<br />

The lower part shows the predicted (multiple) and ord<strong>in</strong>ary coherence<br />

functions. See the text for discussion. . . . . . . . . . . . . . . . . . . . . 47<br />

4.1 MT stations superimposed over the geology of the measurement corridor<br />

<strong>in</strong> South India (Geology adapted from GSI [1995]). See text for discussion 51<br />

4.2 Typical MT time series measured <strong>in</strong> the SGT for two frequency ranges a)<br />

the frequency range 256 Hz to 32kHz (band 1) and b) the period range 8<br />

Hz to 256 Hz (band2) .See the text for discussion. . . . . . . . . . . . . . 54<br />

4.3 Typical MT time series measured <strong>in</strong> the SGT for two frequency ranges c)<br />

the frequency range 8Hz to 0.25 Hz (band 3) and d) the period range 4<br />

sec to 128 sec.See the text for discussion. . . . . . . . . . . . . . . . . . 56<br />

4.4 Plot of apparent resistivity, phase, predicted coherence and degree of freedom<br />

(DOF) vs frequency for four stations measured over South India.<br />

Data are computed <strong>in</strong> their measured direction. See text for discussion . 58<br />

4.5 Apparent resistivity vs frequency plot for 3 stations tht are near and away<br />

from the dip equator with day and night curves superimposed (Rao et al.<br />

[2002]). (a) OK3 - site 300 km north of dip equator, (b) IRA site 100 km<br />

north of dip equator and (3) KAR - site near the dip equator. . . . . . . 60<br />

4.6 Averaged Kp <strong>in</strong>dices [Courtesy NGDC], K <strong>in</strong>dices [HYB] dur<strong>in</strong>g the MT<br />

measurements are compared with telluric predicted coherence [4 sec 128<br />

sec]. The mean coherence is drawn as a l<strong>in</strong>e. . . . . . . . . . . . . . . . 61


xviii<br />

4.7 The band averaged telluric predicted coherence plotted as a contoured map<br />

for all the measured MT sites. (a) Band1 (b) Band2 (c) Band 3 and (d)<br />

Band4. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63<br />

4.8 Contoured map of telluric Noise/(Signal+Noise) ratio, superimposed on<br />

the major geographical elements of the region. Note N/(S+N) ratio on<br />

either sides of Cauvery river near Sankari. . . . . . . . . . . . . . . . . . 65<br />

5.1 Common signal and noise patterns <strong>in</strong> long period MT time series. Samples<br />

are collected from different sites. Note the change <strong>in</strong> amplitudes. (a) Signal<br />

patterns; Samples of E x and H y shows the geomagnetic pulsations. Other<br />

channels also show signals but at longer periods. (b) Noise patterns; E y &<br />

H x shows different types of spikes. A step and its decay is shown <strong>in</strong> H y .<br />

Sample of random noise is shown <strong>in</strong> H z . . . . . . . . . . . . . . . . . . . 69<br />

5.2 Simple three layer feed forward neural network. The <strong>data</strong> is processed at<br />

each neuron <strong>in</strong> the layers. Each neuron performs a summ<strong>in</strong>g of <strong>in</strong>puts multiplied<br />

with a weight parameter and outputs the <strong>data</strong> through its sigmoid<br />

transfer function. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

5.3 Data flow through the feed forward artificial neural network (FANN) based<br />

edit<strong>in</strong>g scheme. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />

5.4 Stacked amplitude and phase spectra of the tra<strong>in</strong><strong>in</strong>g <strong>data</strong>base. The spectra<br />

of noisy <strong>data</strong> (class 0.1) clearly different from the signals (class 0.9). . . 76<br />

5.5 Results from the pattern tra<strong>in</strong><strong>in</strong>g. (a) The SSE as a function of epoch.<br />

The error reached the m<strong>in</strong>imum floor after 400 epochs. Stability of convergence<br />

is demonstrated up to 1000 epochs. (b) Deviation between manually<br />

classified and network predicted classes for 500 test time series segments.<br />

The scattered po<strong>in</strong>ts shows the major deviations. The correct pick<strong>in</strong>g<br />

constitute 94 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77<br />

5.6 Network output versus signal content. The network was simulated by<br />

<strong>in</strong>puts with vary<strong>in</strong>g signal content. A narrow region of high variance exists<br />

when signal content is between 65 . . . . . . . . . . . . . . . . . . . . . . 78<br />

5.7 Pattern, amplitude ratios and correlation coefficients of 900 stacks which<br />

form the <strong>data</strong>base for <strong>in</strong>ter channel tra<strong>in</strong><strong>in</strong>g and test<strong>in</strong>g . The thick l<strong>in</strong>e is<br />

the runn<strong>in</strong>g average over 10 po<strong>in</strong>ts. The broken bar <strong>in</strong>dicates the overall<br />

stack quality - black good (0.9) and white bad (0.1). (a) E x (squares) and<br />

E y (triangles) pattern quality predicted as a function of stack number.<br />

(b) E x - H y (squares) and E y − H x (triangles) amplitude ratio. (c) The<br />

correlation coefficients of E x to H y (squares) and E y to H x (triangles). . 79<br />

5.8 Results from <strong>in</strong>ter channel tra<strong>in</strong><strong>in</strong>g. (a) The SSE as a function of tra<strong>in</strong><strong>in</strong>g<br />

epoch. (b) Deviation between manually classified stack quality and network<br />

predicted for 250 stacks. 235 stacks were classified similar to manual<br />

classification. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81<br />

5.9 Relative significance of various <strong>in</strong>puts to the networks, viz amplitude ratios<br />

(A1 and A2), correlation coefficients (C1 and C2) and five pattern<br />

qualities (E x , E y , H x , H y andH z ). The error deviation aga<strong>in</strong>st each <strong>in</strong>put is<br />

a measure of its significance to the Neural Network. . . . . . . . . . . . 82


xix<br />

5.10 Comparison of MT apparent resistivity and phase computed from different<br />

mode of edit<strong>in</strong>g of <strong>data</strong> from site G12 . Filled circles represent xy and<br />

diamonds represent yx components. (a) Us<strong>in</strong>g all stacks available. (b) By<br />

neural network edit<strong>in</strong>g. (c) By manual edit<strong>in</strong>g. . . . . . . . . . . . . . . 83<br />

5.11 Comparison of MT apparent resistivity and phase computed from different<br />

mode of edit<strong>in</strong>g of <strong>data</strong> from site VP12 . Filled circles represent xy and<br />

diamonds represent yx components. (a) Us<strong>in</strong>g all stacks available. (b) By<br />

neural network edit<strong>in</strong>g. (c) By manual edit<strong>in</strong>g. . . . . . . . . . . . . . . 84<br />

5.12 Comparison of MT apparent resistivity and phase computed from different<br />

mode of edit<strong>in</strong>g of <strong>data</strong> from site JN10 . Filled circles represent xy and<br />

diamonds represent yx components. (a) Us<strong>in</strong>g all stacks available. (b) By<br />

neural network edit<strong>in</strong>g. (c) By manual edit<strong>in</strong>g. . . . . . . . . . . . . . . 84<br />

5.13 Comparison of MT apparent resistivity and phase computed from different<br />

mode of edit<strong>in</strong>g of <strong>data</strong> from site TT08 . Filled circles represent xy and<br />

diamonds represent yx components. (a) Us<strong>in</strong>g all stacks available. (b) By<br />

neural network edit<strong>in</strong>g. (c) By manual edit<strong>in</strong>g. . . . . . . . . . . . . . . 85<br />

5.14 Comparison of manual and neural signal pick<strong>in</strong>g for site TT8. The diamonds<br />

present the neural pick<strong>in</strong>g and crosses, the manual. . . . . . . . . 86<br />

6.1 Examples where robust statistical methods are desirable: (a) A one dimensional<br />

distribution with heavy tails (b) A distribution <strong>in</strong> two dimensions<br />

fitted to straight l<strong>in</strong>es. Adapted from Flannery et al. [1992] . . . . . . . . 92<br />

6.2 Schematic diagram show<strong>in</strong>g loss, <strong>in</strong>fluence and weight functions for Least<br />

Square (LS or L2) , Least Absolute (L1) Huber and Tukey estimators.<br />

Values shown <strong>in</strong> y axis are arbitrary See text for discussion. Adapted from<br />

Zhang [1996] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94<br />

6.3 Responses of different <strong>in</strong>fluence functions to a set of residuals from MT<br />

<strong>data</strong> process<strong>in</strong>g. Station VP13 Shows the Ex residuals for 0.1875 Hz for<br />

all the 105 stacks. (b) Shows the response of three <strong>in</strong>fluence functions to<br />

the residuals. See text for discussion. . . . . . . . . . . . . . . . . . . . . 95<br />

6.4 Comparison of Least Square and Jackknife estimation of MT transfer functions<br />

for station VP10 (a) Jackknife difference for the first iteration. (b)<br />

Variance as a function of iteration number. (c) and (d) comparison of ρ<br />

and φ values from LS & JK process<strong>in</strong>g . . . . . . . . . . . . . . . . . . . 98<br />

6.5 Comparison of Least Square and Jackknife estimation of MT transfer functions<br />

for station VP13 (a) Jackknife difference for the first iteration. (b)<br />

Variance as a function of iteration number. (c) and (d) comparison of ρ<br />

and φ values from LS & JK process<strong>in</strong>g . . . . . . . . . . . . . . . . . . . 99<br />

6.6 Comparison of Least Square and Robust process<strong>in</strong>g of magnetotelluric <strong>data</strong><br />

station TT08 for frequency 0.0791Hz . Triangles represent LS process<strong>in</strong>g,<br />

and stars represent robust (RB) process<strong>in</strong>g (a) and (b) time series of Ex<br />

residuals. C) Quantile Quantile plot of Ex residuals d) MT apparent resistivity<br />

and phase values from LS and robust process<strong>in</strong>g. See text for<br />

discussion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102


xx<br />

6.7 Flow chart for robust process<strong>in</strong>g scheme. The gray area represents the proposed<br />

<strong>in</strong>itialization of the transfer functions <strong>us<strong>in</strong>g</strong> Jackknife. This rout<strong>in</strong>e,<br />

concentrates on the section averag<strong>in</strong>g of MT <strong>data</strong>. See text for discussion. 103<br />

6.8 Data flow through the flow chart represent<strong>in</strong>g robust process<strong>in</strong>g of MT<br />

<strong>data</strong>. The gray area represents the proposed Jackknife <strong>in</strong>itialization. (a)<br />

Process A uses Jackknife (JK) as <strong>in</strong>itial guess and (b) process B uses Least<br />

Square (LS) as <strong>in</strong>itial guess. See text for discussion. . . . . . . . . . . . . 104<br />

6.9 Comparison of robust process<strong>in</strong>g results <strong>us<strong>in</strong>g</strong> Least Square (LS) and Jackknife<br />

(JK) <strong>in</strong>itialization for stations JN12, VP16, OK16 and VP12. Process<br />

A refers to robust process<strong>in</strong>g with JK <strong>in</strong>itialization, where as Process<br />

B refers robust process<strong>in</strong>g with LS <strong>in</strong>itialization See legend for symbol<br />

identification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106<br />

6.10 Comparison of robust process<strong>in</strong>g results <strong>us<strong>in</strong>g</strong> Least Square (LS) and Jackknife<br />

(JK) <strong>in</strong>itialization for stations VP14,TT08,TT04 and OK18. Process<br />

A refers to robust process<strong>in</strong>g with JK <strong>in</strong>itialization, where as Process B<br />

refers robust process<strong>in</strong>g with LS <strong>in</strong>itialization. See legend for symbol identification<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107<br />

6.11 Spectrum estimation <strong>in</strong> MT <strong>us<strong>in</strong>g</strong> band averag<strong>in</strong>g. (a) to (d) w<strong>in</strong>dows <strong>in</strong><br />

frequency with different radii. Though the w<strong>in</strong>dow radius seems to <strong>in</strong>crease<br />

as period <strong>in</strong>creases, effective bandwidth rema<strong>in</strong>s the same, as spectrum <strong>in</strong><br />

long period conta<strong>in</strong>s fewer Fourier harmonics. (e) Sample E x Ex ∗ spectrum<br />

<strong>in</strong> the band 4, with target frequencies projected as dotted l<strong>in</strong>es. It is<br />

common to have 10 12 target frequencies per band. . . . . . . . . . . . . 110<br />

6.12 Apparent resistivities as a function of frequency w<strong>in</strong>dow length. Triangles<br />

represent ρ xy and circles ρ yx . Error bars represents 95% confidence <strong>in</strong>terval.<br />

See text for discussion. . . . . . . . . . . . . . . . . . . . . . . . . . . . 111<br />

6.13 Concepts of robust band averag<strong>in</strong>g. (a) Magnitude of cross spectrum between<br />

H x and H y for station VP13, around a target frequency 6 Hz. The<br />

spectra are multiplied by a Parzen w<strong>in</strong>dow. LS least square, RB Robust.<br />

(b) Quantile Quantile plot for real part of the same cross spectrum.<br />

Inverted triangles unweighted, Circles robust weighted. See text for discussion.<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112<br />

6.14 Comparison of robust and least square band averag<strong>in</strong>g for different frequencies<br />

and band (Parzen) radius for station VP13. Magnitude of H x Hy<br />

∗<br />

(nT 2 /Hz) cross spectra are plotted for all the cases. X-axis show the frequency<br />

b<strong>in</strong>s. Solid l<strong>in</strong>e represents robust average and broken l<strong>in</strong>e represents<br />

Least Square average.The left column represent the the band averag<strong>in</strong>g for<br />

different target frequencies, where as the right column represents the band<br />

averag<strong>in</strong>g with different radius length for the same target frequency. See<br />

text for discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121<br />

6.15 Flow chart represent<strong>in</strong>g robust band averag<strong>in</strong>g. Gray area represents the<br />

proposed process<strong>in</strong>g rout<strong>in</strong>e (a) Shows the different steps <strong>in</strong> robustly estimat<strong>in</strong>g<br />

the spectral matrix from raw time series Schematic diagrams (b) to<br />

(e) shows four process<strong>in</strong>g schemes def<strong>in</strong>ed. SS s<strong>in</strong>gle station, RR remote<br />

reference, RB robust band averag<strong>in</strong>g, LS Least square . . . . . . . . . 122


xxi<br />

6.16 Results of s<strong>in</strong>gle station (SS) and remote reference (RR) process<strong>in</strong>g for<br />

station VP10. Process C (SS) and E (RR) use least squares band averag<strong>in</strong>g,<br />

whereas process D (SS) and F (RR) use robust band averag<strong>in</strong>g. The<br />

MT transfer functions were derived from a robust section averag<strong>in</strong>g §6.2.4<br />

from the spectra sets produced by Process C to F. See text for discussion. 123<br />

6.17 Comparison of upward and downward biased estimates of apparent resistivities<br />

for JN12 for processes C and D (a) XY component and (b) YX<br />

component. UP - up biased (E- reference) and DN down biased (Hreferences).<br />

Symbols are same for both plots. See text for discussion.<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123<br />

6.18 Comparison of upward and downward biased estimates of apparent resistivities<br />

for VP16 for processes C and D (a) XY component and (b) YX<br />

component. UP - up biased (E- reference) and DN down biased (Hreferences).<br />

Symbols are same for both plots. See text for discussion.<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124<br />

6.19 Comparison of upward and downward biased estimates of apparent resistivities<br />

for KG02 for processes C and D (a) XY component and (b) YX<br />

component. UP - up biased (E- reference) and DN down biased (Hreferences).<br />

Symbols are same for both plots. See text for discussion.<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124<br />

6.20 Comparison of upward and downward biased estimates of apparent resistivities<br />

for OK18 for processes C and D (a) XY component and (b) YX<br />

component. UP - up biased (E- reference) and DN down biased (Hreferences).<br />

Symbols are same for both plots. See text for discussion.<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125<br />

6.21 Comparison of telluric predicted coherency, apparent resistivity and phase<br />

values from robust process<strong>in</strong>g with and without robust band averag<strong>in</strong>g for<br />

stations VP14 and TT08. The explanation for symbols are given <strong>in</strong> legend.<br />

See text for discussion. . . . . . . . . . . . . . . . . . . . . . . . . . . . 126<br />

6.22 Comparison of telluric predicted coherency, apparent resistivity and phase<br />

values from robust process<strong>in</strong>g with and without robust band averag<strong>in</strong>g for<br />

stations TT04 and OK18. The explanation for symbols are given <strong>in</strong> legend.<br />

See text for discussion. . . . . . . . . . . . . . . . . . . . . . . . . . . . 127<br />

6.23 Comparison of telluric predicted coherency, apparent resistivity and phase<br />

values from robust process<strong>in</strong>g with and without robust band averag<strong>in</strong>g for<br />

stations JN12 and VP16. The explanation for symbols are given <strong>in</strong> legend.<br />

See text for discussion. . . . . . . . . . . . . . . . . . . . . . . . . . . . 128<br />

6.24 Comparison of telluric predicted coherency, apparent resistivity and phase<br />

values from robust process<strong>in</strong>g with and without robust band averag<strong>in</strong>g for<br />

stations OK16 and VP12. The explanation for symbols are given <strong>in</strong> legend.<br />

See text for discussion. . . . . . . . . . . . . . . . . . . . . . . . . . . . 129<br />

6.25 Comparison of results from two different process<strong>in</strong>g scheme for MT <strong>data</strong>:<br />

Left panel shows the results from ProcMT and the right panel shows the<br />

results from the robust process<strong>in</strong>g and ANN edit<strong>in</strong>g. . . . . . . . . . . . 130


xxii<br />

7.1 Comparison of robust process<strong>in</strong>g (RB) with and without neural network<br />

(NN) edit<strong>in</strong>g for Station JN12. (a) Apparent resistivity and phase values<br />

from both process<strong>in</strong>g schemes. (b) Comparison of robust and neural<br />

network weights. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135<br />

7.2 Comparison of robust process<strong>in</strong>g (RB) with and without neural network<br />

(NN) edit<strong>in</strong>g for Station OK18. (a) Apparent resistivity and phase values<br />

from both process<strong>in</strong>g schemes. (b) Comparison of robust and neural<br />

network weights. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136


List of Tables<br />

1.1 Frequency bands for measurements, GMS05 (after Metronix [1997]). . . . 14<br />

2.1 Classification of noise sources that affect MT measurements. . . . . . . . 25<br />

4.1 The locations and measurement details for the MT stations . . . . . . . 53<br />

6.1 Few commonly used <strong>in</strong>fluence functions. Adapted from Zhang [1996]. . . 93<br />

xxiii


Chapter 1<br />

<strong>Magnetotelluric</strong>s: Basic Theory,<br />

Sensors and Field Procedures<br />

1


2<br />

1.1 Introduction<br />

<strong>Magnetotelluric</strong> signals orig<strong>in</strong>ate <strong>in</strong> Earth’s magnetosphere and atmosphere from different<br />

phenomena. The constant bombardment of solar w<strong>in</strong>d on Earth’s magnetosphere leads<br />

to its deformation and which <strong>in</strong> turn disturb the terrestrial magnetic field. This time<br />

vary<strong>in</strong>g phenomena constitutes MT signals, which have periods usually above 1 second.<br />

The worldwide thunderstorm activity generates high frequency ( > 1 Hz) electromagnetic<br />

signals, which propagate around the globe <strong>in</strong> the Earth-ionosphere waveguide and constitute<br />

the higher frequency part of magnetotelluric signals. These electromagnetic waves<br />

reach Earth’s surface as quasi-homogeneous waves and a small part of it penetrates the<br />

conductive Earth as quasi-planar waves. The <strong>in</strong>duced electromagnetic response of Earth,<br />

both amplitude and phase, depends on the subsurface electrical conductivity structure.<br />

The relation between electric and magnetic fields at the surface of Earth forms frequency<br />

doma<strong>in</strong> transfer functions, which can be <strong>in</strong>terpreted <strong>in</strong> terms of the subsurface structure.<br />

<strong>Magnetotelluric</strong>s is a method of electromagnetic exploration that uses natural electromagnetic<br />

waves as source field (Vozoff [1972]). In the follow<strong>in</strong>g sections the basic ideas<br />

of magnetotellurics viz, planar assumption for source field, properties of EM field <strong>in</strong><br />

conductive earth, estimation of magnetotelluric transfer functions/impedances and their<br />

behavior over different type of media are discussed. Accurate and simultaneous measurement<br />

of the time vary<strong>in</strong>g electromagnetic fields is the prelim<strong>in</strong>ary requirement to<br />

obta<strong>in</strong> MT transfer functions. A brief discussion about sensors, record<strong>in</strong>g devices and<br />

field procedures commonly adapted for MT <strong>data</strong> acquisition is given <strong>in</strong> § 1.2 and § 1.3.<br />

1.1.1 Source fields<br />

<strong>Magnetotelluric</strong> methods evolved out of the observation of similar variation <strong>in</strong> Earth<br />

current and magnetic fields. Telluric methods were already <strong>in</strong> use as an exploration<br />

tool. Tikhonov [1950] and Cagniard [1953] exam<strong>in</strong>ed the relationship between horizontal<br />

orthogonal electromagnetic components and developed formulae to estimate impedance<br />

of subsurface from the simultaneous measurement of these electromagnetic field components.<br />

In essence, the formulae between the two components are valid only if the fields<br />

do not have a lateral gradient over scale lengths that varies with frequency. For example<br />

for an electromagnetic wave, with a frequency 10 −3 Hz and <strong>in</strong> a media of resistivity 10 3<br />

ohm.m, the scale length = ∼350 km (Wait [1954]). The scale length and uniformity of<br />

harmonic source field (with frequencies < 1 Hz) <strong>in</strong> the Earth’s ionosphere and magnetosphere<br />

has been established by Dungey [1955]. Monitor<strong>in</strong>g and study of atmospheric<br />

electricity / lightn<strong>in</strong>g generation, propagation of sub-ionospheric waves provide adequate<br />

characterization of audio-frequency variations <strong>in</strong> electromagnetic waves.<br />

At a distance from the source, the electromagnetic wave front becomes locally planar,<br />

<strong>in</strong> the sense the oscillation of the wave is only <strong>in</strong> a plane that is perpendicular to the<br />

propagation direction. Such waves are called plane waves or plane polarized waves (Figure<br />

1.1). <strong>Magnetotelluric</strong>s envisages its <strong>in</strong>duc<strong>in</strong>g electromagnetic fields as plane polarized.<br />

When such waves are <strong>in</strong>cident on Earth’s surface, maximum energy is reflected. In the<br />

context of MT, it was shown by Cagniard [1953] that the refracted wave propagates down<br />

nearly vertically, due to the large contrast <strong>in</strong> the speed of electromagnetic waves <strong>in</strong> the


3<br />

Figure 1.1: Quasi planar nature of electromagnteic wave front at great distance from<br />

source<br />

Figure 1.2: Behaviour of plane EM wave at the surface of a conduct<strong>in</strong>g Earth<br />

atmosphere and <strong>in</strong> the Earth. (or the large contrast of conductivities between the two<br />

media). It follows that the direction of propagation of electromagnetic fields with<strong>in</strong> the<br />

Earth does not depend on the angle at which it hit the Earth’s surface. Figure 1.2 depicts<br />

this situation.<br />

1.1.2 Maxwell’s equations<br />

Once refracted the electromagnetic fields propagate through the Earth. The electromagnetic<br />

fields <strong>in</strong> isotropic and homogeneous media (of constant electric conductivity σ<br />

[S/m]) of uniform electric permitivity ɛ [As/Vm] and magnetic permeability µ [Vs/Am]<br />

are described by the Maxwell’s equations. Consider<strong>in</strong>g the fields with harmonic temporal


4<br />

variation (e iwt ) these equations are;<br />

∇ × E = iωµH<br />

∇ × H = iωɛE + σE ≈ σE<br />

∇.H = 0<br />

∇.E = q/ɛ ∼ = 0<br />

(1.1)<br />

The electric current density j [A/m 2 ] is proportional to the electric field accord<strong>in</strong>g to<br />

Ohm’s law;<br />

j = σE (1.2)<br />

where q [As/m 3 ] is the volume density of charge, H the magnetic field [Vs/m 2 ],<br />

E [V/m] the electric field and ω=2πf, the angular frequency. Both permittivity and<br />

permeability <strong>in</strong> the earth are assumed to have approximately constant values.<br />

In the Cartesian system of coord<strong>in</strong>ates, the first two equations of system 1.1 can be<br />

written as<br />

∣ i j k ∣∣∣∣∣<br />

iωµH =<br />

∂ ∂ ∂<br />

∂x ∂y ∂x<br />

∣ E x E y H x<br />

and<br />

∣<br />

(1.3)<br />

i j k ∣∣∣∣∣<br />

σE =<br />

∂ ∂ ∂<br />

∂x ∂y ∂x<br />

∣ H x H y H x<br />

It follows that<br />

∂E x<br />

∂z<br />

= iωµH y<br />

∂H y<br />

∂z<br />

= −σE x<br />

(1.4)<br />

By tak<strong>in</strong>g partial derivative of each equations and mak<strong>in</strong>g appropriate substitutions,<br />

∂ 2 F<br />

∂z 2 + k2 F = 0 (1.5)<br />

Where F = E or H and with the assumption that <strong>in</strong> a homogeneous Earth σ is a<br />

constant. The constant k describes the complex penetration depth 1/k [m] of the EM<br />

field.<br />

1.1.3 Sk<strong>in</strong> depth.<br />

In terms of the diffusion factor describ<strong>in</strong>g the penetration <strong>in</strong> depth of the fields, the<br />

so called sk<strong>in</strong> depth (δ(ω) [m]) <strong>in</strong> a homogeneous earth is def<strong>in</strong>ed as<br />

√ √<br />

2 2<br />

δ(ω) =<br />

|k 2 | = (1.6)<br />

ωµσ


5<br />

which represents the exponential decay of the EM field amplitude with depth. At<br />

depth δ(ω) the EM-field amplitude will drop by 1/e with respect to its value at the<br />

surface. The penetration <strong>in</strong> depth of the EM field for a stratified Earth is def<strong>in</strong>ed as a<br />

response function C(ω) = E x /iωH y . In case of a homogeneous Earth C(ω) = 1/k.<br />

Case 1, σ varies along z.<br />

For a 1D stratified Earth of N layers, the penetration <strong>in</strong> depth of the EM fields<br />

measured at the surface is solved iteratively, with a recursive formula described by the<br />

EM – response function C i (ω). This <strong>in</strong>dex refers to the EM-response measured at the<br />

top of the layer I (Weaver [1994]);<br />

where I = N-1,N-2,. . . .1, and<br />

C i (ω) =<br />

1 − r ie −2k it i<br />

k i [1 + r i e −2k i t i ]<br />

(1.7)<br />

r i = 1 − k iC i+1 (ω)<br />

1 + k i C i+1 (ω)<br />

(1.8)<br />

and t i is the thickness of the layer i and k i = ρ i (iωµ) −1 , the diffusion factor <strong>in</strong> the<br />

layer and ρ i is the resistivity of the layer i (See Figure 1.3). The bottom most layer has<br />

the response function 1/k N .<br />

Case 2, σ varies along z and either x or y.<br />

In a 2D earth with strike along the horizontal x-axis and conductivity σ as a function<br />

of z and y, the Maxwell’s equations are de-coupled <strong>in</strong>to two polarization modes. The<br />

decoupl<strong>in</strong>g is valid s<strong>in</strong>ce the EM fields are treated as plane waves. In this context, the<br />

so called TE- polarization mode refers to the tangential electrical field and the TMpolarization<br />

mode to the tangential magnetic field; both components are tangential with<br />

respect to the strike (x axis) of the conductivity structure.<br />

TE – polarization : (E x H y );<br />

TM – polarization: (E y H x );<br />

∂H z<br />

∂y<br />

− ∂Hy<br />

∂z<br />

= σE x<br />

−∂E x<br />

∂z<br />

= iωµH y<br />

−∂E x<br />

∂y<br />

= iωµH z<br />

(1.9)<br />

− ∂Ez<br />

∂y<br />

1.1.4 Impedance tensor:<br />

+ ∂Ey<br />

∂z<br />

= iωµH x<br />

−∂H x<br />

∂z<br />

= σE y<br />

−∂H x<br />

∂y<br />

= σE z<br />

(1.10)<br />

The electrical impedance Z [mV/T] is the ratio between the electric and magnetic<br />

field components, which can be represented as<br />

E=Z H (1.11)


6<br />

Figure 1.3: Diagrammatic representation of 1D, 2D and 3D situations.<br />

In a homogeneous media the ratio of the orthogonal components is<br />

Z = iω/k (1.12)<br />

However, earth cannot be approximated to perfect homogeneous media and it is usual<br />

to assume structural <strong>in</strong>homogeneity for subsurface. Figure 1.3 depicts the 1D, 2D and<br />

3D <strong>in</strong>homogeneity <strong>in</strong> conductivity with<strong>in</strong> earth.<br />

In a general 3D earth, the impedance (transfer function of Earth) is expressed <strong>in</strong>


7<br />

Figure 1.4: Planar view. Off diagonal tensor elements become unequal when geology<br />

has a preferred direction. On rotat<strong>in</strong>g the tensor to the strike, diagonal elements get<br />

m<strong>in</strong>imized. See text for discussion<br />

matrix form <strong>in</strong> Cartesian coord<strong>in</strong>ates,<br />

[ ] [ ]<br />

Ex Zxx Z<br />

=<br />

xy<br />

∗<br />

E y Z yx Z yy<br />

[<br />

Hx<br />

H y<br />

]<br />

(1.13)<br />

Thus each tensor element is Z ij = E i /H j (i,j = x,y).<br />

For a 1-D layered Earth, besides hav<strong>in</strong>g diagonal elements of Z ≈0, the off diagonal<br />

elements are related <strong>in</strong> the form,<br />

Z xy = −Z yx orZ 1D = Z xy = −Z yx . (1.14)<br />

In a 2D earth the diagonal elements of Z vanish (<strong>in</strong> the 2D-strike coord<strong>in</strong>ate system).<br />

And the off diagonal elements become unequal.<br />

[ ] [ ] [ ]<br />

Ex 0 Zxy Hx<br />

=<br />

∗<br />

(1.15)<br />

Ey Z yx 0 Hy<br />

However, if the measurement axes are not <strong>in</strong> alignment with the geological strike, the<br />

diagonal elements may not vanish. The simplest case one can imag<strong>in</strong>e is a geological<br />

fabric <strong>in</strong> a preferred direction, or a structural orientation such as a fold or fault system<br />

(Figure 1.4).<br />

In such cases the tensor can be rotated to an angle θ with the rotation matrix R to align<br />

it with geological strike.<br />

[ ]<br />

cos θ s<strong>in</strong> θ<br />

Zm = RZR T , where R =<br />

(1.16)<br />

− s<strong>in</strong> θ cos θ<br />

with positive θ.


8<br />

Once rotated optimally, the tensor may be written as,<br />

[ ] 0 ZT<br />

Z m =<br />

E<br />

Z T M 0<br />

(1.17)<br />

Where Z T E (transverse electric) is the impedance tensor element parallel to the strike<br />

and Z T M (transverse magnetic) is impedance element perpendicular to the strike direction.<br />

1.1.5 Amplitude and phase of impedance:<br />

The complex impedance (transfer function) Z is usually represented by its amplitude and<br />

phase. The electrical resistivity (<strong>in</strong>verse of σ) as a function of depth can be <strong>in</strong>ferred<br />

by the EM field of the correspond<strong>in</strong>g penetration depths. Resistivity obta<strong>in</strong>ed from the<br />

ratio of the measured electric and magnetic fields is called “apparent” resistivity, which<br />

is frequency dependent. The apparent resistivity ρ aij [Ohm.m] (i,j = x,y) is def<strong>in</strong>ed <strong>in</strong><br />

terms of the transfer function element by the form:<br />

ρ aij = µjZ ij j 2 /ω (1.18)<br />

In case of a homogeneous Earth, the apparent resistivity reflects the true value of<br />

the Earth’s resistivity. Apparent resistivity is what we sense at the surface; the true<br />

resistivity of objects at some depth is masked by <strong>in</strong>terven<strong>in</strong>g material. The phase of<br />

the transfer function element describes the phase shift between the electric and magnetic<br />

field components.<br />

( )<br />

φ = tan −1 im(Zij )<br />

(1.19)<br />

re(Z ij )<br />

In a homogeneous Earth the impedance phase is π/4 (45 0 ). In a 1D layered Earth,<br />

the phase <strong>in</strong>creases over 45 0 when the EM response penetrates <strong>in</strong>to higher conductivity<br />

media. By the same convention, phase decays below 45 0 for the EM response penetrat<strong>in</strong>g<br />

<strong>in</strong>to a less conductive media.<br />

1.2 Sensors and Field Procedures<br />

MT parameters such as apparent resistivity and phase values are derived from simultaneously<br />

measured time vary<strong>in</strong>g components of electric and magnetic fields. <strong>Magnetotelluric</strong><br />

time series consist of five components of the Earth’s electromagnetic field i.e., two orthogonal<br />

components of horizontal electric fields (E x , E y ) and three orthogonal components of<br />

magnetic field (H x , H y , H z ). The measurements are usually made <strong>in</strong> wide bands of overlapp<strong>in</strong>g<br />

frequency ranges, with different sampl<strong>in</strong>g <strong>in</strong>tervals. In the follow<strong>in</strong>g section, an<br />

overview of magnetotelluric field sensors, <strong>data</strong> acquisition equipment and field procedures<br />

given.


9<br />

1.2.1 Introduction<br />

A variety of <strong>in</strong>struments are used to measure the time vary<strong>in</strong>g components of Earth’s<br />

natural electromagnetic fields. Primary requirement of any measur<strong>in</strong>g device is that the<br />

electric and magnetic field measurements must cover a frequency range that is appropriate<br />

to the exploration problem and with sensitivity and accuracy, which is required for the<br />

computation of magnetotelluric transfer functions (Kaufman and Keller [1981]). The<br />

strength and frequency characteristics of natural electromagnetic waves are statistically<br />

known but cannot be predicted <strong>in</strong> advance. The field equipment should have the sufficient<br />

dynamic range and sensitivity to adequately measure the fields, even though, its strength<br />

may vary from hour to hour at a site. For gett<strong>in</strong>g long period (>10 sec) <strong>in</strong>formation,<br />

record<strong>in</strong>gs must be carried out for time rang<strong>in</strong>g from hours to days. This necessitates<br />

the record<strong>in</strong>g equipment to be stable and resistant to adverse weather conditions. The<br />

procedure of acquir<strong>in</strong>g magnetotelluric <strong>data</strong> can be divided <strong>in</strong>to two parts : 1) that of<br />

sensors, which are capable of convert<strong>in</strong>g the electric and magnetic field variations to<br />

voltages and 2) record<strong>in</strong>g of these voltages <strong>in</strong> a way that is suitable for recovery later on.<br />

The modern record<strong>in</strong>g devices are even capable of <strong>in</strong> site <strong>data</strong> process<strong>in</strong>g and prelim<strong>in</strong>ary<br />

<strong>in</strong>version of observed <strong>data</strong>. This facilitates redef<strong>in</strong><strong>in</strong>g of acquisition plans <strong>in</strong> the field itself,<br />

which may save lot of money and time. The follow<strong>in</strong>g sections are devoted to 1) sensors<br />

2) record<strong>in</strong>g equipment and 3) field procedures.<br />

1.2.2 Electric field sensors<br />

At frequencies that are measured <strong>in</strong> magnetotelluric sound<strong>in</strong>g, the electric field is detected<br />

by measur<strong>in</strong>g the voltage drop between pairs of electrode that are <strong>in</strong> contact with<br />

Earth. The electric field can be thought as the limit of voltage drop when electrode<br />

spac<strong>in</strong>g tends to zero. In this way, it may be advantageous to have short electrode separation<br />

if one wants to measure electric field correctly. But <strong>in</strong> practice there are two<br />

drawbacks for this. The current density may change at contact of rocks with different<br />

conductivity. A large separation of electrode will average out such small gradients to<br />

obta<strong>in</strong> a regional picture. More over, there is a chance that the voltage measured over<br />

small electrode spac<strong>in</strong>g may be below the noise level (∼ 1µV/Hz) of the sensor. Typical<br />

separation between electrode contact ranges from 50 to 200 m. The electrodes that are<br />

commonly used are non-polariz<strong>in</strong>g electrodes. Such electrodes consist of a metal immersed<br />

<strong>in</strong> saturated solution of one of its salts. The advantage of <strong>us<strong>in</strong>g</strong> non-polariz<strong>in</strong>g<br />

electrodes is that the voltage difference between a pair of electrodes is relatively stable.<br />

If metal electrodes such as copper or steel are used, relatively large potential difference<br />

result<strong>in</strong>g from electrochemical reactions at the metal surface can be present and these can<br />

vary with time <strong>in</strong> an unpredictable manner. Several types of electrodes are now <strong>in</strong> use<br />

(Petiau and Dupis [1980]) the <strong>data</strong> analysed <strong>in</strong> the thesis was collected with Cd-CdCl 4<br />

cell. One disadvantage of this cell is the high toxicity of CdCl 4 solution. The Electric<br />

Field Probe (EPF05 from M/s Metronix) showed good stability even dur<strong>in</strong>g long period<br />

measurements (1-2 days).<br />

The electrical field sensors are planted <strong>in</strong> small pits (∼0.2 m) at least few hours before<br />

they are used. The pits need to be water saturated and covered to retard evaporation


10<br />

Figure 1.5: Implant<strong>in</strong>g of Electrical Field sensor<br />

temperature changes. In some types of soil, a few milli-Amperes of current can quickly<br />

cause a significant polarization of the ground around the electrodes, lead<strong>in</strong>g to long<br />

exponential decay of this potential. Contact resistance of the telluric electrodes should<br />

be as low as possible; a typical value measured between two electrodes spaced at 100 m<br />

<strong>in</strong> the highly resistive Southern Granulite Terra<strong>in</strong> (§ 4.2) was between 200 - 1,000 and<br />

10,000Ohm.m. Field photograph <strong>in</strong> Figure 1.5 shows a typical implant of electrical field<br />

sensor.<br />

1.2.3 Magnetic Sensors<br />

The problem of detect<strong>in</strong>g magnetic field variations with the required accuracy is much<br />

more difficult than that of detect<strong>in</strong>g electric field variations (Kaufman and Keller [1981]).<br />

The magnetic sensors are required to detect variation <strong>in</strong> magnetic field for the frequency<br />

range 10 −4 Hz to 10 4 Hz. The dynamic range of magnetic field varies from as low as few<br />

tenths of a nano-Tesla for micro-pulsations to as high as few nTs for diurnal variations.<br />

Also the magnetometers should ideally have same response over such a wide range of<br />

frequency and amplitude as well as it should be robust to the severe fluctuations <strong>in</strong><br />

temperature. These requirements make design<strong>in</strong>g of magnetometers difficult. Some of<br />

the commonly used magnetometers are Super Conduct<strong>in</strong>g Quantum Interference Devices<br />

(SQUID see Clarke et al. [1983], Fluxgate and <strong>in</strong>duction coil (Karmann [1977]). A short<br />

description is given on the <strong>in</strong>duction coil magnetometers, which were used to collect the<br />

<strong>data</strong> analysed <strong>in</strong> this thesis.<br />

1.2.3.1 Induction coil magnetometers<br />

At its simplest, the <strong>in</strong>duction coil is a loop of wire which produces a voltage proportional<br />

to its area multiplied by the time derivative of B across the cross sectional area<br />

of the loop. For the low field strength and low frequencies of <strong>in</strong>terest <strong>in</strong> magnetotelluric<br />

surveys, the <strong>in</strong>duction coils are fabricated with a high permeability µ metal core about<br />

which the coil is wound. A very large number of turns must be used to provide a measurable<br />

voltage output from a coil. For example, an <strong>in</strong>duction coil with 50,000 turns of


11<br />

Figure 1.6: Installation of <strong>in</strong>duction coil magnetometer<br />

Figure 1.7: Simplified equivalent circuit for <strong>in</strong>duction coil magnetometer<br />

1m 2 area produces 31.41 nV at 1000 sec (Kaufman & Keller, 1981). A further problem<br />

<strong>in</strong> the design of sensitive <strong>in</strong>duction coil is that if a very f<strong>in</strong>e wire gauge is used, the total<br />

resistance is very high. On the other hand a heavier gauge wire with lower resistivity will<br />

make the <strong>in</strong>duction coil heavier, limit<strong>in</strong>g its mobility. One such type of magnetometer<br />

namely MFS05 (M/s Metronix) was used to measure magnetic field <strong>in</strong> the present study<br />

(Field photograph given <strong>in</strong> Figure 1.6). In addition to hav<strong>in</strong>g a large resistance, an <strong>in</strong>duction<br />

coil will also have an appreciable <strong>in</strong>ductance and capacitance between w<strong>in</strong>d<strong>in</strong>gs. A<br />

simplified equivalent circuit is shown <strong>in</strong> the Figure 1.7. This circuit will have a resonant<br />

frequency above which voltage output decreases with <strong>in</strong>creas<strong>in</strong>g frequency.<br />

The frequency dependence of <strong>in</strong>duction coil response is ideally suited to the measurement<br />

of natural field, where the 1/f dependence of natural field strength (Figure 2.1)


12<br />

Figure 1.8: Theoretical response function for MFS05 magnetometer (from Pulz and Ritter<br />

[2001])<br />

compensate for the low output from the <strong>in</strong>duction coil for the longer periods (Figure 1.8,<br />

adapted from Pulz and Ritter [2001]). In higher frequencies frequency- <strong>in</strong>dependent magnetometer<br />

response is obta<strong>in</strong>ed by feed<strong>in</strong>g back <strong>in</strong>to coil, a magnetic field proportional to<br />

the current. These techniques effectively extend the dynamic range of the magnetometers.<br />

1.2.4 Record<strong>in</strong>g systems<br />

Historically, MT studies have been concerned with the determ<strong>in</strong>ation of the electrical<br />

resistivity of the Earth’s crust or upper mantle on a regional scale. For that purpose,<br />

signals <strong>in</strong> the period range between 10 s and 10000 s have been recorded with <strong>data</strong><br />

sampl<strong>in</strong>g rates <strong>in</strong> the order of several seconds to m<strong>in</strong>utes. Accord<strong>in</strong>gly, the record<strong>in</strong>g<br />

times for long period MT (LMT) <strong>data</strong> have been <strong>in</strong> the order of weeks to months at a site.<br />

The amount of <strong>data</strong> that could be collected dependents on the progress <strong>in</strong> <strong>data</strong> storage<br />

technology. In the 1950’s and 1960’s, analogue <strong>data</strong> were recorded on photographic<br />

film or paper chart recorders and digitised manually at a later stage (For eg. See Vozoff<br />

[1972]). From the 1970’s digital <strong>data</strong> were stored directly, However, on modified analogue<br />

audiocassette recorders (Allsopp et al. [1973]). In the 1980’s, with the arrival of digital<br />

tape and floppy disks, a new era of digital record<strong>in</strong>g systems began and the most advanced<br />

MT systems today use rugged hard drives (Ritter et al. [1998]). Robustness, compactness<br />

and low weight are desirable features for all geophysical <strong>in</strong>struments. Because of their long<br />

deployment times, <strong>data</strong> loggers must have good long-term stability; they must operate<br />

very reliably and have low power consumption. The <strong>in</strong>vention of the microprocessor <strong>in</strong> the<br />

1980’s opened the field to the <strong>in</strong> field <strong>data</strong> process<strong>in</strong>g and modell<strong>in</strong>g (Clarke et al. [1983]).<br />

This also facilitated the measurement of high frequency magnetotelluric <strong>data</strong> (audiomagnetotelluric,<br />

AMT). AMT (Hoover et al. [1976]) <strong>data</strong> are natural electromagnetic<br />

variation signals, typically <strong>in</strong> the frequency range 20,000 Hz - 0.01 Hz. Because a large<br />

volume of <strong>data</strong> is produced <strong>in</strong> a relatively short time, most AMT <strong>in</strong>struments are designed<br />

as real-time systems. Short, non-cont<strong>in</strong>uous segments of time series <strong>data</strong> are processed<br />

simultaneously, enabl<strong>in</strong>g on-l<strong>in</strong>e quality control of the stacked results. The correspond<strong>in</strong>g<br />

electromagnetic fields penetrate only the first hundred meters or first few kilometres of the<br />

Earth’s crust, which are economically important (Egbert and Livelybrooks [1996], Garcia


13<br />

and Jones [2002]). The signal strength <strong>in</strong> the high frequency range varies enormously over<br />

the spectrum shown <strong>in</strong> Figure 2.1. Artificially generated signals from the ma<strong>in</strong> power<br />

supplies or electric railways exceed the natural signals by several magnitudes <strong>in</strong> narrow<br />

frequency ranges. To elim<strong>in</strong>ate these narrow band noises, MT <strong>in</strong>struments are equipped<br />

with special hardware filters (notches). In the frequency range of the dead band around<br />

1 Hz, the natural signal activity is at a m<strong>in</strong>imum. Hence, record<strong>in</strong>g <strong>in</strong> narrow frequency<br />

bands is necessary to ensure optimum dynamic range of the signals, while LMT (>10 sec)<br />

<strong>data</strong> are usually recorded <strong>in</strong> one broad frequency band. The remote reference technique<br />

(Gamble et al. [1979], also see § 3.3.6) that brought major improvement <strong>in</strong> the MT<br />

impedance estimates requires synchronized record<strong>in</strong>g <strong>in</strong> two MT <strong>data</strong> acquisition devices.<br />

Clarke et al. [1983] describes a radio telemetry system for simultaneous record<strong>in</strong>g of <strong>data</strong><br />

from two stations. Hard wir<strong>in</strong>g and synchronized record<strong>in</strong>g <strong>us<strong>in</strong>g</strong> a crystal clock are the<br />

other options. With the advent of global position<strong>in</strong>g system (GPS), it is now possible to<br />

have both the measur<strong>in</strong>g devices synchronized to a very high precision. The two sets of<br />

MT equipment should follow a common record<strong>in</strong>g timetable, so as to facilitate remote<br />

reference process<strong>in</strong>g. Very recently the concept of computer network<strong>in</strong>g is be<strong>in</strong>g adopted<br />

for MT measur<strong>in</strong>g devices, with each MT unit is envisaged as a node <strong>in</strong> a network,<br />

connected by cheap coaxial network cables/telephone cables.<br />

In the present study, the <strong>data</strong> were collected <strong>us<strong>in</strong>g</strong> Geophysical Measur<strong>in</strong>g System<br />

(GMS05) manufactured by M/s Metronix GmBH, Germany. On standard, the GMS 05<br />

is equipped for the 5 channel MT and AMT method. GMS05 consists of follow<strong>in</strong>g components.<br />

1. Geophysical Process<strong>in</strong>g Unit – GPU 05<br />

2. Sensor Connection Box – SDB 05<br />

3. Display Unit – HDU06<br />

4. 3 Magnetic Field Sensors MFS 06<br />

5. 4 Electric field components<br />

The sensors used by this system is already been discussed <strong>in</strong> § 1.2. The first two<br />

components are described below.<br />

The Geophysical Process<strong>in</strong>g Unit(GPU) houses the major electronic circuitry for condition<strong>in</strong>g,<br />

digitis<strong>in</strong>g and stor<strong>in</strong>g of magnetotelluric <strong>data</strong>. It can receive up to 8 channels<br />

of <strong>data</strong> and digitises it <strong>us<strong>in</strong>g</strong> 16-bit A/D converter, with a dynamic range of 132 dB.<br />

Data are stored <strong>in</strong> hard drive of 1GB capacity. The <strong>in</strong>tel 486-driven mother board works<br />

on DOS platform and performs <strong>in</strong>field <strong>data</strong> process<strong>in</strong>g and storage. This also allows the<br />

operator to <strong>in</strong>spect the time series and computed MT parameters (§ 3) onl<strong>in</strong>e. Communication<br />

is provided through serial or parallel connectors with a maximum <strong>data</strong> transfer at<br />

115 k B/s. External power supply required by the system (12 V DC) is usually provided<br />

with standard car batteries (100 Ah or better). For remote reference applications a highly<br />

stable precision real time clock (PRC) synchronizes the GMS05. The clocks synchronize<br />

automatically as soon as the signals from the GPS satellites are received. The accuracy<br />

of the PRC is better than 10 −12 seconds <strong>in</strong> long term as it runs synchronously with the


14<br />

Figure 1.9: Block diagram depict<strong>in</strong>g different signal process<strong>in</strong>g steps <strong>in</strong> GMS05 (adapted<br />

from Metronix [1997])<br />

Figure 1.10: GPU05 (ma<strong>in</strong> unit) and HDU05 (display unit) dur<strong>in</strong>g <strong>data</strong> acquisition<br />

Caesium clocks on board of the satellites. Figure 1.9 shows a schematic diagram of the<br />

digital process<strong>in</strong>g steps <strong>in</strong> GMS05. Figure 1.10 shows a field photograph of GMS05, <strong>in</strong><br />

operation.<br />

The GMS05’s total frequency range from (4096 sec) −1 to 8192 Hz is split <strong>in</strong>to a total<br />

of 5 bands. The factor between lowest and highest frequency for one band is always 32.<br />

In the table 1.1, an overview of the channel-sampl<strong>in</strong>g rates is given.<br />

The SDB 05 is used for <strong>in</strong>terconnection of different sensors with the GPU 05. The<br />

sensor connection box has three <strong>in</strong>puts for the magnetometers and <strong>in</strong>puts to connect the<br />

Band Lower freq. Hz Upper freq. Hz Sampl<strong>in</strong>g rate Hz<br />

1 256 Hz 8,192 Hz 32,768 Hz<br />

1 8 Hz 256 Hz 1024 Hz<br />

3 1/4s 8 Hz 32 Hz<br />

4 1/128s 1/4s 1 Hz<br />

5 1/4096s 1/128s 1/32s<br />

Table 1.1: Frequency bands for measurements, GMS05 (after Metronix [1997]).


15<br />

electric field probes for E x and E y measurement. The small waterproof box also conta<strong>in</strong>s<br />

the preamplifiers for the electric fields.<br />

1.3 Field Procedures<br />

1.3.1 Survey design<br />

As any modern geophysical survey, magnetotelluric surveys are also planned before fieldwork<br />

starts. Though the survey must accommodate changes due certa<strong>in</strong> unforseen circumstances,<br />

the overall plan rema<strong>in</strong>s the same. Two major decisions to be made before<br />

start of survey are station spac<strong>in</strong>g and m<strong>in</strong>imum record<strong>in</strong>g time. M<strong>in</strong>imum record<strong>in</strong>g<br />

time depends on the depth of <strong>in</strong>terest and the geological condition. Forward modell<strong>in</strong>g<br />

of an assumed model can provide m<strong>in</strong>imum safe record<strong>in</strong>g time. For deep crustal studies,<br />

m<strong>in</strong>imum record<strong>in</strong>g of 2 to 3 days are required over a resistive terra<strong>in</strong>. For basement<br />

configuration <strong>in</strong> hydrocarbon exploration, a record<strong>in</strong>g times range from 1 day to 2 days.<br />

2<br />

M<strong>in</strong>imum site spac<strong>in</strong>g also depends on the conductivity structure and required lateral<br />

resolution. Spac<strong>in</strong>g of 5 – 10 km is required for a regional survey, where as a spac<strong>in</strong>g of 1<br />

km or less is required for geothermal exploration (for eg. Har<strong>in</strong>arayana et al. [2002]). In<br />

the current survey, the stations were spaced 8-10 km (§ 4.3.1) and more densely spaced<br />

measurements were carried out over suspected geological contacts.<br />

1.3.2 Site selection<br />

Site selection is one of the major factors govern<strong>in</strong>g the <strong>data</strong> quality. A good site can<br />

yield quality <strong>data</strong> with<strong>in</strong> a short duration of record compared to a noisy site. Cultural<br />

noises aris<strong>in</strong>g from various sources such as power l<strong>in</strong>es, electric railways, irrigation pumps,<br />

vehicular traffic, radar transmitters etc may act as non planar source (§ 1.1.1) field and<br />

corrupt the computed transfer function (see § 2.4 for more details). An overview of noise<br />

sources affect<strong>in</strong>g MT measurements is given <strong>in</strong> § 2.3. The scenario becomes complicated<br />

<strong>in</strong> areas like south India, where such noises travel great distances as the upper crust<br />

is highly resistive (Har<strong>in</strong>arayana et al. [2003]). It is a practice to avoid high-tension<br />

power l<strong>in</strong>es and electric rail l<strong>in</strong>es by at least 2 km and vehicular traffic by 200 m. It is<br />

advisable to avoid any natural feature that might affect the electric field such as abrupt<br />

topographical relief, water bodies and rock outcrops. In most surveys carried out <strong>in</strong><br />

India, dry and ploughed fields provide an ideal site.<br />

1.3.3 Sensor deployment<br />

Proper deployment of sensors is also essential for good quality MT <strong>data</strong>. Once a site is<br />

selected, a pattern for sensor alignment is planned by survey<strong>in</strong>g. Most preferable is a cross<br />

pattern aligned with geomagnetic directions. The directions are marked with the help of<br />

a compass and sensors are laid <strong>in</strong> alignment with the geomagnetic directions. Geography<br />

and topography some times prevent the array from be<strong>in</strong>g aligned to the geomagnetic axes<br />

or the spread for electrodes be<strong>in</strong>g exactly equal. These factors are taken <strong>in</strong>to consideration


16<br />

Figure 1.11: Typical lay out for MT sensors and <strong>data</strong> acquisition (after Metronix [1997])<br />

while comput<strong>in</strong>g system transfer functions. A typical set-up followed for <strong>data</strong> collection<br />

is shown <strong>in</strong> Figure 1.11.<br />

Electrode pits are prepared few hours ahead of <strong>data</strong> acquisition to allow for stabilization.<br />

The dipole wires and connect<strong>in</strong>g cables are weighted/covered with soil so as to<br />

m<strong>in</strong>imize w<strong>in</strong>d caused movement. Accord<strong>in</strong>g to Clarke et al. [1983], a 1 m length of wire<br />

vibrat<strong>in</strong>g at 1 Hz with amplitude of only 1 mm generates about 0.5 µV, a voltage that<br />

is comparable to the electric field signals at that frequency. Magnetometers are leveled<br />

<strong>us<strong>in</strong>g</strong> a spirit level and buried <strong>in</strong> pit (∼30 cm depth) with a cover of plastic sheet and soil<br />

to protect it from w<strong>in</strong>d, animals and temperature transients. Induction coil H z , requires<br />

a vertical hole of about 1 m depth. The <strong>data</strong> acquisition crew consist<strong>in</strong>g of six people,<br />

select and prepare the site dur<strong>in</strong>g daytime. The record<strong>in</strong>g usually starts once the sensors<br />

are stabilized and connections are properly made. More crewmembers are needed under<br />

difficult field conditions; for example if the area <strong>in</strong>accessible by vehicles.<br />

1.3.4 Calibration<br />

The natural electromagnetic fields, measured with a suite of sensors and record<strong>in</strong>g<br />

devices (system), are modified by the properties of the system itself. A proper knowledge<br />

of the measurement system’s response over the entire frequency range of <strong>in</strong>terest is<br />

essential <strong>in</strong> obta<strong>in</strong><strong>in</strong>g accurate magnetotelluric transfer functions. Any modern MT <strong>data</strong><br />

acquisition system employs a suite of analogue filters to precondition the signal prior<br />

to digitisation. The sensors, particularly the magnetometers, also respond differently


17<br />

<strong>in</strong> different frequency ranges. Even the cables that carry the signals have appreciable<br />

responses at higher frequencies. All these factors make the MT system’s response a<br />

complex non-l<strong>in</strong>ear function. The GMS05 system may be calibrated <strong>in</strong> two ways: 1) A<br />

theoretical calibration table could be computed with the knowledge of different <strong>data</strong> acquisition<br />

parameters, filter sett<strong>in</strong>g etc and 2) By send<strong>in</strong>g a Pseudo Random B<strong>in</strong>ary Signal<br />

(PRBS – basically white noise) to the magnetometers and electric field preamplifiers and<br />

by obta<strong>in</strong><strong>in</strong>g the ratio between the calibration signal and the system output <strong>in</strong> frequency<br />

doma<strong>in</strong>. In the first mode f<strong>in</strong>e effects caused by tolerances <strong>in</strong> the components are not<br />

accomodated. The accuracy that can be achieved with the model function is about ±10%<br />

or better for band 1 and ± 2% for band 2 to 5, whereas <strong>in</strong> the second mode accuracy<br />

better than +/- 0.5% <strong>in</strong> magnitude and +/- 1 ◦ <strong>in</strong> phase can be achieved. The procedures<br />

given by Ell<strong>in</strong>ghaus [1997] were used to calibrate the measured MT time series, for the<br />

present study.<br />

1.3.5 Data acquisition.<br />

Prior to <strong>data</strong> record<strong>in</strong>g, a thorough check of sensors and <strong>in</strong>struments is a good practice.<br />

At each site parameters such as noise levels, amplifier ga<strong>in</strong> and filter performance are<br />

checked usually <strong>us<strong>in</strong>g</strong> short calibration runs (§ 4.3.1). Dur<strong>in</strong>g <strong>data</strong> acquisition <strong>in</strong> South<br />

India, the electric field amplitudes were so high <strong>in</strong> the frequency range 8 Hz to 256 Hz<br />

(band 2) for few sites that the measurement were made with lowest ga<strong>in</strong> sett<strong>in</strong>gs. As<br />

many modern digital MT equipment, GMS05 also allows the storage of time series <strong>data</strong><br />

<strong>in</strong> addition to the computed MT parameters. The storage of time series was crucial <strong>in</strong><br />

obta<strong>in</strong><strong>in</strong>g better estimates of magnetotelluric impedance, as will be shown <strong>in</strong> the follow<strong>in</strong>g<br />

chapters. The quantity of <strong>data</strong> to be collected at each site depends on the requirements of<br />

the exploration program (§ 1.3.1). However, time required <strong>in</strong> collect<strong>in</strong>g the <strong>data</strong> depends<br />

on the prevail<strong>in</strong>g natural signal strength as well. Once the record<strong>in</strong>g beg<strong>in</strong>s, <strong>data</strong> quality<br />

is monitored cont<strong>in</strong>uously. Monitor<strong>in</strong>g consists of<br />

1)Manual Edit<strong>in</strong>g: Inspection of magnetic and electric time series for obvious outliers<br />

(See § 5.2 for more details regard<strong>in</strong>g noise patterns of time series). A cont<strong>in</strong>uous<br />

disturbance <strong>in</strong> one or more channels and dead / muted traces may be due to equipment<br />

malfunction<strong>in</strong>g (Figure 5.1).<br />

2)Coherency: This is a measure of the correlation between the E and H fields and<br />

predictability of the E fields from the H fields (§ 3.3.5). Coherency less than about 0.8<br />

generally <strong>in</strong>dicates problems associated with local noise sources associated with <strong>in</strong>strumentation<br />

or cultural effects (for e.g. see Figure 6.19). It can also <strong>in</strong>dicate localized<br />

natural sources such as lightn<strong>in</strong>g that do not conform to the MT assumption of spatially<br />

uniform source fields.<br />

3)Smoothness: Well-behaved MT curves will show smooth variations between estimated<br />

magnetotelluric transfer functions. Sudden offsets, rapid changes <strong>in</strong> curvature<br />

such as slope reversals across less than 0.5 log cycle of frequency and curve slopes greater<br />

than 45-degrees on log-frequency versus log-apparent resistivity are physically implausible<br />

and <strong>in</strong>dicate a “near field” source (§ 2.4). It is desirable to have low scatter, moderate<br />

curvature and well-jo<strong>in</strong>ed frequency-band curve segments.


Chapter 2<br />

Signal and Noise Sources for<br />

magnetotellurics<br />

18


19<br />

2.1 Introduction<br />

The def<strong>in</strong>ition for signal and noise vary from one geophysical method to another. In<br />

the context of magnetotellurics, Stodt (1983) def<strong>in</strong>ed signals as those components of<br />

measured electric and magnetic fields, which are determ<strong>in</strong>istically related through the<br />

transfer functions of a multi-<strong>in</strong>put multi-output l<strong>in</strong>ear system (§ 3.3.1.1). Noise <strong>in</strong> MT<br />

may be then def<strong>in</strong>ed as the additive components <strong>in</strong> the measured fields that are not<br />

related <strong>in</strong> such a way. This def<strong>in</strong>ition carries over to the frequency doma<strong>in</strong> as well as<br />

the Fourier transform is a l<strong>in</strong>ear operator (Stodt [1983]). From a different po<strong>in</strong>t of view<br />

Madden [1964] termed all unpredictable part of MT <strong>data</strong> as noise. One consequence of<br />

this def<strong>in</strong>ition is that, determ<strong>in</strong>istic noises such as 50 Hz (power l<strong>in</strong>e) harmonics will not<br />

be treated as noise, even though they will not qualify by Stodt [1983] criteria for signal.<br />

However, any unpredictable changes <strong>in</strong> amplitude and /or phase of a determ<strong>in</strong>istic noise<br />

will be treated as noise. A third way of def<strong>in</strong><strong>in</strong>g signal and noise <strong>in</strong> MT would be check<strong>in</strong>g<br />

the source field morphology. However, if the signal and noise components have overlaps <strong>in</strong><br />

time, frequency or spatial doma<strong>in</strong>s, their separation is im-perfect. MT signal process<strong>in</strong>g<br />

algorithms try to differentiate the signal and noise parts by transform<strong>in</strong>g the measured<br />

<strong>data</strong> to a suitable doma<strong>in</strong> and or with the help of additional <strong>in</strong>dependent measurements.<br />

The derivations made <strong>in</strong> § 1.1.4 for MT transfer functions assumes a planar source field.<br />

Any process (natural or man-made) that makes the source field <strong>in</strong>homogeneous may thus<br />

be treated as noise source and its effect <strong>in</strong> MT records as noise. One way to better<br />

understand the signal and noise <strong>in</strong> MT <strong>data</strong> is to study their sources. Many signal and<br />

noise sources have its own characteristics, which may be used to separate their effects<br />

on measured MT <strong>data</strong>. The first section of this chapter deals with the signal sources for<br />

MT. A discussion on various noise sources is given <strong>in</strong> the second part.<br />

2.2 Signal sources<br />

2.2.1 Introduction<br />

<strong>Magnetotelluric</strong> method uses naturally occurr<strong>in</strong>g electromagnetic field as source signal.<br />

The advantage of <strong>us<strong>in</strong>g</strong> natural field is to have unlimited power available throughout the<br />

frequency range of <strong>in</strong>terest (∼ 10– 4 to 10 4 [Hz]). This is particularly important, <strong>in</strong> the<br />

low frequencies (< 0.01 [Hz]), where a large power source and equipment set up would be<br />

needed otherwise to generate signals. This ensures the prob<strong>in</strong>g range of MT to 100s of<br />

kilometers. But to depend on natural source means spend<strong>in</strong>g sufficient time <strong>in</strong> the field<br />

to measure enough <strong>data</strong>. This is necessary as the signal level of natural electromagnetic<br />

waves is unpredictable and <strong>in</strong>terfered with man-made and other noises (§ 2.3) especially<br />

near the frequency range 10 −1 to 1 [Hz].<br />

Natural electromagnetic waves come from a variety of processes and from sources<br />

rang<strong>in</strong>g from Earth’s core to distant galaxies (Vozoff [1991]). With<strong>in</strong> the frequency range<br />

of <strong>in</strong>terest to magnetotellurics, only two sources are important. These are the atmosphere<br />

and magnetosphere. World wide lightn<strong>in</strong>g activities <strong>in</strong> the lower atmosphere are causes<br />

for MT signals from 1 Hz to 10 4 Hz, whereas below 1 Hz the fields orig<strong>in</strong>ate ma<strong>in</strong>ly <strong>in</strong><br />

the magnetosphere due to its <strong>in</strong>teraction with solar w<strong>in</strong>d and energy exchange between


20<br />

Figure 2.1: Earths horizontal magnetic field spectrum (Modified after Macnae et al.<br />

[1984]) The Black and gray bars <strong>in</strong>dicate the frequency range of measurement bands for<br />

GMS05. (see 1.1)<br />

particles and waves. Figure 2.1illustrates the components <strong>in</strong> the natural electromagnetic<br />

spectrum (modified after Macnae et al. [1984]) the amplitude and shape shown <strong>in</strong> the<br />

plot may vary with location and time. The bars <strong>in</strong>dicate the measurement bands used<br />

for the study (see also table 1.1).<br />

As can be seen from the Figure2.1, the amplitude of natural magnetic fields spectra<br />

varies by almost 5 decades with<strong>in</strong> the frequency range ∼10 −2 to 10 5 Hz. The relatively<br />

higher amplitude <strong>in</strong> the lower frequency ranges are due to geomagnetic pulsations. The<br />

signals from world wide thunder storm activity (sferics) is the ma<strong>in</strong> source of natural<br />

signal for frequency > 1 Hz. These two phenomena do not overlap <strong>in</strong> frequency and there<br />

exists ‘dead band’ around 1 Hz, with very low signal amplitude. We may def<strong>in</strong>e another<br />

dead band around 1KHz as well. However, <strong>in</strong> magnetotelluric context, the effect of first<br />

‘dead band’ has been recognized widely and <strong>in</strong> the present thesis, the term ‘dead band’<br />

refers to the low energy frequency band around 1 Hz. The sharp peaks <strong>in</strong> the spectra<br />

between 10 to 1000 Hz are due to power l<strong>in</strong>e <strong>in</strong>terference. This clearly shows how narrow<br />

band noise can override the natural signals. In the follow<strong>in</strong>g sections, an overview of the<br />

two sources for magnetotelluric signals are given.


21<br />

2.2.2 Geomagnetic pulsations<br />

Geomagnetic pulsations are temporal variations <strong>in</strong> the Earth’s magnetic field that have<br />

a quasi-periodic structure with frequencies rang<strong>in</strong>g from 10 −3 [Hz] to 2 [Hz] (Kaufman<br />

and Keller [1981]). The magnetosphere is the region around the Earth <strong>in</strong> which the ma<strong>in</strong><br />

magnetic field is conf<strong>in</strong>ed by solar w<strong>in</strong>d. Solar w<strong>in</strong>d consists of charged particles, mostly<br />

hydrogen and helium nuclei, ejected from sun and it take four days to reach Earth. Earth’s<br />

magnetosphere conta<strong>in</strong>s ionised oxygen and nitrogen, which make the magnetosphere<br />

conductive. The conductive part (100 to 250 km) of magnetosphere is called ionosphere<br />

and the resistive lower part is called atmosphere. This very complex system of Earth’s<br />

magnetosphere is constantly bombarded by solar w<strong>in</strong>d. The magnetosphere responds to<br />

magnetic pressure of solar w<strong>in</strong>d plasma by generat<strong>in</strong>g waves. The EM hydrodynamic<br />

waves thus set up travels towards the Earth. To reach Earth’s surface these waves<br />

must cross the ionosphere and resistive atmosphere. Ionosphere, with its anisotropy does<br />

not allow the vertical components of E and H . This not only modifies the horizontal<br />

components of the EM field, but set up horizontal currents <strong>in</strong> the ionosphere. It is believed<br />

that geomagnetic pulsation observed on surface of the Earth is ma<strong>in</strong>ly due to this current<br />

system <strong>in</strong> ionosphere. The time behavior of magnetic pulsations is episodic but <strong>in</strong>cludes<br />

features localized <strong>in</strong> frequency / space as a result of local conditions (Vozoff [1991]).<br />

Geomagnetic pulsations are divided <strong>in</strong>to two classes; cont<strong>in</strong>uous (Pc) and irregular (Pi).<br />

The cont<strong>in</strong>uous pulsations are subdivided <strong>in</strong>to six classes (2 – 0.001 [Hz]). They are Pc1<br />

(2 – 0.2 [Hz]), Pc2 (0.2 – 0.1[Hz]), Pc3 (0.1 to 0.022 [Hz] ), Pc4 (0.022 – 0.0066 [Hz]),<br />

Pc5 (0.0066 – 0.00166 [Hz]) and Pc6 with frequency less than 0.00166 [Hz]. The second<br />

class of pulsations, Pi, which have an irregular from, is divided <strong>in</strong>to three types: Pi-1 (1<br />

– 0.025 [Hz]), Pi-2 (0.025 – 0.0066 [Hz]) and Pi-3 with frequency less than 0.0066 [Hz]<br />

(Kaufman and Keller [1981]). As discussed earlier, due to the unpredictable occurrence<br />

character of the geomagnetic pulsation, long time of measurement is required to obta<strong>in</strong><br />

a rich spectrum of the signals <strong>in</strong> full bandwidth.<br />

To obta<strong>in</strong> signals up to 0.001 [Hz] it is usual to record for more than one day at a site.<br />

A suite of statistical methods is then employed to obta<strong>in</strong> smooth spectra <strong>in</strong> the frequency<br />

range of geomagnetic pulsations. Such an example from MT <strong>data</strong> collected at site VP16<br />

is presented here. The <strong>data</strong> was collected <strong>in</strong> period range 4sec to 128 sec, with a sampl<strong>in</strong>g<br />

rate 1 Hz. All the 5 MT components were measured for approximately 34 hours (25 - 27<br />

Feb.2000). Signal activity was strong <strong>in</strong> the range of Pc3 to Pc5 dur<strong>in</strong>g last half of the<br />

record<strong>in</strong>g session. Figure 2.2 shows a record for 4096 seconds of the measurement. The<br />

transient s<strong>in</strong>usoids, apparent <strong>in</strong> the records are geomagnetic pulsation and can easily be<br />

identified (see §5 ). The <strong>data</strong> was subjected to spectral <strong>analysis</strong> (§ 3.2.3). The total<br />

record was divided <strong>in</strong>to 30 segments of 4096 samples each. Each of the segments were<br />

Fourier transformed and an average spectral amplitude for H x is given <strong>in</strong> Figure 2.3. As<br />

a whole, the spectral power decreases monotonically with <strong>in</strong>creas<strong>in</strong>g frequency, much <strong>in</strong><br />

agreement with the natural EM spectra as given <strong>in</strong> Figure 2.1. Superimposed on the<br />

ma<strong>in</strong> trends, at least three lobes associated with Pc3, Pc4 and Pc5 are prom<strong>in</strong>ent.


22<br />

Figure 2.2: Plot for MT time series recorded at site VP16 show<strong>in</strong>g geomagnetic pulsations<br />

Figure 2.3: Averaged magnetic amplitude spectra (Hx) at station VP16. Peaks related<br />

to geomagnetic pulsations are labeled.Thick (gray) l<strong>in</strong>e represents smoothed spectra


23<br />

Figure 2.4: Plot of MT time series show<strong>in</strong>g impulse signals related to sferics recorded at<br />

site C12.<br />

2.2.3 Thunder storm activity<br />

<strong>Magnetotelluric</strong> signals with frequency above 1 Hz ma<strong>in</strong>ly orig<strong>in</strong>ate from worldwide thunderstorms.<br />

About 100 to 1000 lightn<strong>in</strong>g storms happen at any given moment worldwide.<br />

The fields as seen by the MT system depend on the strengths, path lengths (cloud heights)<br />

occurr<strong>in</strong>g frequencies etc (Vozoff [1991]). These signals, which are called Atmospherics<br />

or sferics, die off with distance. Most of these storms are located <strong>in</strong> the tropical region.<br />

Each lightn<strong>in</strong>g produces a current flow <strong>in</strong> the atmosphere with a peak <strong>in</strong>tensity of 300<br />

[Am]. The lightn<strong>in</strong>g excites the electric field <strong>in</strong> the resistive atmosphere that is sandwiched<br />

between relatively conductive ionosphere and Earth. While propagat<strong>in</strong>g from<br />

the light<strong>in</strong>g source, electromagnetic waves get reflected at lower and upper boundaries<br />

due to the high resistivity contrasts. They effectively travel round the glob 7.75 times a<br />

second (A circumference of 38,400 [km] with a speed of 297600 [km/s]). The atmosphere<br />

acts as a resonator and filters the lightn<strong>in</strong>g spike to multiples of resonant frequencies.<br />

These resonance frequencies are called ‘Schumann resonance’. In 1952 W.O. Schumann<br />

predicted (Meloy, scientific report available at http://space.t<strong>in</strong>.it/scienza/rromero) the<br />

existence of electromagnetic resonance <strong>in</strong> the Earth - Ionosphere cavity as,<br />

f n = 7.49 [n(n + 1)] 1 2<br />

for n = 1, 2, 3... (2.1)<br />

The equation yields f1 = 10.6 Hz, f2 = 18.4 Hz and so on. However, the first observational<br />

evidence of resonance came after the studies of Balser and Wagner [1960], when<br />

they measured the resonance of electromagnetic pulse generated from nuclear explosions.<br />

They found the resonance at f1 = 7.8 Hz, f2 = 14.2 Hz etc. The difference <strong>in</strong> values


24<br />

Figure 2.5: Electric Field (Ex) spectral amplitude depict<strong>in</strong>g the Schumann resonance<br />

between prediction and observation was attributed to the ionospheric loses. Sufficiently<br />

far from the light<strong>in</strong>g source, the electromagnetic fields behave like a plane wave - a requirement<br />

for MT (see § 1.1.1). A plot of MT time series measured at a sampl<strong>in</strong>g rate<br />

1024 Hz (Band 2) is given <strong>in</strong> Figure 2.4. An impulsive source and its decay are evident<br />

between <strong>data</strong> samples 100 – 150. When signal activity is strong, the observed MT<br />

<strong>data</strong> conta<strong>in</strong>s superposition of <strong>in</strong>dividual sferics orig<strong>in</strong>at<strong>in</strong>g from different thunderstorm,<br />

throughout the world. Schumann resonance frequencies become prom<strong>in</strong>ent <strong>in</strong> such cases<br />

and an example is given <strong>in</strong> Figure 2.5. The amplitude spectra plot for E x component<br />

of electric field clearly shows the resonance frequencies related to thunder storm activity<br />

especially at frequencies 8 Hz and 16 Hz.<br />

As discussed <strong>in</strong> § 2.2.1, the typical natural electromagnetic signals with frequencies<br />

around 1 kHz have very low signal strength (Figure 2.1) and are below the background<br />

noise from the measur<strong>in</strong>g <strong>in</strong>struments. After analyz<strong>in</strong>g the seasonal and diurnal behavior<br />

of a large set of high frequency MT <strong>data</strong> collected from Canada and northern Germany,<br />

Garcia and Jones [2002] conclude that, to better estimate high frequency MT transfer<br />

functions, the measurements should be made <strong>in</strong> the nighttime. Dur<strong>in</strong>g daytime, the<br />

atmospheric conductivity is larger, compared to night, as direct sun light ionizes the<br />

atmosphere. The <strong>in</strong>creased conductivity results <strong>in</strong> strong attenuation of sferics. The<br />

highest observable signal occurs when the whole propagation path from the lightn<strong>in</strong>g to<br />

the site is unlit.


25<br />

Active Passive Natural Instrument<br />

Electric power Conductors: Magnetic storms, Failure <strong>in</strong> electronics,<br />

transmission, power l<strong>in</strong>es, Lightn<strong>in</strong>g, Microseisms,<br />

Sensor<br />

Electrified Rail Pipel<strong>in</strong>es, Fences,<br />

W<strong>in</strong>d, malfunction<strong>in</strong>g,<br />

l<strong>in</strong>es, Electric Large iron structures<br />

Equatorial elec-<br />

Dropped bits, sen-<br />

switch<strong>in</strong>g, Factories,Vehicle<br />

Resistors: trojet.<br />

sor misalignment,<br />

Roads, ditches,<br />

muted traces.<br />

movement. culverts<br />

Table 2.1: Classification of noise sources that affect MT measurements.<br />

2.3 Noise Sources<br />

2.3.1 Introduction<br />

Compared to signal sources, the number of mechanisms that produce noise <strong>in</strong> the measured<br />

magnetotelluric <strong>data</strong> are numerous. This make it difficult to review the various<br />

noise sources for MT. Further the classifications of the noise sources are different <strong>in</strong><br />

different research documents. An attempt is made here to classify the noise sources for<br />

magnetotellurics, with<strong>in</strong> the limit of the scope of the thesis. Thus the terms ‘geologic’ and<br />

‘terra<strong>in</strong>’ noises are not described <strong>in</strong> detail. Ward [1967] divided noise <strong>in</strong> measured electromagnetic<br />

<strong>data</strong> <strong>in</strong>to <strong>in</strong>strumental, terra<strong>in</strong> & disturbance field. McCraken and Hohman<br />

[1986] classified EM noises <strong>in</strong>to ‘geologic’ and ‘electromagnetic’. While review<strong>in</strong>g man<br />

made electromagnetic noises <strong>in</strong> geophysics, Szarka [1988] classified them <strong>in</strong>to passive,<br />

active and other effects. Hatt<strong>in</strong>g [1989] classified them <strong>in</strong>to mechanical equipment and<br />

electromagnetic. A compilation of the noise sources is given <strong>in</strong> the table 2.1.<br />

The various sources listed <strong>in</strong> the table above describes the major noises that affect<br />

electromagnetic methods and magnetotelluric <strong>in</strong> particular. The list is by no means<br />

comprehensive. For a detailed review on noise sources see Szarka [1988], Junge [1996].<br />

2.3.2 Noises from power l<strong>in</strong>e signals<br />

Man made noise, <strong>in</strong> the EM spectrum (Figure 2.1) comes ma<strong>in</strong>ly from the electric power.<br />

Figure 2.6 shows the distribution of electromagnetic field strength of three phase power<br />

transmission l<strong>in</strong>e <strong>in</strong> the immediate vic<strong>in</strong>ity of power l<strong>in</strong>es (Szarka [1988]). The <strong>in</strong>terference<br />

from power l<strong>in</strong>e exponentially decreases with distance from it. Several authors studied<br />

EM harmonics <strong>in</strong> the vic<strong>in</strong>ity of 50 Hz power l<strong>in</strong>es (Szarka [1988]). This <strong>in</strong>terference is<br />

<strong>in</strong>ductive <strong>in</strong> nature and will not harm MT <strong>data</strong>, if the site is sufficiently far. However,<br />

there are other types of <strong>in</strong>terference that causes a current flow <strong>in</strong> the Earth. An ideal and<br />

perfectly balanced AC power transmission system does not cause any active EM noise <strong>in</strong><br />

the earth. In the case of unbalanced networks (ie <strong>in</strong> practice) a current component appears<br />

hav<strong>in</strong>g equal amplitudes and directions <strong>in</strong> each conductor. This is the so-called zerosequence<br />

current (Szarka [1988]) which usually flows through soil. The noisy <strong>in</strong>terference<br />

from this current may travel a great distance and affect MT <strong>data</strong>, depend<strong>in</strong>g on the<br />

resistivity of the soil. Motor loads operate non-synchronously and can also produce


26<br />

Figure 2.6: Distribution of electromagnetic field strength of three phase power transmission<br />

l<strong>in</strong>e <strong>in</strong> the immediate vic<strong>in</strong>ity of power l<strong>in</strong>es (adopted from Szarka [1988]). H total<br />

magnetic field, E horizontal electric field.<br />

side bands and sub-harmonics of the ma<strong>in</strong> frequencies. Added to this are the problems<br />

aris<strong>in</strong>g from f<strong>in</strong>ite observations. This results <strong>in</strong> ‘spectral leakage’ and can corrupt the<br />

transfer function estimate <strong>in</strong> the vic<strong>in</strong>ity of power l<strong>in</strong>e harmonics as well. Results of<br />

such an <strong>analysis</strong> is shown <strong>in</strong> Figure 2.7. The <strong>data</strong> were acquired at OK14, <strong>in</strong> band 2<br />

(sampl<strong>in</strong>g rate 1024) and was affected by noise from nearby power l<strong>in</strong>es. Power spectrum<br />

of electric field computed from three sets of <strong>data</strong> length viz. 128, 256 and 1024 are plotted.<br />

The dist<strong>in</strong>ctive peaks of the spectrum are related to ma<strong>in</strong> power l<strong>in</strong>e frequency and its<br />

harmonics. Leakage of power from ma<strong>in</strong> lobe is visible for spectrum calculated from<br />

128 and 256 <strong>data</strong> samples compared to the spectrum computed from 1024 <strong>data</strong> samples.<br />

This shows how the narrow band noise can leak <strong>in</strong>to their vic<strong>in</strong>ity. The transfer function<br />

estimations near to 50 Hz and its harmonic can get affected, if they are computed from<br />

a low-resolution spectrum.<br />

2.3.3 Noise from electric traction<br />

Electrified railways are another noise source for magnetotellurics. The Indian Railways<br />

adopted 25,000 V as standard for its electric traction <strong>in</strong> 1957. The electric traction is<br />

powered by AC power l<strong>in</strong>es, which hang above the tracks. The current loop is completed<br />

through the electric motors which are <strong>in</strong> turn connected to the wheels and thus to the<br />

rail tracks. Rail tracks are grounded at regular <strong>in</strong>tervals for safety reasons. A part of the<br />

current will travel through the ground to the electric substation. In this way, rail related<br />

noises have two components. While the <strong>in</strong>duction effect from the power l<strong>in</strong>es can be felt<br />

at distances less than few hundred meters, the electric impulse through ground can travel<br />

a great distance. In the context of South India, where the upper crust is highly resistive<br />

these noises can carry great distances from the source. Chaize and Lavergne [1970] give


27<br />

Figure 2.7: Amplitude spectra of electric field computed over different <strong>data</strong> length for<br />

the site OK14. Th<strong>in</strong> black l<strong>in</strong>e - 1024, Thick gray l<strong>in</strong>e 256 and Gray dots - 128. See text<br />

for discussion<br />

a brief description of source mechanism for noise <strong>in</strong>terference. As described <strong>in</strong> the Figure<br />

2.8 current I 2 hav<strong>in</strong>g an impulse like character due to the power enter<strong>in</strong>g the Earth<br />

where the rail track is grounded and it is this erroneous current that causes distant EM<br />

disturbances. Power for underground railways is usually supplied at the voltage of 1000<br />

v and its load<strong>in</strong>g may result <strong>in</strong> an impulse with an amplitude 7000 A (Szarka [1988]).<br />

2.3.4 Noises from <strong>in</strong>strument & sensors<br />

Noises from failure <strong>in</strong> electronics <strong>in</strong> equipment, quantization errors at A/D converter<br />

(Hatt<strong>in</strong>g [1989]), malfunction<strong>in</strong>g of sensors and errors <strong>in</strong> align<strong>in</strong>g the sensors are common<br />

<strong>in</strong> magnetotelluric <strong>data</strong>. In addition to this the sensors and the circuitry has its own<br />

background noise, which may vary with time as the electronic component develop fatigue.<br />

One way to identify such a problem is to calibrate the system (§ 1.3.5). Another way is to<br />

set up electric and magnetic sensors <strong>in</strong> parallel (Pedersen [1988]). The sensors should be<br />

placed sufficiently far away from each other that they do not <strong>in</strong>teract but sufficiently close<br />

that the signals are same. Then by averag<strong>in</strong>g over a number of frequency components we<br />

may def<strong>in</strong>e the coherence γ 2 between two sensor outputs. A correspond<strong>in</strong>g expression for<br />

the signal to noise ratio is obta<strong>in</strong>ed as S/N = γ 2 /(1-γ 2 ). A very low coherence <strong>in</strong>dicates<br />

malfunction<strong>in</strong>g of one of the sensors/circuitry. Disorientation of sensors can also <strong>in</strong>duce<br />

noise to the transfer function estimation. Pedersen [1988] showed that a 5 0 deviation <strong>in</strong><br />

magnetic sensor alignment would correspond to a skew of 0.1.


28<br />

Figure 2.8: Current circuit electric railway systems. I is the current <strong>in</strong> the overhead<br />

power l<strong>in</strong>e I1 is the return current <strong>in</strong> the rails and I2 is the return current <strong>in</strong> the Earth (<br />

modified after Chaize and Lavergne [1970]).<br />

2.3.5 Noises from the other sources<br />

Switch<strong>in</strong>g of submersible pumps causes disturbances <strong>in</strong> measured fields and is observed <strong>in</strong><br />

farm areas and villages. Electric channels are more prone to this type of noise. Movements<br />

of ferrous metals or other magnetic material <strong>in</strong> the vic<strong>in</strong>ity of the magnetic field sensor can<br />

<strong>in</strong>troduce noise <strong>in</strong>to the magnetic channels. Vehicular traffic generates both magnetic and<br />

seismic noise. In most cases, the magnetic effects are negligible, when the magnetometers<br />

are more than 200 m from the road (Clarke et al. [1983]). Seismic noise transforms <strong>in</strong>to<br />

magnetic noise through the movement of magnetometer <strong>in</strong> the Earth’s field. Clarke et al.<br />

[1983] reports that for the worst case, when the sensor is aligned perpendicular to earth’s<br />

ma<strong>in</strong> magnetic field, a rotation of 0.002 0 produced by seismic vibration produces a field<br />

change of 1 nT.<br />

Passive noise sources of EM measurement generally mean superficial resistivity <strong>in</strong>homogeneities<br />

of man made orig<strong>in</strong>. Conductive constructions (Pipel<strong>in</strong>es, metal fences<br />

etc.) may cause a redistribution of extreme natural electromagnetic phenomena such as<br />

magnetic storms, lightn<strong>in</strong>g etc. Passive distortion effects of man made construction can<br />

surely be avoided by choos<strong>in</strong>g the MT site away from them (§ 1.3.2). But this may not<br />

be possible always, especially <strong>in</strong> <strong>in</strong>dustrial areas (as discussed <strong>in</strong> § 4). The largest and<br />

most common source of natural noise is w<strong>in</strong>d, which can either move the magnetometer<br />

directly or <strong>in</strong>duce seismic noise by blow<strong>in</strong>g on trees or bushes whose roots then move the<br />

ground. Thus ideally the magnetometers should be <strong>in</strong> a flat area and buried away from<br />

trees. The cable connect<strong>in</strong>g the sensors to the <strong>data</strong> acquisition system should be secured<br />

to ground, to protect it from mov<strong>in</strong>g <strong>in</strong> w<strong>in</strong>d. A 1 meter length of wire vibrat<strong>in</strong>g at 1 z<br />

with an amplitude of only 1 mm generates about 0.5 µ V, a voltage that is comparable<br />

to the telluric signal at that frequency.


29<br />

Figure 2.9: Near and far field effect of a grounded dipole on MT measurements (adapted<br />

from Zonge and Hughes [1991]). The two panels on left and right describes the apparent<br />

resistivity and phase values for far field and near field measurements<br />

2.4 Effect of active electrical noise <strong>in</strong> magnetotelluric<br />

<strong>data</strong><br />

Most of the active noise sources discussed above manifest <strong>in</strong> MT measurements as correlated<br />

and un-correlated noise that can be attributed to the near field of a grounded<br />

dipole (Oett<strong>in</strong>ger et al. [2001]). Zonge and Hughes [1991] describe the electromagnetic<br />

field grounded <strong>in</strong> a homogeneous half space. Figure 2.9 gives a diagrammatic sketch of<br />

such a setup. Consider the near field and far field case. The near field case applies when<br />

the dipole – measurement site distance is far smaller than the sk<strong>in</strong> depth of the EM wave<br />

(r A


30<br />

Figure 2.10: Effect of a near field source on MT <strong>data</strong>, as recorded at the station OK14.<br />

See text for discussion<br />

One such example from the <strong>data</strong> collected at station OK14 is presented here. The<br />

station is located 80 km south of Erode <strong>in</strong> the measurement corridor (Figure 4.1). The<br />

<strong>data</strong> was collected <strong>in</strong> a visibly excellent site, but was found to be affected noise from<br />

unknown source. Time series collected <strong>in</strong> band 2 (sample rate 1024 Hz) were sub segmented<br />

<strong>in</strong>to stacks of 1024 <strong>data</strong> samples and were processed <strong>us<strong>in</strong>g</strong> least square technique<br />

(Sims et al. [1971]). The magnetotelluric transfer functions, especially the xy component<br />

showed the near source effect. The apparent resistivity and phase are plotted <strong>in</strong> Figure<br />

2.10 follows the exact description of Zonge and Hughes [1991] of near source effect. The<br />

error bars computed have dimension less than the symbols <strong>in</strong> the plot. In the far field<br />

(r C >> δ) the electric and magnetic field decay as 1/r 3 and the apparent resistivity and<br />

phase depends on frequency and resistivity (Figure 2.9). The ratio E/H is <strong>in</strong>dependent<br />

of the dipole – site distance r. The phase difference between E and H is 45 0 and apparent<br />

resistivity equals the resistivity of the homogeneous half space.<br />

The effect of near source noise <strong>in</strong> MT is one such example of how noise affects the<br />

computed MT parameters. In the presence of noise, it becomes necessary to apply statistical<br />

signal process<strong>in</strong>g tools to m<strong>in</strong>imize its effect on computed MT parameters. Many<br />

of the standard spectrum <strong>analysis</strong> techniques on random <strong>data</strong> have been applied to MT<br />

for this purpose. In the next chapter, an overview of various signal process<strong>in</strong>g methods<br />

that are commonly employed <strong>in</strong> magnetotelluric time series <strong>analysis</strong> is given.


Chapter 3<br />

<strong>Magnetotelluric</strong> time series <strong>analysis</strong><br />

31


32<br />

3.1 Introduction<br />

Various natural and anthropogenic processes contribut<strong>in</strong>g to magnetotelluric signal observed<br />

at Earth’s surface were described <strong>in</strong> the previous chapter. Assum<strong>in</strong>g a quasiuniform<br />

source field, the l<strong>in</strong>ear relation between measured electric and magnetic <strong>data</strong> may<br />

yield usable <strong>in</strong>formation about the distribution of electrical conductivity of earth. Unfortunately<br />

not all electromagnetic signals that are measured at Earth’s surface comprise<br />

the <strong>in</strong>duction process or <strong>in</strong>clude the <strong>in</strong>formation we seek. Data reduction and process<strong>in</strong>g<br />

techniques are necessary to convert the measured time series to an <strong>in</strong>terpretable form.<br />

Generally <strong>in</strong>terpretation of magnetotelluric <strong>data</strong> is done <strong>in</strong> frequency doma<strong>in</strong>. Therefore,<br />

the first step <strong>in</strong> <strong>in</strong>terpret<strong>in</strong>g magnetotelluric <strong>data</strong> <strong>in</strong>volves evaluat<strong>in</strong>g time series<br />

to identify and reject variations that seem to be of non-<strong>in</strong>ductive sources. The next is<br />

to estimate 10 1 – 10 2 frequency doma<strong>in</strong> complex transfer functions Z( ω) from the raw<br />

electric and magnetic field time series E(t) and H(t) with approximately 10 6 real numbers<br />

/ site (Egbert and Livelybrooks [1996]). MT <strong>data</strong> process<strong>in</strong>g is straight forward when<br />

there is no noise present <strong>in</strong> the measurements and the equations relat<strong>in</strong>g <strong>in</strong>duc<strong>in</strong>g and<br />

<strong>in</strong>duced variations through the impedance tensor (§ 2) are directly applicable. When<br />

noise is present a large number of process<strong>in</strong>g methods have been proposed (Sims et al.<br />

[1971], Goubau et al. [1978], Gamble et al. [1979], Stodt [1983]); Formerly MT time series<br />

<strong>analysis</strong> was done as an extension of classical random <strong>data</strong> <strong>analysis</strong> (Bendat and<br />

Piersol [1971]). Much of the same <strong>data</strong> reduction and transformation techniques are still<br />

followed. However, the last two decades have seen considerable improvement <strong>in</strong> the procedural<br />

and computational aspects of magnetotelluric method (Park and Chave [1984],<br />

Egbert and Booker [1986], Chave and Thomson [1989], Larsen [1989], Sutarno and Vozoff<br />

[1991], Egbert [1997], Banks [1998], Ritter et al. [1998], Oett<strong>in</strong>ger et al. [2001], Smirnov<br />

[2003], Chave and Thomson [2003], Manoj and Nagarajan [2003]).<br />

In this chapter an <strong>in</strong>troduction to various signal process<strong>in</strong>g steps that are used <strong>in</strong><br />

MT <strong>data</strong> <strong>analysis</strong> are described. Natural signals like magnetotelluric fields are usually<br />

treated as random process and statistical methods are applied to derive their properties.<br />

Basic concepts of random <strong>data</strong> and its properties are discussed <strong>in</strong> § 3.2.1. In the context<br />

of magnetotellurics it is also desirable to describe certa<strong>in</strong> jo<strong>in</strong>t properties of <strong>data</strong> from<br />

two or more random processes, namely components of horizontal electric and magnetic<br />

fields. The concept of jo<strong>in</strong>t probability distribution functions, cross correlation functions<br />

and cross-spectral density functions are described <strong>in</strong> § 3.2.2. Computational aspects of<br />

these functions from real <strong>data</strong> are described <strong>in</strong> § 3.2.3. Perhaps the dual <strong>in</strong>put – dual<br />

output l<strong>in</strong>ear system with additive noise <strong>in</strong> all components is the best way to describe the<br />

relation between magnetotelluric fields. The classical least square method of solv<strong>in</strong>g for<br />

the magnetotelluric transfer functions, its properties and associated errors are described<br />

towards the end of the chapter (§ 3.3). Violations to the assumption of noise free <strong>in</strong>put<br />

signals may bias the LS estimation. A brief discussion on certa<strong>in</strong> disadvantages of LS<br />

estimators is also <strong>in</strong>cluded <strong>in</strong> § 3.3.


33<br />

3.2 Random <strong>data</strong> <strong>analysis</strong><br />

When considered <strong>in</strong>dividually, both signal and noise processes <strong>in</strong> magnetotelluric field<br />

components can be considered as <strong>in</strong>dependent random processes. All determ<strong>in</strong>istic parts<br />

are noise and can easily be removed. Random process may be categorized as stationary<br />

and non stationary. Statistical properties of stationary random processes do not vary<br />

with time, where as these properties will change with time for non-stationary processes.<br />

3.2.1 Properties of random <strong>data</strong><br />

The ma<strong>in</strong> types of statistical functions used to represent a random process are (1) probability<br />

distribution function (2) mean square value and median (3) auto-correlation function<br />

and (4) power spectral density function.<br />

3.2.1.1 Probability density function<br />

Probability density function or pdf describes the probability that the <strong>data</strong> will assume<br />

certa<strong>in</strong> amplitude with<strong>in</strong> some def<strong>in</strong>ed range at any <strong>in</strong>stant of sampl<strong>in</strong>g. The probability<br />

that sample r(n) assumes a value between r and r+∆r may be obta<strong>in</strong>ed by tak<strong>in</strong>g the<br />

ratio N r /N where N r is the total number of occurrences that r(n) had <strong>in</strong> the range (r,<br />

r+∆r) from N <strong>in</strong>dependent field observations. In equation form,<br />

Pr ob[r < r(n) ≤ ∆r] =<br />

lim<br />

(3.1)<br />

N → ∞ N<br />

The probability that the current sample r(n) is less than or equal to some value r is<br />

def<strong>in</strong>ed by P(r), which is equal to the <strong>in</strong>tegral of the probability density function from<br />

m<strong>in</strong>us <strong>in</strong>f<strong>in</strong>ity to r. A typical plot of probability density versus <strong>in</strong>stantaneous values for<br />

magnetotelluric transfer function is presented <strong>in</strong> Figure 3.1 to demonstrate the property.<br />

The <strong>data</strong> recorded <strong>in</strong> band 3 (table 1.1) were sub-segmented to 174 blocks of 512 <strong>data</strong><br />

samples each. Transfer functions were estimated for each segments <strong>us<strong>in</strong>g</strong> least square<br />

technique (equation 3.37) , with m<strong>in</strong>imum 5 adjacent Fourier coefficients used for each<br />

estimate. The derived estimates were sorted <strong>in</strong>to 10 b<strong>in</strong>s of ranges. The ma<strong>in</strong> peak (near<br />

27.5 mV/(km.nT)) is flanked both sides by near symmetric decay.<br />

The bell shaped probability density plots <strong>in</strong>dicated <strong>in</strong> the Figure 3.1 are typical of<br />

either narrow or wide band random processes. These probability density plots would<br />

ideally be of the classical Gaussian form as given by the equation for variable x,<br />

N r<br />

p(x) = e−x2 /2σ x 2<br />

σ x<br />

√<br />

2π<br />

(3.2)<br />

Where σ is the standard deviation of x (§ 3.2.1.2). Gaussian or normal density function<br />

is most common <strong>in</strong> natural processes as is evident from MT measurements. Central limit<br />

theorem postulates that when a large number of processes contribute to s<strong>in</strong>gle random<br />

process, its pdf will tend to be a Gaussian, irrespective of probability density functions<br />

of <strong>in</strong>dividual process (Menke [1984]).


34<br />

Figure 3.1: Ordered —Zxy— plotted aga<strong>in</strong>st number of occurrences observed at station<br />

C13<br />

3.2.1.2 Mean square and variance<br />

The mean square value is simply the average of squared values <strong>in</strong> the time series. Consider<br />

the horizontal magnetic field h x (t), the mean square value ψ 2 of the time series is given<br />

as,<br />

Ψ 2 =<br />

lim 1<br />

T → ∞ T<br />

∫ T<br />

0<br />

h 2 x(t)dt (3.3)<br />

Many random processes have a static time <strong>in</strong>variant component and a dynamic or<br />

fluctuat<strong>in</strong>g component. The static component mean (µ) is the simple average of all the<br />

values and the fluctuat<strong>in</strong>g component variance is the mean square value about mean.<br />

Variance is given as,<br />

σ 2 =<br />

lim 1<br />

T → ∞ T<br />

∫ T<br />

0<br />

[h x (t) − µ] 2 dt (3.4)<br />

The positive square root of the variance is called standard deviation.<br />

3.2.1.3 Median and average absolute deviation.<br />

If a set of N observations are sorted <strong>in</strong>to the ascend<strong>in</strong>g order, h(1)≤h(2) ≤h(3)<br />

≤. . . .≤h(N), where h(j) is called j th order statistic. Median x is the middle sample


35<br />

is T is odd. The sample median is ambiguous for N even but it is typically chosen as<br />

(h [N/2] +h [N/2+1] )/2. Another class for the estimation of dynamic fluctuation is average<br />

absolute deviation and is given by<br />

σ =<br />

lim 1<br />

T → ∞ T<br />

∫ T<br />

0<br />

|h(t) − x|dt (3.5)<br />

Mean and median are various types of averages. It is less commonly realized that<br />

such averages are the result of m<strong>in</strong>imiz<strong>in</strong>g various norms. The mean µ is obta<strong>in</strong>ed by<br />

m<strong>in</strong>imiz<strong>in</strong>g the L 2 (or LS) norm or least squares norm (§ 3.3) of the samples. Whereas<br />

the median is obta<strong>in</strong>ed by m<strong>in</strong>imiz<strong>in</strong>g the L 1 norm of samples (Chave et al. [1987]).<br />

3.2.1.4 Autocorrelation function<br />

The autocorrelation functions of random <strong>data</strong> describes the dependence of the values of<br />

<strong>data</strong> at one <strong>in</strong>stance on the values at another time. In equation form,<br />

R x (τ) =<br />

lim 1<br />

N → ∞ N<br />

∫ T<br />

n=0<br />

h(n)h(n + τ)dn (3.6)<br />

where τ is the time lag. Autocorrelograms for a random process like magnetotelluric<br />

will be sharply peaked at lag τ = 0 and rapidly dim<strong>in</strong>ishes to zero on either side of the<br />

correlogram. In the limit<strong>in</strong>g case of hypothetical white noise, the autocorrelogram is a<br />

dirac delta function at zero lag (Bendat and Piersol [1971]).<br />

3.2.1.5 Power spectral density function<br />

Frequency <strong>analysis</strong> of random <strong>data</strong> essentially <strong>in</strong>volves the implementation of the Fourier<br />

transform. If F( ω) and f(t) are frequency and time doma<strong>in</strong> expression of a random <strong>data</strong><br />

its relation given by Fourier transform are,<br />

and<br />

f(t) = 1<br />

2π<br />

∫ ∞<br />

−∞<br />

F (ω)e iωt dω (3.7)<br />

F (ω) =<br />

∫ ∞<br />

f(t)e −iωt dt. (3.8)<br />

−∞<br />

The transform def<strong>in</strong>ed <strong>in</strong> the equations 3.7 & 3.8 are only valid for functions with<br />

f<strong>in</strong>ite energy and a different treatment is necessary for <strong>data</strong> which exist for all the time so<br />

that the total energy (Σf 2 (t) , t = -∞. . . .∞) is not f<strong>in</strong>ite. In such cases one must consider<br />

the spectral properties of the energy density rather than the spectral properties of the<br />

amplitude fluctuations. Power spectral density function of a random <strong>data</strong> describes the<br />

frequency composition of the <strong>data</strong> <strong>in</strong> terms of the spectral density of its mean square


36<br />

value. Its concept is well understood by consider<strong>in</strong>g a narrow band analog filter. The<br />

mean squared out put of the filter will converge on average over a long period of time.<br />

In equation form,<br />

Ψ 2 x(f, ∆f) =<br />

lim 1<br />

T → ∞ T<br />

∫ T<br />

0<br />

h 2 (t, f, ∆f)dt (3.9)<br />

where T is the length of the record, H(t)(Bendat and Piersol [1971]). For small ∆f a<br />

power spectral density function S xx (f) can be def<strong>in</strong>ed such that,<br />

Ψ 2 x(f, ∆f) ≈ Sxx(f)∆f (3.10)<br />

Accord<strong>in</strong>g to so-called Wiener-Kh<strong>in</strong>ch<strong>in</strong> theorem (Ghil et al. [2002]) the power spectral<br />

density is equal to the Fourier transform of the autocorrelation function. Hence power<br />

spectrum of H(t) can be estimated by the Fourier transform S hh (f) of an estimate of<br />

autocorrelation function R x ( τ).<br />

S x (f) = 2<br />

∫ ∞<br />

R x (τ)e −j2πfτ dτ (3.11)<br />

The random part of the error <strong>in</strong> the estimation of S xx<br />

error,<br />

−∞<br />

is given by the normalized<br />

ɛ = √ 2/dof, (3.12)<br />

where, dof is the numbers of degree of freedom. As dof = 2 for each Fourier harmonics,<br />

ɛ >1 imply<strong>in</strong>g that the standard deviation of the estimate S xx is greater than the estimate<br />

itself. There for the need of spectral smooth<strong>in</strong>g is seen. Spectral smooth<strong>in</strong>g is discussed<br />

while address<strong>in</strong>g the computational aspects of power spectral density <strong>in</strong> § 3.2.3.<br />

3.2.2 Jo<strong>in</strong>t signal properties<br />

In many applications such as magnetotellurics, it is desirable to study the jo<strong>in</strong>t signal<br />

properties of two or more random processes. Cross-spectral density between Earth’s<br />

natural electric and magnetic field gives the relation between the fields <strong>in</strong> amplitude and<br />

phase, which <strong>in</strong> turn is useful to understand the subsurface conductivity distribution (§<br />

1.1). The jo<strong>in</strong>t statistical properties like jo<strong>in</strong>t probability distribution, cross correlation,<br />

cross-spectral density etc are essentially the extensions of § 3.2.1.<br />

3.2.2.1 Cross correlation functions<br />

Cross correlation function of two sets of random <strong>data</strong> describes the general dependence<br />

of the values of one set of <strong>data</strong> on the other (Bendat and Piersol [1971]). Consider two


37<br />

Figure 3.2: Crosscorrelogram of two magnetotelluric time series components.<br />

random series H(t) and E(t), their cross correlation is given as<br />

R hy (τ) =<br />

lim 1<br />

T → ∞ T<br />

∫ T<br />

0<br />

h(t)y(t + τ)dt (3.13)<br />

The function R hy (τ) is always a real valued function which is symmetrical about the<br />

ord<strong>in</strong>ate when h and y are <strong>in</strong>terchanged. That is,<br />

R hy (−τ) = R yh (τ) (3.14)<br />

A Typical plot of the cross correlation versus time lag plot for two random processes<br />

(Ex and Hy) is given <strong>in</strong> Figure 3.2. Note the sharp peak at lag = 0 second and few other<br />

less def<strong>in</strong>ed peaks which shows the existence of correlation between two series Ex(t) and<br />

Hy(t) at specific displacements. The value of cross correlogram at lag=0 may be used<br />

to measure the similarity between two processes. For example, <strong>in</strong> § 5, correlation of<br />

orthogonal electric and magnetic field components were used to as a quality parameter<br />

of that section of time series.<br />

3.2.2.2 Cross-spectral density functions<br />

The cross spectral density (csd) function of two time series can be def<strong>in</strong>ed as the Fourier<br />

transform of their cross correlation function, <strong>in</strong> the same manner we def<strong>in</strong>ed power spectral<br />

density of a random process. As the cross correlation is an odd function, csd is a<br />

complex valued function unlike psd.


38<br />

3.2.2.3 Coherence<br />

When consider<strong>in</strong>g physical measurement of two random processes, it is often desirable<br />

to compute a real valued function γ 2 (f ), called ord<strong>in</strong>ary coherence function, which is a<br />

measure of l<strong>in</strong>ear relation between the processes as a function of frequency and frequency<br />

doma<strong>in</strong> equivalent of cross correlation. It may def<strong>in</strong>ed from the psd and csd of the<br />

processes as,<br />

S<br />

γhy(f) 2 =<br />

hy 2 (f)<br />

∣S hh (f)S yy (f) ∣ (3.15)<br />

The coherence function satisfies the limit 0


39<br />

3.2.3.3 W<strong>in</strong>dow<strong>in</strong>g<br />

F<strong>in</strong>ite observation of an <strong>in</strong>f<strong>in</strong>ite process can be viewed as multiply<strong>in</strong>g the <strong>in</strong>f<strong>in</strong>ite process<br />

with a boxcar function. Boxcar function has a value one with<strong>in</strong> the observation period and<br />

zero outside the observation period. Then the estimated power spectral density function<br />

S xx (f ) is the convolution of the true power spectral density with the frequency doma<strong>in</strong><br />

transform of the boxcar function. Thus convolution causes spectral energy to leak from<br />

the orig<strong>in</strong>al frequency to adjacent frequencies by add<strong>in</strong>g <strong>in</strong>f<strong>in</strong>ite number of smaller side<br />

lobes. Near power l<strong>in</strong>e frequency and its harmonics as observed <strong>in</strong> magnetotelluric time<br />

series, the leakage may totally corrupt the adjacent frequencies as discussed <strong>in</strong> § 2.3.2<br />

(see Figure 2.7). Many alternate w<strong>in</strong>dow functions were <strong>in</strong>troduced to obta<strong>in</strong> smooth<br />

spectra (Bendat and Piersol [1971]). A simple ‘Hann<strong>in</strong>g’ w<strong>in</strong>dow can be realized as,<br />

W (i) = 0.5<br />

(<br />

1 + cos<br />

( πi<br />

N<br />

))<br />

(3.17)<br />

Where I = 0,1,2,3. . . ..N samples of time series.<br />

As the w<strong>in</strong>dow function effectively suppresses the time series <strong>data</strong> at both ends,<br />

w<strong>in</strong>dow<strong>in</strong>g results <strong>in</strong> loss of <strong>in</strong>formation. In order to retrieve the <strong>data</strong> that are lost due to<br />

w<strong>in</strong>dow<strong>in</strong>g, it is a common practice to overlap, adjacent w<strong>in</strong>dows. In the magnetotelluric<br />

time series <strong>analysis</strong> carried out <strong>in</strong> this thesis Hann<strong>in</strong>g w<strong>in</strong>dow with 50 % overlap was<br />

used. However, the overlapp<strong>in</strong>g will decrease the effective degrees of freedom (dof) and<br />

corrections are given as,<br />

( 1<br />

dof = dof ·<br />

2 + 1 )<br />

(3.18)<br />

π<br />

3.2.3.4 Discrete Fourier Transform<br />

Discrete Fourier Transform (DFT) is obta<strong>in</strong>ed by frequency doma<strong>in</strong> sampl<strong>in</strong>g of the<br />

cont<strong>in</strong>uous Fourier Transform presented <strong>in</strong> equations 3.7 and 3.8. DFT realized on N<br />

discrete samples of <strong>data</strong> H(t) is given by,<br />

F DF T (f k ) = 1 T<br />

∑T −1<br />

kn<br />

−2πi<br />

W (t)h(t)e T (3.19)<br />

t=0<br />

With a spectral resolution ∆f=f s /N. Here k = 1,2,. . . T-1 and f s, is the sampl<strong>in</strong>g<br />

frequency. W(t) is the w<strong>in</strong>dow<strong>in</strong>g function given <strong>in</strong> equation 3.17. This summarization<br />

does not change the units the raw DFT should be divided by the sampl<strong>in</strong>g frequency,<br />

F (ω) ⇔ F DF T (ω k )/f s (3.20)<br />

to obta<strong>in</strong> the same result as <strong>in</strong> equation 3.8. Thus H(t) measured <strong>in</strong> the unit V will<br />

get the V/Hz after the transformation. To remove the effects of w<strong>in</strong>dow<strong>in</strong>g, the DFT<br />

has to be multiplied with <strong>in</strong>verse of the <strong>in</strong>tegral of the w<strong>in</strong>dow<strong>in</strong>g function, here Hann<strong>in</strong>g<br />

w<strong>in</strong>dows with an <strong>in</strong>tegral value 2. Power spectral density S xx (f k ) of time series H(t) is


40<br />

Figure 3.3: The Parzen spectral w<strong>in</strong>dow over the target frequency plotted on background<br />

of a sample spectra<br />

thus computed by,<br />

S xx (f k ) = 2<br />

T f s<br />

|F DF T (f k )| 2 (3.21)<br />

with the units V 2 /Hz. With the same reason, the amplitude spectrum get the unit<br />

V/ √ Hz.<br />

3.2.3.5 Smooth<strong>in</strong>g of spectra by band and section averag<strong>in</strong>g<br />

The estimate S xx (f) itself is a random function and from a loose def<strong>in</strong>ition of central limit<br />

theorem, it can be said that the S xx (f) will have a Gaussian distribution even though<br />

the time series may have different distribution. It is thus usual to attribute 2 degrees<br />

of freedom to each spectral density values (except for first (mean) and last (nyquist)<br />

harmonics, which have only 1 dof each). From equation 3.12 it is seen that the power<br />

spectrum estimate is <strong>in</strong>consistent, as the estimate is less than its standard deviation. In<br />

order to reduce bias and variance <strong>in</strong> estimated power spectrum, a number of techniques<br />

are proposed (Jenk<strong>in</strong>s and Watts [1968], Ghil et al. [2002]). A comb<strong>in</strong>ation of frequency<br />

band smooth<strong>in</strong>g and segment averag<strong>in</strong>g is generally followed <strong>in</strong> magnetotellurics to reduce<br />

variance <strong>in</strong> psd estimations (Sims et al. [1971], Chave et al. [1987]). As the MT transfer<br />

functions are slowly vary<strong>in</strong>g functions of frequency, the frequency band smooth<strong>in</strong>g is<br />

performed over a w<strong>in</strong>dow <strong>in</strong> frequency doma<strong>in</strong>, here the Parzen (Jenk<strong>in</strong>s and Watts<br />

[1968]) w<strong>in</strong>dow,<br />

⎧<br />

⎨<br />

W (f) =<br />

⎩<br />

1 if f z − f = 0<br />

( s<strong>in</strong>(u) ) 4 if 0 < f<br />

u z − f < f r<br />

(3.22)<br />

0 if f z − f ≥ f r<br />

where f z is the target frequency and f r is the radius of the Parzen w<strong>in</strong>dow. A<br />

schematic diagram is presented <strong>in</strong> Figure 3.3 show<strong>in</strong>g band averag<strong>in</strong>g over Parzen w<strong>in</strong>dow.


41<br />

Jenk<strong>in</strong>s and Watts [1968] discuss the use of other types of ‘spectral w<strong>in</strong>dows’ to reduce<br />

the variance <strong>in</strong> spectral estimation. They have shown that given a particular radius M,<br />

Parzen w<strong>in</strong>dow achieves the smallest variance among other types of w<strong>in</strong>dows.<br />

However, frequency smooth<strong>in</strong>g reduces spectral resolution. A trade off between frequency<br />

resolution and estimated variance dictates the choice of Parzen radius. Another<br />

problem with band averag<strong>in</strong>g is the bias error. Parzen w<strong>in</strong>dow like other spectral w<strong>in</strong>dows,<br />

may result <strong>in</strong> biased estimates of cross and auto spectrum and can affect the<br />

computed MT transfer functions. As will be shown <strong>in</strong> § 6.3, robust weight<strong>in</strong>g of spectra<br />

with<strong>in</strong> a band can vastly improve the estimates of auto and cross spectra and thereby<br />

yield less biased MT transfer functions. Section averag<strong>in</strong>g, which is def<strong>in</strong>ed as averag<strong>in</strong>g<br />

of spectra computed from the sub- segments of time series, may be applied to further<br />

smooth the spectra. Also called Bartlett’s smooth<strong>in</strong>g procedure, this will also result <strong>in</strong><br />

decrease <strong>in</strong> variances associated with the estimate of spectra (Jenk<strong>in</strong>s and Watts [1968]).<br />

Cross-spectral densities are also computed <strong>in</strong> the same fashion as for psd as described<br />

above. Frequency band averag<strong>in</strong>g assumes the <strong>in</strong>dividual spectral values have a Gaussian<br />

distribution (§ 3.2.1.1). Whereas the problem with non-Gaussian distribution <strong>in</strong> section<br />

averag<strong>in</strong>g is well recognized, the same for band averag<strong>in</strong>g has been overlooked. It will be<br />

shown <strong>in</strong> § 6.3 that a better estimate of cross and auto spectral densities can be obta<strong>in</strong>ed<br />

by robust band averag<strong>in</strong>g.<br />

3.3 <strong>Magnetotelluric</strong> transfer functions<br />

3.3.1 Least Square Solution<br />

3.3.1.1 Multi <strong>in</strong>put, multi output l<strong>in</strong>ear system<br />

<strong>Magnetotelluric</strong> <strong>data</strong> can be considered as a result of a system of random process, with<br />

the signal components determ<strong>in</strong>istically related between the <strong>in</strong>puts and outputs. Under<br />

the assumption of a plane source field, concepts appropriate for multiple-<strong>in</strong>put / multiple<br />

out put l<strong>in</strong>ear system can be applied (Jones et al. [1989]). The estimation of weight<strong>in</strong>g<br />

response functions, or their frequency doma<strong>in</strong> equivalent transfer functions for such a system<br />

is described <strong>in</strong> this section. Conventionally, the horizontal components of magnetic<br />

field Hx(t) and Hy(y) [nT] are treated as the <strong>in</strong>puts, and the horizontal components of<br />

electric field Ex(t) and Ey(t) [mV/km] are treated as outputs (Figure 3.4. The <strong>in</strong>puts<br />

and outputs are related by a convolution operation to the four time doma<strong>in</strong> weight<strong>in</strong>g<br />

functions Z( τ). The magnetotelluric <strong>data</strong> can be <strong>in</strong>terpreted <strong>in</strong> terms of its weight<strong>in</strong>g<br />

response functions Z( τ) (Kunetz [1972], Yee et al. [1988]) or its frequency doma<strong>in</strong><br />

representation, Z( ω) (see Vozoff [1991]). Many authors questioned the time doma<strong>in</strong> approach<br />

due to unstable statistical properties of the weight<strong>in</strong>g response functions (Jenk<strong>in</strong>s<br />

and Watts [1968], Egbert [1992]). With some exceptions the transfer functions are now<br />

computed <strong>in</strong> the frequency doma<strong>in</strong>.


42<br />

Figure 3.4: <strong>Magnetotelluric</strong> l<strong>in</strong>ear system with two horizontal magnetic components as<br />

<strong>in</strong>puts and horizontal electric components as outputs. Adapted from Jones et al. [1989].<br />

3.3.1.2 Solution with noise free <strong>data</strong><br />

In frequency doma<strong>in</strong> the complex frequency dependent relation between the horizontal<br />

MT fields can be written as given by Swift [1986],<br />

[ ] [ ] [ ]<br />

Ex Zxx Z<br />

=<br />

xy Hx<br />

(3.23)<br />

E y Z yy H y<br />

Z yx<br />

Where Z is the MT transfer function tensor (impedance tensor) [mV.km −1 .nT −1 ],<br />

E [mV.km −1 .Hz − 1 2 ] and H [nT.Hz − 1 2 ] are Fourier transforms ( of electric and magnetic<br />

fields E(t) and H(t) respectively at a particular frequency ω <strong>in</strong> radian. The relationship<br />

between vertical (H z ) and horizontal magnetic field can also be expressed <strong>in</strong> the same<br />

way as,<br />

H z = [ ] [ ]<br />

H<br />

T x T x<br />

y , (3.24)<br />

H y<br />

Where T x and T y are called tipper functions (Vozoff [1991]). As the solution for<br />

tipper functions are similar to that of impedance functions, they are not separately dealt<br />

with <strong>in</strong> the follow<strong>in</strong>g sections. When the measurements are noise-free, two <strong>in</strong>dependent<br />

observations of field components E and H are enough to estimate the transfer functions<br />

(Sims et al. [1971]). For example a solution for one tensor element is obta<strong>in</strong>ed as,<br />

∣ H ∣<br />

x1 E x1 ∣∣∣<br />

H x2 E x2<br />

Z xy =<br />

∣ H ∣<br />

x1 H y1 ∣∣∣<br />

(3.25)<br />

H x2 H y2<br />

where the subscript 1 and 2 <strong>in</strong>dicate <strong>data</strong> from two measurements. An additional


43<br />

requirement is that H x1 H y2 ≠ H x2 H y1 , which physically mean the two field measurements<br />

must have different source polarization (Kaufman and Keller [1981]).<br />

3.3.1.3 Solution with noise <strong>in</strong> measurement<br />

In the measured magnetotelluric <strong>data</strong>, we have no knowledge of the true signal components<br />

<strong>in</strong> Ex, Ey, Hx, Hy or Hz and it is desirable to make more than two <strong>in</strong>dependent<br />

observations for each components. With more than two measurements equation 3.23 is<br />

over determ<strong>in</strong>ed and a solution for transfer functions is sought which m<strong>in</strong>imizes the errors<br />

<strong>in</strong> observations. An equivalent matrix form replaces a component of the tensor equation<br />

3.22 as,<br />

E = ZH + r (3.26)<br />

where there are N observations so that E and r are N vectors, H is an N X 2 matrix<br />

and Z is a two vector. The last variable <strong>in</strong> equation (3.26), r is the difference between<br />

measured and predicted output field (here, Electric) and is the residual parameter to be<br />

m<strong>in</strong>imized. Classical way of solv<strong>in</strong>g MT equation is by the least squares technique (Swift<br />

[1986], Sims et al. [1971]) where <strong>in</strong> the square of the residual power <strong>in</strong> equation (3.26) is<br />

m<strong>in</strong>imized to yield a solution for Z. Writ<strong>in</strong>g one component of equation (3.26),<br />

E xi = Z xx H xi + Z xy H yi + r i (3.27)<br />

Squar<strong>in</strong>g this equation and summ<strong>in</strong>g for all the N records sets we obta<strong>in</strong>,<br />

r 2 =<br />

N∑<br />

(E xi − Z xx H xi − Z xy H yi ).(Exi ∗ − ZxxH ∗ xi ∗ − ZxyH ∗ yi) ∗ (3.28)<br />

i=1<br />

(Asterisk <strong>in</strong>dicates complex conjugate). Sett<strong>in</strong>g the derivative of r 2 with respect to<br />

Z xx and Z xy <strong>in</strong> turn to zero yield two <strong>in</strong>dependent equations,<br />

< E x H ∗ y > = Z xx < H x H ∗ y > + Z xy < H y H ∗ y > (3.29)<br />

< E x H ∗ y > = Z xx < H x H ∗ y > + Z xy < H y H ∗ y > (3.30)<br />

If we m<strong>in</strong>imize noise <strong>in</strong> magnetic field <strong>in</strong> equation (3.27) two more equations arise.<br />

They are,<br />

< E x Ex ∗ > = Z xx < H x Ex ∗ > +Z xy < H y Ex ∗ ><br />

< E x Ey ∗ > = Z xx < H x Ey ∗ > +Z xy < H y Ey ∗ ><br />

(3.31)<br />

where the quantities like H x H x and H y H y are power spectral densities (or auto<br />

spectra) and quantities E x H x and H x H y are cross-spectral densities (or cross spectra<br />

- § 3.2.2.2) between different field components. The bars <strong>in</strong>dicate segment averag<strong>in</strong>g<br />

and braces () <strong>in</strong>dicate an average over a small frequency band (§ 3.2.3.5). From now<br />

onwards the use of these symbols are omitted for simplicity. Any of the two equations<br />

(3.28 through 3.29) must be solved simultaneously for Z xx and Z xy . Thus there are six


44<br />

possible estimates for each of the tensor elements. For example, estimates of Z xy<br />

obta<strong>in</strong>ed as,<br />

are<br />

Ẑ xy = H xE ∗ xE x E ∗ y − H x E ∗ yE x E ∗ x<br />

H x E ∗ xH y E ∗ y − H x E ∗ yH y E ∗ x<br />

Ẑ xy = H xE ∗ xE x H ∗ x − H x H ∗ xE x E ∗ x<br />

H x E ∗ xH y H ∗ x − H x H ∗ xH y E ∗ x<br />

Ẑ xy = H xE ∗ xE x H ∗ y − H x H ∗ yE x E ∗ x<br />

H x E ∗ xH y H ∗ y − H x H ∗ yH y E ∗ x<br />

Ẑ xy = H xE ∗ yE x H ∗ x − H x H ∗ xE x E ∗ y<br />

H x E ∗ yH y H ∗ x − H x H ∗ xH y E ∗ y<br />

Ẑ xy = H xE ∗ yE x H ∗ y − H x H ∗ yE x E ∗ y<br />

H x E ∗ yH y H ∗ y − H x H ∗ yH y E ∗ y<br />

Ẑ xy = H xH ∗ xE x H ∗ y − H x H ∗ yE x H ∗ x<br />

H x H ∗ xH y H ∗ y − H x H ∗ yH y H ∗ x<br />

(3.32)<br />

(3.33)<br />

(3.34)<br />

(3.35)<br />

(3.36)<br />

(3.37)<br />

Where the hat corresponds to an estimate. Equation (3.32) is the solution for Z xy<br />

<strong>in</strong> the equations 3.31, which assumes that all noise is <strong>in</strong> magnetic fields. Two of the<br />

equations (3.34 and 3.35) are relatively unstable <strong>in</strong> the 1-D case where the fields are<br />

unpolarized. Any noise <strong>in</strong> the electrical field will cause the estimate for Z xy to bias<br />

upward. Equation 3.37 is the solution to set of equations 3.29 & 3.30, which treat the<br />

electrical field as output, (as <strong>in</strong> Figure 3.1). Noises <strong>in</strong> magnetic field components will<br />

down bias this estimate. However, MT transfer functions are rarely estimated <strong>us<strong>in</strong>g</strong><br />

equation (3.32), as i) The magnetic measurements are made with better accuracy than<br />

electric field and ii) Electric fields can get polarized due to electrical resistivity anisotropy<br />

and denom<strong>in</strong>ator of equation 3.32 can become very small or zero.<br />

3.3.2 Concept of bias<br />

A crucial factor <strong>in</strong> the Least Square method described above is the assumption of noise<br />

– free <strong>data</strong> <strong>in</strong> either the electric or magnetic <strong>data</strong>. In the equations for the estimations<br />

Z xy (eq 3.30), this assumption is implicit <strong>in</strong> use of auto spectra. Under the assumption<br />

that noise is uncorrelated with signal, A S B N ∗ =0, where A, B ∈ {E x , E y , H x , H y }, A S<br />

is the signal <strong>in</strong> component A and A N is the noise component <strong>in</strong> A, one can write,<br />

{<br />

AB ∗ = (A S + A N )(B S∗ + B N∗ ) =<br />

A S B S∗ when A ≠ B<br />

A S B S∗ + A N B N ∗ when A = B<br />

(3.38)<br />

While estimates for the cross spectra are statistically distributed and yield a good<br />

approximation to the true value, the estimates of the auto spectra are systematically too<br />

large (Müller [2000]). Thus noise <strong>in</strong> E x E ∗ x cause Z xy <strong>in</strong> equation (3.32) to bias upwards,<br />

where as noise <strong>in</strong> H x H ∗ x and H y H ∗ y <strong>in</strong> the denom<strong>in</strong>ator of equation 3.37 biases Z xy


45<br />

downwards. The downward and upward biased estimates give an envelope with<strong>in</strong> which<br />

true transfer functions should lie (Jones et al. [1989]). To reduce the bias effect Sims et al.<br />

[1971] proposed to take averages over four stable estimates, the argument be<strong>in</strong>g that the<br />

negative and positive bias would cancel each other. Kao and Rank<strong>in</strong> [1977] devised an<br />

iterative approach to remove the bias<strong>in</strong>g effects. Goubau et al. [1978] and Gamble et al.<br />

[1979] used magnetic <strong>data</strong> from a remote station (§ 3.3.6) <strong>in</strong> order to avoid the use of<br />

auto powers <strong>in</strong> solutions such as equations 3.32 to 3.37.<br />

3.3.2.1 Predicted coherence<br />

The coherence between measured and predicted output <strong>data</strong> describes the optimality of<br />

LS estimate such as given <strong>in</strong> equations 3.32 to 3.37. For example, squared predicted<br />

coherence, when consider Ex as output accord<strong>in</strong>g to equation (3.37) is given by,<br />

∣ ∣∣∣∣<br />

γ 2 E pEx p = |E x Ex| p 2<br />

(E x Ex)(E ∗ xE p p∗<br />

(3.39)<br />

∣<br />

where,<br />

x )<br />

E p x = ẐxxH x + ẐxyH y (3.40)<br />

(modified after Swift [1986]). Predicted coherence (or multiple coherence) functions<br />

will strongly <strong>in</strong>dicate the presence or absence of l<strong>in</strong>ear relationship between <strong>in</strong>put and<br />

output. This should be differentiated from the ord<strong>in</strong>ary coherency def<strong>in</strong>ed <strong>in</strong> equation<br />

3.15 (§ 3.2.2.3). A value of 1.0 <strong>in</strong>dicates the derived impedances are <strong>in</strong> perfect l<strong>in</strong>e with<br />

the observations of E and H. Presence of uncorrelated noise <strong>in</strong> the fields reduces the<br />

predicted coherence from unity. The three conditions under which the coherences can<br />

have non-unity values (Jones [1981]) are, when,:<br />

1. noise is present <strong>in</strong> both electric and magnetic fields<br />

2. the system relat<strong>in</strong>g the magnetic and electric field are non l<strong>in</strong>ear<br />

3. processes other than electromagnetic <strong>in</strong>duction are <strong>in</strong>volved<br />

Many approaches were made to weight MT spectra sets from different observations,<br />

accord<strong>in</strong>g to predicted coherence (Stodt [1983], Egbert and Livelybrooks [1996], Larsen<br />

et al. [1996]). However, as po<strong>in</strong>ted out by Dekker and Hastie [1981], the multiple coherence<br />

functions can get biased upwards when computed from fewer number of observations<br />

and/or <strong>in</strong> the presence of correlated noise between measured fields. Correction to bias<br />

error <strong>in</strong> coherence is discussed <strong>in</strong> § 4.6.<br />

3.3.3 Coherent and <strong>in</strong>coherent noises<br />

Noises <strong>in</strong> MT <strong>data</strong> can be classified <strong>in</strong>to coherent and <strong>in</strong>-coherent. Noises <strong>in</strong> electric or<br />

magnetic channels (or both) that cannot be related via transfer functions (out of l<strong>in</strong>e)<br />

are termed <strong>in</strong>coherent. An example is the thermal noises <strong>in</strong> sensors. Predictive coherence


46<br />

may be used to detect them. Coherent noises are noises that affect both the <strong>in</strong>put (say<br />

magnetic) and out put (say electrical) channels <strong>in</strong> l<strong>in</strong>e with the l<strong>in</strong>ear model (impedance).<br />

An example is a correlated spike. It may be detected by exam<strong>in</strong><strong>in</strong>g the time variations<br />

of the impedance functions (§6.2).<br />

A rigid coherency gate may not be useful <strong>in</strong> dead band as the signal-to-noise ratio is<br />

least here. Coherency Weighted Estimate (Stodt [1983]) or Coherency sort<strong>in</strong>g will give<br />

better results. However as shown by Egbert and Livelybrooks [1996], a comb<strong>in</strong>ation of<br />

these schemes with robust process<strong>in</strong>g may significantly improves the results (§6.2).<br />

With no coherent noises, coherency gate is arguably the best choice. However, coherence<br />

cannot always be relied for the discrim<strong>in</strong>ation, when the measured <strong>data</strong> conta<strong>in</strong>s<br />

coherent noises (Banks [1998]). W<strong>in</strong>dows that are contam<strong>in</strong>ated with coherent noises<br />

can generate transfer functions significantly different from the normal ones. A realistic<br />

approach would be a comb<strong>in</strong>ation of the coherence and robust schemes ( Egbert and<br />

Livelybrooks [1996]). As shown <strong>in</strong> §6, with a proper <strong>in</strong>itialization, robust process<strong>in</strong>g can<br />

remove the noisy <strong>data</strong> (see §6.2.4.2).<br />

3.3.4 Variance & Errors<br />

Estimates of error <strong>in</strong> computed transfer function are important, as they are required at<br />

the model<strong>in</strong>g or <strong>in</strong>version stage to assess the adequacy of fit of a derived model to the<br />

measured <strong>data</strong> (Eisel and Egbert [2001]) Statistical variance for each component of the<br />

transfer function tensor (equation 3.1) may be as,<br />

(∆Z xy ) 2 = k F (k, 2N − 4, δ = 0.05)[1 − γ2 ExExp ]E xE ∗ x<br />

(2N − 4) [1 − γ 2 HxHy ]H yH ∗ y<br />

(3.41)<br />

Where δ = 0.05 is the significance level correspond<strong>in</strong>g to a 0.95 or 95 per cent confidence<br />

limit, N is the number of Fourier coefficients used, 2N-4 is the degree of freedom,<br />

F(m,n,δ) is the factor for F distribution (Bendat and Piersol [1971]) and k is 4 for confidence<br />

limit of |Z xy | 2 (Schmucker [1978]). The errors for apparent resistivity and phase are<br />

∆ρ xy = 2µ |Z ω xy| 2 ∆Z xy and<br />

∆ϕ xy =<br />

∆Z xy<br />

|Z xy|<br />

(3.42)<br />

The estimate of variance <strong>in</strong> equation (3.41) is exact under the assumption that the<br />

noise <strong>in</strong> output channel (r <strong>in</strong> equation 3.25) are normally distributed and statistically<br />

<strong>in</strong>dependent while at the same time the <strong>in</strong>puts are error free. Chave and Thomson [1989]<br />

describe many <strong>in</strong>stances where<strong>in</strong> these assumptions can break down result<strong>in</strong>g <strong>in</strong> biased<br />

estimation of confidence limit.<br />

3.3.5 Coherence and bias of transfer functions<br />

Figure 3.5 shows MT transfer function (Z xy ) computed <strong>us<strong>in</strong>g</strong> equations 3.32 (up biased)<br />

and 3.37 ( down biased) respectively, compared to their associated coherence and variance<br />

for station VP20 (see section § 4.3.3). The upward and downward biased estimates


47<br />

Figure 3.5: MT Transfer function (Z xy ) computed <strong>us<strong>in</strong>g</strong> upward and downward biased<br />

estimators compared with the coherence functions for station VP20. The upper part<br />

of the diagram shows real and imag<strong>in</strong>ary components of upward and downward biased<br />

estimates. The variance of Z xy is shown as solid l<strong>in</strong>e. The lower part shows the predicted<br />

(multiple) and ord<strong>in</strong>ary coherence functions. See the text for discussion.<br />

behave identically over most of the period range except near 8 seconds and periods ><br />

200 sec. In these ranges, the two different estimates deviate from each other. The up<br />

biased estimates show considerable bias from the common trend. Though the apparent<br />

magnitude of bias seems to be less, the down biased estimates show reduced values <strong>in</strong><br />

these period ranges. On <strong>in</strong>spection on the coherence plots, it can be seen that these<br />

period- ranges are associated with low predicted coherence. The ord<strong>in</strong>ary coherence<br />

function E x H y decreases with <strong>in</strong>crease <strong>in</strong> period whereas coherence E x H x behaves <strong>in</strong><br />

exactly opposite way. The <strong>in</strong>creased dependency of E x on H x <strong>in</strong> the longer period might<br />

be caused by the polarization of electric fields by geological structures at depth. However,<br />

the predicted coherence function is undisturbed by the polarization. This is due to the<br />

fact that predicted coherence takes <strong>in</strong>to consideration, the transfer function between<br />

coll<strong>in</strong>ear fields as well (<strong>in</strong> this case Z xx ).<br />

The iterative scheme proposed by Kao and Rank<strong>in</strong> [1977] averages the up and down<br />

biased estimates to obta<strong>in</strong> bias free transfer functions. But this assumes that both electrical<br />

and magnetic channels are equal <strong>in</strong> noise content. Any deviation from this (as<br />

practically seen, electric channels are more noisy than magnetic channels Travassos and<br />

Beamish [1988]) will results <strong>in</strong> <strong>in</strong>ferior estimate than that of conventional ones. This was<br />

agreed by the authors while reply<strong>in</strong>g to a question by Hernandez and Jacobs [1979].


48<br />

3.3.6 Remote reference<br />

In § 3.3.6, we have seen the bias error <strong>in</strong> estimated transfer functions due to auto-power<br />

terms <strong>in</strong> equations such as 3.32 and 3.37. One effective way to reduce bias errors has been<br />

the remote reference (RR, it is usual to denote s<strong>in</strong>gle station process<strong>in</strong>g as SS) method<br />

(Gamble et al. [1979]), <strong>in</strong> which two <strong>in</strong>dependent signal channels are recorded for use as<br />

the complex conjugate part of cross and auto powers <strong>in</strong> equation such as 3.37.<br />

Ẑ xy = H xR ∗ xE x R ∗ y − H x R ∗ yE x R ∗ x<br />

H x R ∗ xH y R ∗ y − H x R ∗ yH y R ∗ x<br />

(3.43)<br />

Equation 3.43 gives the RR estimate of Z xy . By provid<strong>in</strong>g a k<strong>in</strong>d of synchronous<br />

detection, the method helps compensate for noise, both <strong>in</strong>ternal and external to the measur<strong>in</strong>g<br />

device (Vozoff [1991]). The remote channels usually designated R x and R y are<br />

the most commonly H x and H y components measured at a remote, noise free site. In<br />

pr<strong>in</strong>ciple the electric field also can be used as a remote reference. However, as shown by<br />

Travassos and Beamish [1988], this may not result <strong>in</strong> better transfer function estimate, as<br />

the electric fields are more affected (polarized) by local geology and the spatial coherence<br />

between electric field components are often small. Comb<strong>in</strong>ed with robust (§ 6.2) estimations,<br />

this technique is widely used at present. Jones et al. [1989] proved its superiority<br />

<strong>in</strong> a comparison of different process<strong>in</strong>g techniques on MT time series. In a recent work<br />

by Shalivahan and Bhattacharya [2002] the question of ‘how remote can the far remote<br />

reference site be?’ is addressed. They processed MT <strong>data</strong> from one permanent site with<br />

remote fields collected at distances 80, 115 and 215 kilometers away from it. Only by<br />

<strong>us<strong>in</strong>g</strong> the farthest station <strong>data</strong>, they could improve quality of MT <strong>data</strong> <strong>in</strong> all frequency<br />

ranges. However, the <strong>data</strong> used for the thesis were collected <strong>in</strong> s<strong>in</strong>gle station mode,<br />

though more than one systems were used <strong>in</strong> the field. Instrument problems prevented<br />

the synchronization <strong>data</strong> acquisition system’s <strong>in</strong>ternal clock with that of GPS (Global<br />

Position<strong>in</strong>g System) satellite. For this reason, the various approaches made <strong>in</strong> the thesis<br />

are for s<strong>in</strong>gle station process<strong>in</strong>g, though it can easily be extended to dual station process<strong>in</strong>g.<br />

In some cases there is scope to improve remote reference process<strong>in</strong>g as shown <strong>in</strong><br />

the § 6.3.


Chapter 4<br />

Signal and Noise Characteristics of<br />

MT <strong>data</strong> measured over the<br />

Southern Granulite Terra<strong>in</strong><br />

49


50<br />

4.1 Introduction.<br />

<strong>Magnetotelluric</strong> studies were carried out as part of an <strong>in</strong>tegrated geophysical study of<br />

the Southern Granulite Terra<strong>in</strong> (SGT) dur<strong>in</strong>g 1998-2000. The region south of the 13 o<br />

N latitude mark<strong>in</strong>g the high-grade granulite zone was traversed by a N-S corridor of<br />

geophysical observations (Figure 4.1). Of these, MT studies were conducted along 2 N-S<br />

profiles extend<strong>in</strong>g from Kuppam, near Bangalore <strong>in</strong> the north to Palani and Kodaikanal<br />

<strong>in</strong> the south. The <strong>Magnetotelluric</strong> measurements carried out over the Southern Granulite<br />

Terra<strong>in</strong> (SGT) <strong>in</strong> two field campaigns dur<strong>in</strong>g September 1998 – December 1998 and<br />

January 2000 - March 2000 form the basic <strong>data</strong> used <strong>in</strong> the thesis. These MT studies<br />

form part of a major geophysical and geological study on the SGT <strong>in</strong>itiated by the Deep<br />

Cont<strong>in</strong>ental Studies programme of the Department of Science and Technology. There are<br />

three factors that make this region important <strong>in</strong> terms of MT signal <strong>analysis</strong>. 1) The<br />

majority of the upper crustal rocks <strong>in</strong> the SGT belong to the Archaean and the Proterozoic<br />

age (Naqvi and Rogers [1987]) and exhibit high electrical resistivity as other shield<br />

regions <strong>in</strong> the world (Mareschal et al. [1994]). Highly resistive upper crust offers very little<br />

attenuation to EM signals and <strong>in</strong> pr<strong>in</strong>ciple noise can propagate larger distance on the<br />

SGT as compared to regions where younger rocks are exposed. 2) The high population<br />

density <strong>in</strong> the southern states of India, especially <strong>in</strong> Tamil Nadu, where most of the MT<br />

sites are located give rise to various cultural noises (§ 2.3). The <strong>in</strong>dustrial belt along the<br />

two banks of Cauvery River is another noise source for MT. And 3) while study<strong>in</strong>g the<br />

effect of Equatorial Electrojet (EEJ) on magnetotelluric <strong>data</strong>, measured further south<br />

of the ma<strong>in</strong> measurement corridor, Rao et al. [2002] reports perceptible decrease <strong>in</strong> daytime<br />

apparent resistivity curves compared to night-time estimates <strong>in</strong> periods excess of<br />

100 Sec. The <strong>analysis</strong> of MT <strong>data</strong>, <strong>in</strong> context of its signal and noise characteristics <strong>in</strong><br />

time, frequency and space doma<strong>in</strong> is discussed <strong>in</strong> this chapter. Har<strong>in</strong>arayana et al. [2003]<br />

discusses prelim<strong>in</strong>ary <strong>in</strong>terpretation based on 2D MT model<strong>in</strong>g along three NS profiles.<br />

4.2 Geological objectives of MT <strong>in</strong>vestigations <strong>in</strong> the<br />

SGT<br />

Deep structure of South Indian Shield Region (SISR) (Figure 4.1) has warranted the<br />

attention of earth scientists ow<strong>in</strong>g to its association with various tectonic features. It is<br />

characterized by expanses of high-grade crystall<strong>in</strong>e prov<strong>in</strong>ces south of 13 o N. Regional geology<br />

has been well mapped and studied, result<strong>in</strong>g <strong>in</strong> the identification of gradual <strong>in</strong>crease<br />

<strong>in</strong> high-grade metamorphic assemblages. It comprises of Archaean and Proterozoic terra<strong>in</strong><br />

and exposes major crustal scale thrust faults and tectonic l<strong>in</strong>eaments. The geology of<br />

the South Indian Pen<strong>in</strong>sula could be envisaged as a northward plung<strong>in</strong>g structure that exposes<br />

the Archaean craton and adjacent shallow green stone belts to the north and deeper<br />

high grade granulites to the South <strong>in</strong> an oblique section (Naqvi and Rogers [1987]. These<br />

features seem to have played a major role <strong>in</strong> the evolution of the cont<strong>in</strong>ental lithosphere<br />

of the region. Several models of tectonic evolution of the region, based on metamorphic,<br />

structural and geophysical <strong>data</strong> have been proposed (Drury et al. [1984], Radhakrishna<br />

[1989]). Earlier geophysical studies <strong>in</strong>clude regional gravity (Mishra [1988]), MAGSAT


Figure 4.1: MT stations superimposed over the geology of the measurement corridor <strong>in</strong><br />

South India (Geology adapted from GSI [1995]). See text for discussion<br />

51


52<br />

(Mishra and Venkatarayudu [1985]), aeromagnetic studies (Reddi et al. [1988]), seismic<br />

tomographic studies (Rai et al. [1993]) and electromagnetic <strong>in</strong>duction studies (Nityananda<br />

et al. [1977], Nityananda and Jayakumar [1981]). Dur<strong>in</strong>g a magnetometer array study <strong>in</strong><br />

South India, no large scale electrical conductivity anomaly was observed <strong>in</strong> Kuppam –<br />

Salem area, <strong>in</strong> contrast to crustal electrical anomalies observed along coastal marg<strong>in</strong>s <strong>in</strong><br />

the Southern granulite block ( Thakur et al. [1986]). A detailed discussion of the more<br />

recent <strong>in</strong>sights <strong>in</strong>to crustal evolution of South Indian shield has been compiled by Mahadevan<br />

[1994]. In order to obta<strong>in</strong> more <strong>in</strong>formation about the deep crust of south India<br />

and its evolution, <strong>in</strong>tegrated geophysical studies <strong>in</strong> the Southern Granulite Terra<strong>in</strong> (SGT)<br />

were <strong>in</strong>itiated under Deep Cont<strong>in</strong>ental Studies program of Department of Science and<br />

Technology, Government of India. Studies <strong>in</strong>volv<strong>in</strong>g co<strong>in</strong>cident seismic refraction and<br />

reflection profil<strong>in</strong>g, magnetotellurics, gravity and deep resistivity sound<strong>in</strong>gs have been<br />

completed (For a detailed report, see Memoir of Geol. Soc. Ind. No 50, 2003).<br />

4.3 The Data<br />

4.3.1 Acquisition<br />

Figure 4.1 shows the locations of the selected MT stations superimposed on the geology<br />

of the measurement corridor. The corridor, which roughly trends NS, lies <strong>in</strong> the southern<br />

pen<strong>in</strong>sula of India, with an approximate dimension of 300 X 100 km. The MT <strong>data</strong> were<br />

acquired ma<strong>in</strong>ly along three profiles, viz. Kuppam-Bommidi (KB), Omalur- Kodaikanal<br />

(OK) and Kolattur-Palani (KP), all be<strong>in</strong>g roughly oriented NS, with a station spac<strong>in</strong>g of<br />

around 10 km. The corridor cuts across major geologic and tectonic elements of the region<br />

– highly resistive high grade granulite gniess, ultra-basics, granite –gniesses and low-grade<br />

amphibolite granulite facies <strong>in</strong> the south (Naqvi and Rogers [1987]). For the MT survey,<br />

a wide-band digital MT system (GMS05 from Metronix GmbH) was used to measure<br />

the five electromagnetic field components <strong>in</strong> s<strong>in</strong>gle station mode. The electric fields were<br />

measured <strong>us<strong>in</strong>g</strong> porous pots with Cd-CdCl 2 electrodes. Four electrodes were laid <strong>in</strong> a<br />

cross arrangement with 90m distance between the pairs. Induction coil magnetometers<br />

with ∼30,000 turns were used to measure the natural magnetic fields (§ 1.2.3). MT time<br />

series were measured <strong>in</strong> four overlapp<strong>in</strong>g frequency bands (§ 1.2.4) for a period of 24-48<br />

hours per site, with a maximum possible frequency range, (4096 sec) −1 to 8192 [Hz].<br />

At most of the sites the <strong>data</strong> were recorded <strong>in</strong> three sessions. Out of three, one of the<br />

session would carry out relatively longer period of measurements than other sessions.<br />

It was typical to have around 100 stacks (1 stack = 1024 <strong>data</strong> po<strong>in</strong>ts) for band 1 to 3<br />

and 80 to 100 stacks for band 4. This effectively means a total record<strong>in</strong>g time of more<br />

than 24 hours. In the other sessions, the quantity of <strong>data</strong> collected would be less. As the<br />

signal and noise <strong>in</strong> MT are time variant processes, the chief session need not always result<br />

<strong>in</strong> best estimate of MT transfer functions. However, the present thesis concentrate on<br />

process<strong>in</strong>g the <strong>data</strong> from the chief session of each site, as this conta<strong>in</strong>s the largest volume<br />

of cont<strong>in</strong>uous <strong>data</strong>. A selected subset of about 24 MT stations is used to demonstrate<br />

the methodologies developed for MT time series process<strong>in</strong>g. MT <strong>data</strong> from sites with<br />

differ<strong>in</strong>g signal/noise distribution and geology were <strong>in</strong>cluded <strong>in</strong> this set of 24 stations.<br />

Table 4.1 shows location, Village names and stations codes, record<strong>in</strong>g date etc.


53<br />

4.3.2 Typical MT Time series<br />

The observed MT time series exhibit vary<strong>in</strong>g amplitude, pattern and frequency content<br />

based on the signal as well as noise sources, which <strong>in</strong> turn are time variant processes.<br />

It is important to <strong>in</strong>spect the time series to identify obvious outliers <strong>in</strong> time series and<br />

remove those segments from further process<strong>in</strong>g (Manoj and Nagarajan [2003]). Time<br />

series samples measured <strong>in</strong> the four frequency bands <strong>in</strong> South India are displayed, as<br />

representative examples <strong>in</strong> Figures 4.2 and 4.3 .<br />

Figure 4.2(a) shows a snapshot of MT time series (1024 po<strong>in</strong>ts) measured <strong>in</strong> the band<br />

1 (sampl<strong>in</strong>g rate was 32 kHz). Here the natural signal is superimposed on power l<strong>in</strong>e<br />

frequency and its harmonics. The pattern of the signal <strong>in</strong> this frequency range is pseudos<strong>in</strong>usoids.<br />

Superimposed on such a background, a burst of signal activity is observed<br />

at 0.015 sec. This activity is correlated well over four of the five measured channels.<br />

The obvious source for this transient signal is lightn<strong>in</strong>g (§ 2.2.2). S<strong>in</strong>ce the transient<br />

envelope is rich <strong>in</strong> high frequency content, the lightn<strong>in</strong>g might have occurred near to the<br />

site. But the term ’near’ has to treated cautiously. If the light<strong>in</strong>g were occurred so near<br />

MT Station Village Latitude Longitude Day Start Day End<br />

JN02 Venkatapalli 12.900917 78.3894167 12-Jan-00 14-Jan-00<br />

JN07 Kochchakallur 12.402861 78.3060556 22-Jan-00 24-Jan-00<br />

JN10 Puliampatti 12.152708 78.3267083 28-Jan-00 29-Jan-00<br />

JN12 Regadahalli 12.031389 78.2559722 29-Jan-00 01-Feb-00<br />

EW01 Vellicahndi 12.335361 78.0815833 03-Feb-00 05-Feb-00<br />

VP01 Vellar 11.892528 77.9628056 07-Feb-00 09-Feb-00<br />

VP03 Ellakutaur 11.779278 77.9203889 11-Feb-00 12-Feb-00<br />

TT06 Ramudaiyanur 11.156917 78.1988889 12-Feb-00 14-Feb-00<br />

KG01 Tottik<strong>in</strong>aru 11.795417 77.7229722 14-Feb-00 16-Feb-00<br />

KG02 Pakapudur 11.718194 77.6813056 16-Feb-00 17-Feb-00<br />

TT08 Valasiramani 11.123028 78.3431944 16-Feb-00 18-Feb-00<br />

KG03 Kottur 11.636722 77.6697222 17-Feb-00 18-Feb-00<br />

TT09 Venkatachalapuram 11.244278 78.5219722 17-Feb-00 19-Feb-00<br />

TT10 Viralipatti 11.12175 78.7151389 19-Feb-00 20-Feb-00<br />

VP11 Pudupalayam 11.16875 77.6191111 22-Feb-00 23-Feb-00<br />

VP10 Siliamaptti 11.300222 77.5708611 23-Feb-00 23-Feb-00<br />

VP13 Muttukkalivalasu 10.968222 77.5275556 23-Feb-00 25-Feb-00<br />

VP12 Rasipalayam 11.072722 77.5706389 24-Feb-00 26-Feb-00<br />

VP16 Talakkarai 0.796778 77.5331111 25-Feb-00 27-Feb-00<br />

VP15 Gollipatti 10.703222 77.5446944 26-Feb-00 27-Feb-00<br />

VP17 Kuppanavalasu 10.635556 77.52925 27-Feb-00 28-Feb-00<br />

OK15 Kuthilabbai 10.601583 77.7785556 28-Feb-00 29-Feb-00<br />

VP19 Karadikuttam 10.452778 77.4468889 28-Feb-00 29-Feb-00<br />

OK18 Viralipatty 10.140778 77.7341389 01-Mar-00 03-Mar-00<br />

Table 4.1: The locations and measurement details for the MT stations


Figure 4.2: Typical MT time series measured <strong>in</strong> the SGT for two frequency ranges a)<br />

the frequency range 256 Hz to 32kHz (band 1) and b) the period range 8 Hz to 256 Hz<br />

(band2) .See the text for discussion.<br />

54


55<br />

that the em fields due to light<strong>in</strong>g behaved non planar at the site (§ 1.1.1), this burst<br />

would be treated as noise. By construct<strong>in</strong>g MT transfer function between the channels<br />

and verify<strong>in</strong>g its consistency between segments, one can easily check this. The telluric<br />

signals fluctuates with<strong>in</strong> ∼100mV/km and magnetic fields with<strong>in</strong> 0.2 nT. But <strong>in</strong> certa<strong>in</strong><br />

<strong>in</strong>stances the telluric fields can have even higher amplitudes. MT time series measured at<br />

site TT08 <strong>in</strong> band 2 (sampl<strong>in</strong>g rate 1024 Hz) is presented <strong>in</strong> Figure 4.2(b). The narrow<br />

band noise present <strong>in</strong> all the channels is noticeable. They are power l<strong>in</strong>e signals (50 Hz),<br />

perfectly manifested <strong>in</strong> MT time series as s<strong>in</strong>usoids. The time series <strong>data</strong> were filtered<br />

with analogue notch filters (50 & 150 Hz) before be<strong>in</strong>g digitized. The high power of the<br />

50 Hz signals even after notch filter<strong>in</strong>g shows the level of 50 Hz signals prevalent <strong>in</strong> the<br />

area. Consider the higher amplitude for telluric and magnetic fields compared to band 1.<br />

The visual <strong>in</strong>spection of this time series will not yield any first hand <strong>in</strong>formation about<br />

the signal content, its distribution over frequency etc. It is common practice to look for<br />

large spikes or discont<strong>in</strong>uity <strong>in</strong> the time series, to exclude such segments from further<br />

process<strong>in</strong>g. The large spike seen at 0.3 sec and is reflected <strong>in</strong> all the channels is one of<br />

such examples. The magnetic fields show a sharp decreas<strong>in</strong>g trend at the beg<strong>in</strong>n<strong>in</strong>g of<br />

each channel. It repeats <strong>in</strong> all the segments and could most probably be an artefact from<br />

measurement system.<br />

MT time series measured <strong>in</strong> Band3 (Figure 4.3(a) <strong>in</strong>cludes the well-known MT ’dead’<br />

band around 1 sec (§ 2.2.1). The natural electromagnetic signal energy (Figure 2.1) is low<br />

<strong>in</strong> this band and the time series usually look like a random process. The <strong>data</strong> displayed<br />

<strong>in</strong> Figure 4.3(a) were measured at VP12, which is relatively a noise free site (see §4.3.3 <strong>in</strong><br />

this chapter). The <strong>data</strong> were measured with a sampl<strong>in</strong>g rate 32 Hz and the duration of<br />

the displayed time series is 128 sec( 4096 samples). The time series fluctuates randomly<br />

from a common mean, with some large spike activity <strong>in</strong> between. Spike activity is more<br />

<strong>in</strong> Ex, Hy and Hz channels and is less <strong>in</strong> Ey and Hx channels. But the major spikes, for<br />

example the one occurs at 95 sec is well correlated between the channels. This <strong>in</strong>dicates<br />

an <strong>in</strong>ductive source for the noise and was probably due to switch<strong>in</strong>g of large power loads.<br />

The amplitude of telluric signals are <strong>in</strong> the order of ∼4 mV/Km and for magnetic fields<br />

it is ∼0.01 nT/s.<br />

Visual <strong>in</strong>spection of time series (§ 5) is more appropriate <strong>in</strong> the measurement band 4<br />

(sampl<strong>in</strong>g rate 1 Hz). The time series measured at JN10 are presented <strong>in</strong> Figure 4.3(b),<br />

conta<strong>in</strong>s 4096 samples. The ma<strong>in</strong> signal source for long period MT measurements are<br />

geomagnetic pulsations. Envelopes of pulsation activity (Pc4 and Pc5 - see §2.2.2) are<br />

observed near 1500, 1900, 2500 and 3500 seconds and is best observed <strong>in</strong> Hx. The spike<br />

activity with much larger magnitude made the pulsation activity less obvious <strong>in</strong> other<br />

channels. In addition to the short period pulsations, fluctuations with longer period<br />

were also observed (and correlated) <strong>in</strong> all the channels. The magnitudes of the signals<br />

displayed were largely controlled by the spike activity.<br />

4.3.3 Examples of MT <strong>data</strong> collected over South India<br />

Station VP12 (western part) and TT08 (eastern part) lie <strong>in</strong> the centre of the measurement<br />

corridor (Figure 4.1) and are near to a major EW trend<strong>in</strong>g shear zone <strong>in</strong> the


Figure 4.3: Typical MT time series measured <strong>in</strong> the SGT for two frequency ranges c) the<br />

frequency range 8Hz to 0.25 Hz (band 3) and d) the period range 4 sec to 128 sec.See<br />

the text for discussion.<br />

56


57<br />

region. JN10 and VP20 are from northern and southern part of the corridor, respectively.<br />

All of the sites are on exposed crystall<strong>in</strong>e gneisses of Archaean / Proterozoic age. The<br />

time series <strong>in</strong> each station were processed <strong>in</strong> the follow<strong>in</strong>g manner:<br />

1. For each measurement band, the total available time series were sliced <strong>in</strong>to sub<br />

segments with the constra<strong>in</strong>ts of lowest frequency of <strong>in</strong>terest and required degree<br />

of freedom for each estimate.<br />

2. The bias and trend of the time series were elim<strong>in</strong>ated by procedures outl<strong>in</strong>ed <strong>in</strong> §<br />

3.2.3.1.<br />

3. To reduce the bias of the spectra due to f<strong>in</strong>ite measurements, a hann<strong>in</strong>g w<strong>in</strong>dow<br />

was applied to each time segments, followed by FFT and calibration.<br />

4. Smooth cross and auto spectra of MT field components were estimated for each<br />

time series segment by frequency band averag<strong>in</strong>g (§ 3.2.3.5).<br />

5. Robust process<strong>in</strong>g (§ 6.2) was applied to obta<strong>in</strong> global spectral matrices from all<br />

the segments.<br />

Apparent resistivity, phase, coherence between predicted observed electrical field (§<br />

3.3.5) and degree of freedom for each estimates are presented for sites VP12 <strong>in</strong> the Figure<br />

4.4(a). Data are presented <strong>in</strong> their measured co-ord<strong>in</strong>ates. The xy and yx components<br />

of apparent resistivity show split that becomes appreciable for frequencies below 0.01Hz.<br />

As the phase components also behave differently from each other, it could be the result of<br />

lateral resistivity contrast. Predicted coherence are above 0.8 for most of the frequency<br />

range, but with a low <strong>in</strong> the dead band (5 Hz to 10 Sec). The degrees of freedom<br />

(dof ) exponentially decrease with frequency <strong>in</strong> each band. Fourier transform results <strong>in</strong><br />

equally spaced harmonics <strong>in</strong> l<strong>in</strong>ear scale, where as the target frequency to compute MT<br />

parameters are equally spaced <strong>in</strong> logarithmic scale. Due to this, the number of spectral<br />

l<strong>in</strong>es available for target frequencies exponentially reduces with decrease <strong>in</strong> frequency.<br />

In addition to this the selective stack<strong>in</strong>g processes modify the dof. But the effect of<br />

degree of freedom seems to be m<strong>in</strong>imal on the predicted coherence. In fact, near 0.1 Hz<br />

the even with an effective dof of 2000, both the coherences are low. The steep rise <strong>in</strong><br />

apparent resistivity and low values for phase observed between 10,000 Hz to1,000 Hz is<br />

clearly the effect of a near field signal. Computed error bars (xy) shows large values <strong>in</strong><br />

the dead band and frequency less than 0.01 Hz. Whereas the large errors for dead band<br />

result from the poor signal strength, for lower frequency range, where the coherences<br />

are relatively high, low dof are the reason for large error bars. Station JN10 lies <strong>in</strong> the<br />

northern part of the measurement corridor and is relatively noisy compared to VP12.<br />

The apparent resistivity values (Figure 4.4(b)) monotonously decrease with <strong>in</strong>crease <strong>in</strong><br />

frequency. Phase smoothly varies except <strong>in</strong> the dead band. The coherences assume low<br />

values between 10 Hz to 0.01 Hz. Quantity of observed <strong>data</strong> is less compared to VP12, as<br />

seen on dof plot. TT08 lies <strong>in</strong> the eastern part of the measurement corridor and is heavily<br />

affected with noise especially <strong>in</strong> band 3 & band 4 (frequencies below 8 Hz). Apparent<br />

resistivity and phase values of xy (Figure 4.4(c)) component are scattered and have larger<br />

associated errors. Though the quantity of measured <strong>data</strong> is relatively high (see the dof


Figure 4.4: Plot of apparent resistivity, phase, predicted coherence and degree of freedom<br />

(DOF) vs frequency for four stations measured over South India. Data are computed <strong>in</strong><br />

their measured direction. See text for discussion<br />

58


59<br />

plot), the predicted coherences (E xp E x ) are less than 0.6 for majority of the frequency<br />

range. Apparent resistivity observed <strong>in</strong> frequencies less than 10 Hz is much lower than<br />

one would expect on a crystall<strong>in</strong>e terra<strong>in</strong>. The MT transfer functions are evidently biased<br />

<strong>in</strong> the frequency range 10 Hz to 0.01 Hz by noise. VP20 is located <strong>in</strong> the southern part<br />

and Figure 4.4(d) presents the MT <strong>data</strong> from this station. The site was relatively noise<br />

free and the apparent resistivity and phase values vary smoothly and are self-consistent.<br />

Still there is a clear evidence for bias error <strong>in</strong> the ρ xy values for frequency less than 0.01<br />

Hz. The coherence plot also shows low values (E xp E x ) for correspond<strong>in</strong>g frequencies. To<br />

conclude, bias and random errors due to noise are evident <strong>in</strong> many of the sites and it is<br />

not always reflected <strong>in</strong> the associated coherences and errors.<br />

4.4 The EEJ effect on MT <strong>data</strong><br />

Equatorial Electrojet (EEJ) is a non-uniform east flow<strong>in</strong>g current <strong>in</strong> the ionosphere,<br />

with<strong>in</strong> 5 0 each side of the magnetic equator. The direct <strong>in</strong>cidence of sun’s radiation at<br />

equatorial regions, ionizes the upper atmosphere, where geomagnetic field is essentially<br />

horizontal. This current amplifies the northward component of slower magnetic variations.<br />

The EEJ will also affect the geomagnetic pulsation (§ 2.2.2) amplitude (Sarma<br />

et al. [1982], Sastry et al. [1983]) of MT signal below 3 Hz. The two reported studies on<br />

the effect of equatorial electrojet on measured MT transfer functions at sites close to dip<br />

equators are by Padilha et al. [1997] and by Rao et al. [2002]. MT <strong>data</strong> were collected<br />

along 1000-km profile <strong>in</strong> Brazil, with the dip equator pass<strong>in</strong>g through the center of the<br />

profile. The time series were measured <strong>in</strong> day and night were processed separately <strong>us<strong>in</strong>g</strong><br />

conventional as well as robust process<strong>in</strong>g, to obta<strong>in</strong> apparent resistivity and phase for<br />

the un-rotated diurnal and nocturnal tensor elements (Padilha et al. [1997]). However,<br />

the comparison between daytime and nighttime results did not show any significant difference<br />

<strong>in</strong> the entire period band. The MT curves were nearly identical, with<strong>in</strong> very low<br />

error bars. This was true for all the measured sites along the profile, <strong>in</strong>dicat<strong>in</strong>g that EEJ<br />

currents did not affect the MT <strong>data</strong>. The theoretical model<strong>in</strong>g of EEJ, approximated<br />

as a conductive layer at 110 km, the authors found EEJ may affect MT responses at<br />

periods greater than 1000s. But they did not have enough observed <strong>data</strong> at comparable<br />

period range. They concluded that the theoretically anticipated EEJ distortions are<br />

probably overestimated and the plane wave assumption of MT signals at equatorial regions<br />

is valid at least <strong>in</strong> the frequency range 1000 to 0.0005 [Hz]. But recent studies by<br />

Rao [2000] and Rao et al. [2002], on analyz<strong>in</strong>g MT <strong>data</strong> collected near and away from<br />

Indian magnetic dip equator showed the effect of EEJ is appreciable above 100 sec. The<br />

<strong>data</strong> were collected dur<strong>in</strong>g the same field campaign <strong>in</strong> South India, the stations be<strong>in</strong>g<br />

south of the ma<strong>in</strong> measurement corridor. The <strong>data</strong> collected at three sites viz OK3, IRA<br />

and KAR, which lie <strong>in</strong> a NS profile, with KAR be<strong>in</strong>g southern most and at the center<br />

of dip equator. MT time series measured at each site were separated <strong>in</strong>to daytime and<br />

nighttime segments. After omitt<strong>in</strong>g the sub-segments with obvious outliers, <strong>in</strong>dependent<br />

daytime and nighttime estimates of MT impedance were made. It was observed (Figure<br />

4.5(a) to (c) that the ρ yx component is very similar between day and night estimates at<br />

OK3 as well as IRA. But at station KAR, which is at the center of EEJ, the daytime


60<br />

Figure 4.5: Apparent resistivity vs frequency plot for 3 stations tht are near and away<br />

from the dip equator with day and night curves superimposed (Rao et al. [2002]). (a)<br />

OK3 - site 300 km north of dip equator, (b) IRA site 100 km north of dip equator and<br />

(3) KAR - site near the dip equator.<br />

estimates were biased down at frequencies less than 0.01 [Hz]. This corroborate the idea<br />

that, <strong>in</strong>creased signal amplitude for H x as a result of the eastward current flow associated<br />

with EEJ might bias the MT transfer functions.<br />

The effect of EEJ on MT transfer functions seem to be decreas<strong>in</strong>g fast as one goes<br />

away from the dip equator, as evidenced by the similarity <strong>in</strong> daytime and nighttime curves<br />

of IRA and OK3, which lies to the north of dip equator and KAR. As the measurement


61<br />

Figure 4.6: Averaged Kp <strong>in</strong>dices [Courtesy NGDC], K <strong>in</strong>dices [HYB] dur<strong>in</strong>g the MT<br />

measurements are compared with telluric predicted coherence [4 sec 128 sec]. The mean<br />

coherence is drawn as a l<strong>in</strong>e.<br />

corridor for the present study is located ∼300 km north of the dip equator, the effect<br />

of EEJ on MT transfer functions may be treated as negligible, at least <strong>in</strong> the frequency<br />

range of measurement.<br />

4.5 Signal Activity dur<strong>in</strong>g the Field Campaign<br />

MT <strong>data</strong> were collected <strong>in</strong> two field campaigns <strong>in</strong> September - December 1998 and January<br />

– March 2000. In order to study the ability of different process<strong>in</strong>g methods to estimate<br />

MT transfer functions (> 1 sec) <strong>in</strong> the presence of both low and high noise levels, compared<br />

to signal levels, the geomagnetic activity dur<strong>in</strong>g the measurement were exam<strong>in</strong>ed. Shown<br />

<strong>in</strong> Figure 4.6 as histograms are the averaged daily global Kp and Hyderabad K <strong>in</strong>dices,<br />

with long period average coherence (§ 4.5). K <strong>in</strong>dices are a local quasi-logarithmic<br />

measure of geomagnetic activity (Jones et al. [1989]) and have 10 classes between K =0<br />

and K =9 (magnetic storm). K =9 correspond to a range of 300 nT for Hyderabad<br />

observatory. Where as Kp <strong>in</strong>dices are global <strong>in</strong>dicators of geomagnetic activity and<br />

are weighted average of a selection of local K <strong>in</strong>dices and with weights reflect<strong>in</strong>g the<br />

geomagnetic latitude and longitude. Note that some stations are omitted. The stations<br />

are plotted <strong>in</strong> sequential order from Jan 12 to Mar 3, 2000. The K and Kp <strong>in</strong>dices<br />

shows almost same magnitude and variation through out the measurements time, with<br />

a maximum at VP3 on 12 th February 2000. The low geomagnetic activity (<strong>in</strong>dices less<br />

than 2) is observed between station KG01 and VP11 (16-23 rd February) which <strong>in</strong> turn is<br />

flanked by two well def<strong>in</strong>ed highs (<strong>in</strong>dices > 3). Predicted coherence functions are good<br />

measure of the quality of MT <strong>data</strong> (§ 3.3.5), with<strong>in</strong> the statistical limit of the least square<br />

solution of MT transfer functions. Procedures used to compute the average bias-corrected<br />

coherence function are discussed <strong>in</strong> § 4.6. The averaged predicted coherence values for<br />

all the station varies with a mean of ∼0.7 for the measurement duration. They are<br />

generally high <strong>in</strong> the region of K > 0.3 and low at stations whose period of measurement<br />

co<strong>in</strong>cided with low geomagnetic activity. But stations KG03 and TT08 whose period of<br />

measurement have low geomagnetic <strong>in</strong>dices, show relatively high coherence values and<br />

station VP13 with K <strong>in</strong>dices > 0.3 shows low coherence. These deviation from the general


62<br />

trend, <strong>in</strong>dicate the presence of other factors which might have also contributed to the<br />

coherence distribution.<br />

4.6 Spatial character of coherence<br />

Predicted coherence function def<strong>in</strong>es optimality of the least square solution for MT transfer<br />

functions (§ 3.3.5). It has widely been used to exam<strong>in</strong>e the quality of the measured<br />

MT <strong>data</strong> (Vozoff [1972]). Though it is possible to compute predicted coherence functions<br />

for all the measured MT fields, telluric coherence functions are more frequently used<br />

for two obvious reasons i.e., 1) The magnetic fields are usually measured at a greater<br />

level of accuracy than the electric fields, 2) Electric fields are more prone to noise from<br />

man-made sources. Due to reasons 1 & 2, the least square estimation of MT transfer<br />

functions usually consider magnetic fields as noise-free <strong>in</strong>puts and m<strong>in</strong>imize the noise <strong>in</strong><br />

electric fields (Sims et al. [1971]). In this section, the distribution of averaged predicted<br />

coherence of E x and E y components for all the stations <strong>in</strong> the measurement corridor are<br />

described. The averaged predicted coherence is def<strong>in</strong>ed as,<br />

γ 2 = 1 2<br />

(<br />

γ<br />

2<br />

Ex−HxHy + γ 2 Ey−HxHy)<br />

(4.1)<br />

The <strong>in</strong>dividual coherence functions with<strong>in</strong> the bracket are def<strong>in</strong>ed <strong>in</strong> § 3.3.5. However,<br />

it is relatively well known that the predicted coherence functions get biased due to noise.<br />

Jones [1981] and Jones et al. [1983] describe a simple way to reduce the bias <strong>in</strong> coherence<br />

functions. For predicted coherence it is def<strong>in</strong>ed as,<br />

( ) 4 (1<br />

γb 2 = γ 2 −<br />

) )<br />

− γ<br />

2 2<br />

(1 + 4γ2<br />

(4.2)<br />

v − 2<br />

v<br />

Where γ b is the bias reduced estimate of γ and v is the degree of freedom for the<br />

estimate. In order to estimate the MT transfer functions, all the time segments available<br />

at the stations were processed and averaged the coherence without <strong>us<strong>in</strong>g</strong> any preferential<br />

stack<strong>in</strong>g algorithms. The idea here is to represent the noise contributions at each site<br />

and time segments with obvious outliers were also kept to estimate impedance tensor and<br />

their associated coherence functions. Consider<strong>in</strong>g the bias <strong>in</strong> coherence estimation <strong>in</strong> the<br />

presence of noise either correlated between the <strong>in</strong>put and output fields or un-correlated<br />

noise <strong>in</strong> the <strong>in</strong>put field, the coherence function over respective frequency bandwidth of<br />

measurements were averaged. At each site we have four estimates of predicted coherence<br />

functions <strong>in</strong> four bands of measurements. The coherence estimates at all the sites are<br />

gridded (with grid surface always go<strong>in</strong>g through the <strong>data</strong> po<strong>in</strong>ts) and presented as a<br />

contour map for the measurement corridor for each band <strong>in</strong> Figure 4.7.<br />

A total of 69 sites were used to prepare the map out of 80 stations. The predicted<br />

coherence functions of the telluric fields vary between 0.3 and 0.9 for most of the sites and<br />

over bands of measurements. The spatial distribution of coherence is however not similar<br />

for all the bands. The distribution patterns are similar for band 1 and band 2 (Figure<br />

4.7(a)&(b)) with coherence assum<strong>in</strong>g high values for most of the sites. Except for the NE<br />

apart and a site <strong>in</strong> the center of the measurement corridor, the coherence values are very


63<br />

Figure 4.7: The band averaged telluric predicted coherence plotted as a contoured map<br />

for all the measured MT sites. (a) Band1 (b) Band2 (c) Band 3 and (d) Band4.<br />

high (>0.8). As band1 and band2 covers the noise contributions from power l<strong>in</strong>es, there<br />

could be two reasons for high coherence for MT <strong>data</strong> <strong>in</strong> this range. If the noise source<br />

can be considered as ‘near field’ (§ 2.4) for that measurement site, the MT coherence may<br />

still be very high, but the apparent resistivity will just be a function of frequency, with<br />

least amount of <strong>in</strong>formation of subsurface resistivity. On the other hand, if measured<br />

sufficiently far from the noise source, so that the plane wave conditions for MT are met,<br />

then the power l<strong>in</strong>e signals can be an ideal source for the high frequency MT. However, it<br />

is impossible to dist<strong>in</strong>guish these two possibilities by analyz<strong>in</strong>g the coherences. The MT<br />

<strong>data</strong> <strong>in</strong> the middle frequency range (Band3 – 0.25 – 6 Hz) exhibits comparatively low<br />

values for coherence (Figure 4.7(c). As the natural signal strength <strong>in</strong> this frequency band<br />

are very low compared to other bands, it is not surpris<strong>in</strong>g to observe the low coherence<br />

for band3. However, the southern part of the measurement corridor (south of 11 0 Lat)<br />

has higher coherence values than the northern part. The low coherence for the NE part<br />

observed <strong>in</strong> Band1 and Band2 (Figure 4.7(a)&(b)) also is repeated <strong>in</strong> band3. Except for a<br />

site <strong>in</strong> the north and for an EW (11.5 0 Latitude) belt <strong>in</strong> the center, the MT measurements<br />

show fairly high coherence values <strong>in</strong> the long period measurements (Figure 4.7(d)). On<br />

a closer look at all the four plots it is apparent that a EW belt <strong>in</strong> the center of the<br />

measurement corridor shows low values throughout the frequency bands.


64<br />

4.7 Geographic relation of noise<br />

As we have seen <strong>in</strong> § 4.5 the geomagnetic <strong>in</strong>dices varied dur<strong>in</strong>g the measurement period.<br />

However, its direct implication on coherence of longer period <strong>data</strong> was not clear. One<br />

reason could be the vary<strong>in</strong>g effect of man-made noise <strong>in</strong> the MT <strong>data</strong>. The spatial<br />

distribution <strong>in</strong> predicted coherence (§ 4.6) consistently showed as EW belt of low values<br />

for all the measurement bands. In order to understand the relation of noises <strong>in</strong> MT<br />

<strong>data</strong> measured over South India to the major geographical elements of the region, the<br />

contoured plot for noise/(signal+noise) ratio for all the stations was superimposed on the<br />

geography map for the measurement corridor (Figure 4.8.The noise/ (signal+noise) ratio<br />

is def<strong>in</strong>ed as ,<br />

n<br />

s + n = (1 − γ2 ) (4.3)<br />

Where γ is the averaged predicted coherence def<strong>in</strong>ed <strong>in</strong> equation 4.1 & 4.2. The<br />

ratio is bounded by values 0 and 1. Major roads (thick black l<strong>in</strong>es) and rail tracks<br />

crisscross the measurement corridor. Major rail/road l<strong>in</strong>ks between two state capitals viz<br />

Bangalore and Chennai passes through the northern part of the corridor. Cauvery river<br />

flows through the center of the corridor, with major towns like Erode and Sankaridurg on<br />

its banks. In addition to the rail and road clusters, this area hosts a number of cement<br />

factories. Limestone is present all along the fr<strong>in</strong>ges of an exposed granite (Sankari)<br />

dome. Availability of limestone and water from Cauvery made this region suitable for<br />

such factories. The measurement corridor is relatively free of anthropogenic disturbance<br />

sources further south (ie south of 11 0 Latitude). The contour plot for N/(S+N) ratio<br />

shows an <strong>in</strong>terest<strong>in</strong>g relation to the distribution of man made electromagnetic disturbance<br />

sources. The northern region of the corridor shows three isolated highs at stations JN08,<br />

JN03 and JN05. Though it is spatially correlated well with the Chennai – Bangalore<br />

rail l<strong>in</strong>es and roads, the other three stations <strong>in</strong> the area viz. JN01, JN02 and EW02<br />

do not show high N/(S+N) ratios. In the middle of the measurement corridor, 8 MT<br />

stations distributed <strong>in</strong> and around the towns Erode and Sankaridurg shows very high<br />

N/(S+N) ratio. This cluster of highs <strong>in</strong> noise ratio clearly shows their aff<strong>in</strong>ity towards<br />

the <strong>in</strong>dustrial belt mentioned above. Though one expects the effect of the <strong>in</strong>dustrial zone<br />

to extend all along the banks of the river Cauvery, except for station TT08, the effect<br />

of noise is not perceptible further West. In l<strong>in</strong>e with the observation that the southern<br />

part of the corridor is devoid of major anthropogenic disturbance source, the N/(S+N)<br />

ratios of southern MT sites have relatively low values. In conclusion, <strong>in</strong> the northern and<br />

central part of the measurement corridor, the noise distribution is related to the known<br />

man-made electromagnetic disturbance sources.<br />

4.8 Discussion<br />

An attempt has been made <strong>in</strong> this chapter to characterize the signal and noise <strong>in</strong> the<br />

measured MT <strong>data</strong> over the Southern Granulite Terra<strong>in</strong>. Process<strong>in</strong>g of MT <strong>data</strong> start<br />

with visual <strong>in</strong>spection of the measured time series. As seen <strong>in</strong> Figure 4.2 and 4.3, the<br />

time series collected over the SGT shows effects of noise from various sources. Moreover


Figure 4.8: Contoured map of telluric Noise/(Signal+Noise) ratio, superimposed on the<br />

major geographical elements of the region. Note N/(S+N) ratio on either sides of Cauvery<br />

river near Sankari.<br />

65


66<br />

the noise is manifested <strong>in</strong> the time series with patterns different from the natural signals.<br />

However, visual <strong>in</strong>spection of time series will not remove all the noise <strong>in</strong> the measured<br />

<strong>data</strong>, as many noise processes have temporal characteristics similar to the signals. A more<br />

quantitative def<strong>in</strong>ition of signal and noise is required to separate them <strong>in</strong> MT <strong>data</strong>. One<br />

straightforward way to do this is to construct a transfer function between the measured<br />

electric and magnetic field. Now the portion of electric and magnetic fields that cannot<br />

be expla<strong>in</strong>ed by such a relation may be treated as noise. As we have seen <strong>in</strong> § 4.3.3,<br />

criteria for separat<strong>in</strong>g signal and noise based on the coherence and errors may not work<br />

always. This necessitated look<strong>in</strong>g at some other properties of the <strong>data</strong>, specifically, two<br />

<strong>in</strong>dependent observations on the processes contribut<strong>in</strong>g to signal and noise. In this respect<br />

the <strong>in</strong>dices of geomagnetic activity dur<strong>in</strong>g the measurement time were compared with<br />

the averaged long period coherence of each site. The agreement of the coherence and the<br />

geomagnetic <strong>in</strong>dices <strong>in</strong>dicate the validity of such an approach (Figure 4.6). However, a few<br />

disagreements are also evident. This <strong>in</strong>dicates the possible noise processes contribut<strong>in</strong>g<br />

to the measured <strong>data</strong>. Spatial relation of the known cultural noise centers and the quality<br />

of MT <strong>data</strong> was exam<strong>in</strong>ed. The strong correlation of the high noise/(signal +noise) ratio<br />

to the major <strong>in</strong>dustrial belt <strong>in</strong>dicate that the MT signals may get consistently degraded<br />

due to presence of an active noise sources, irrespective of the signal activity. However,<br />

the conclusions drawn from this chapter should be treated rather cautiously <strong>in</strong> <strong>in</strong>stances<br />

where the underly<strong>in</strong>g noise processes deviate from a Gaussian distribution. The signal<br />

and noise are def<strong>in</strong>ed <strong>in</strong> this chapter as <strong>in</strong>l<strong>in</strong>e and outl<strong>in</strong>e components of a least square<br />

solution of MT transfer functions. In the next chapter, we will take a closer look at<br />

the temporal properties of MT <strong>data</strong> and propose a new signal discrim<strong>in</strong>ation method,<br />

without the def<strong>in</strong>ition signal and noise from a least square solution.


Chapter 5<br />

The application of the artificial<br />

neural networks to magnetotelluric<br />

time series <strong>analysis</strong><br />

67


68<br />

5.1 Introduction.<br />

In the previous chapter the signal and noise characteristics of MT <strong>data</strong> acquired from<br />

SGT were analyzed. The noises from various sources manifest <strong>in</strong> MT <strong>data</strong> <strong>in</strong> different<br />

forms. A part of noise is obvious <strong>in</strong> the time series itself. The manual <strong>in</strong>spection of the<br />

time series is often the first step <strong>in</strong> MT <strong>data</strong> process<strong>in</strong>g, which selects a subset of time<br />

segments, followed by statistical transfer function estimation such as robust process<strong>in</strong>g.<br />

However, manual <strong>in</strong>spection has its own limitations. Firstly, it is a time consum<strong>in</strong>g process,<br />

which constitutes ∼ 80 % time used <strong>in</strong> MT time series <strong>analysis</strong>. Secondly, like other<br />

human decision mak<strong>in</strong>g processes, MT time series edit<strong>in</strong>g is also a subjective process,<br />

where<strong>in</strong> the same person may output differently, <strong>in</strong> a long session of edit<strong>in</strong>g. Consider<strong>in</strong>g<br />

the large volume of MT <strong>data</strong> collected over SGT, a need to develop a procedure<br />

to automate the manual edit<strong>in</strong>g of MT time series was felt. An approach was made to<br />

this problem from pattern recognition angle demonstrat<strong>in</strong>g the efficacy of artificial neural<br />

networks <strong>in</strong> discrim<strong>in</strong>at<strong>in</strong>g noisy sections of time series aga<strong>in</strong>st the signals. In first<br />

<strong>in</strong>stance, characteristics of long period (4-128sec, band 4). MT time series were used to<br />

build tra<strong>in</strong><strong>in</strong>g / test<strong>in</strong>g <strong>data</strong> base for artificial neural networks.<br />

5.2 Signal and noise <strong>in</strong> the magnetotelluric time series<br />

An exact classification of signal and noise characteristics <strong>in</strong> magnetotelluric time series<br />

is difficult as their sources may have similar spectral content. Unfortunately there are<br />

often more noise sources than signal sources. Still, a generalized classification is possible<br />

accord<strong>in</strong>g to the orig<strong>in</strong>s of both signal and noise. At periods longer than 1 sec, the natural<br />

electromagnetic field orig<strong>in</strong>ates <strong>in</strong> the upper ionosphere and magnetosphere. Random<br />

bursts of energy orig<strong>in</strong>ate <strong>in</strong> charged particles from the Sun and <strong>in</strong>duce s<strong>in</strong>usoidal electromagnetic<br />

waves <strong>in</strong> the magnetosphere and ionosphere (see § 2.2.1 for more details).<br />

The signal amplitude and frequency may vary with the energy and type of activity and<br />

there may be a long gap between the arrival of two tra<strong>in</strong>s of signal. Figure 5.1(a) shows<br />

some typical magnetotelluric signal patterns. <strong>Magnetotelluric</strong> noise sources have been<br />

listed <strong>in</strong> table 2.1. Any or all the these noise sources may be present together with MT<br />

signal result<strong>in</strong>g <strong>in</strong> compounded responses <strong>in</strong> the recorded time series. Some of the most<br />

common MT noise patterns are presented <strong>in</strong> Figure 5.1(b).<br />

5.3 Visual Inspection (edit<strong>in</strong>g) of magnetotelluric<br />

time series <strong>data</strong>; why automation ?<br />

Manual edit<strong>in</strong>g is often the first step <strong>in</strong> magnetotelluric <strong>data</strong> process<strong>in</strong>g. The editor<br />

exam<strong>in</strong>es each stack of time series and labels it as either good or bad accord<strong>in</strong>g to<br />

its signal/noise character. The bad stacks are removed from further process<strong>in</strong>g. The<br />

task <strong>in</strong>volves an <strong>in</strong>tensive amount of pattern recognition. Experience provides a judicial<br />

balanc<strong>in</strong>g of signal characteristics such as shape, amplitude, frequency and correlation.


69<br />

Figure 5.1: Common signal and noise patterns <strong>in</strong> long period MT time series. Samples<br />

are collected from different sites. Note the change <strong>in</strong> amplitudes. (a) Signal patterns;<br />

Samples of E x and H y shows the geomagnetic pulsations. Other channels also show<br />

signals but at longer periods. (b) Noise patterns; E y & H x shows different types of<br />

spikes. A step and its decay is shown <strong>in</strong> H y . Sample of random noise is shown <strong>in</strong> H z .<br />

This process is subjective <strong>in</strong> nature and the same editor may output differently <strong>in</strong> long<br />

sequences of edit<strong>in</strong>g. A rule of thumb is that ‘if <strong>in</strong> doubt throw it out’. If questioned<br />

about a particular decision, however, the editor may offer a few rules for guidance but<br />

can give no obvious systematic reason<strong>in</strong>g. This constitutes the major time and human<br />

resource used <strong>in</strong> MT <strong>data</strong> process<strong>in</strong>g. The number of stacks of time series recorded<br />

sometimes runs <strong>in</strong>to hundreds. With several such sessions of record<strong>in</strong>gs from a station<br />

and many such MT stations occupied <strong>in</strong> each survey, there is a press<strong>in</strong>g need to provide<br />

a more robust alternative, which is less time consum<strong>in</strong>g and more objective.


70<br />

5.4 <strong>Magnetotelluric</strong> noise characterization<br />

Automation of MT time series edit<strong>in</strong>g requires a systematic evaluation of the task performed<br />

by the editor. What parameters <strong>in</strong>fluence an edit<strong>in</strong>g decision? How much importance<br />

does the editor give to each parameter? Can it be quantified? As there is<br />

no <strong>in</strong>formation available <strong>in</strong> this regard, an edit<strong>in</strong>g evaluation exercise was undertaken.<br />

It <strong>in</strong>volved discussion with different editors and re – <strong>analysis</strong> of previously edited <strong>data</strong><br />

and identification of the <strong>in</strong>fluence of different parameters <strong>in</strong> magnetotelluric time series<br />

edit<strong>in</strong>g. From the study, it was found that the follow<strong>in</strong>g factors reasonably represent the<br />

criteria applied <strong>in</strong> edit<strong>in</strong>g:<br />

1. Signal pattern<br />

2. Signal amplitude<br />

3. Correlation of different field components<br />

4. General noise level<br />

5. Quantity of MT <strong>data</strong> available.<br />

When the general noise level is very high and/or the quantity of <strong>data</strong> available for<br />

edit<strong>in</strong>g is limited, compromises are made <strong>in</strong> edit<strong>in</strong>g. As these are exceptional cases, they<br />

were excluded from the present <strong>analysis</strong>. The first three factors were analyzed and each<br />

was given a percentage of <strong>in</strong>fluence. The percentage of <strong>in</strong>fluence for a particular factor<br />

was found by preferential edit<strong>in</strong>g <strong>us<strong>in</strong>g</strong> that criterion and compar<strong>in</strong>g results with the<br />

regular mode of edit<strong>in</strong>g.<br />

5.4.1 Patterns of signal & noise :<br />

Pattern controls a major part of decision mak<strong>in</strong>g. The signal pattern is transient overlapped<br />

s<strong>in</strong>usoids. The noise patterns can be classified as follows.<br />

1. Spikes : High amplitude, maximum duration for few <strong>data</strong> samples. Sometimes a<br />

spike is followed by a transient decay. These are commonly power related.<br />

2. Noise bursts: Plateau or step-like appearance spann<strong>in</strong>g hundreds of <strong>data</strong> samples.<br />

3. Noisy trace: Random fluctuations due to w<strong>in</strong>d, seismic effects and or low signal.<br />

4. Muted or dead trace: Instrument problem such as broken cable or failure <strong>in</strong> electronics.<br />

In most cases, the quality of signal can be deduced from the signal waveform. This,<br />

accord<strong>in</strong>g to our study <strong>in</strong>fluences 60% of edit<strong>in</strong>g decisions.


71<br />

5.4.2 Amplitude of signals<br />

Although naturally vary<strong>in</strong>g electromagnetic fields exhibit a large variation <strong>in</strong> their<br />

strength, a broad range can be specified. In most cases, the amplitude ratio of orthogonal<br />

electric and magnetic signals was found to be a good discrim<strong>in</strong>ator. In long<br />

period time series, the electric field fluctuates with<strong>in</strong> +/-100 mV/km and the magnetic<br />

field fluctuates with<strong>in</strong> +/-0.5 nT <strong>in</strong> a noiseless environment. The channel amplitude may<br />

be <strong>in</strong>creased many times <strong>in</strong> the presence of certa<strong>in</strong> types of noise. Amplitude criterion<br />

often gives the most reliable <strong>in</strong>formation if the contam<strong>in</strong>ated signal has the same pattern<br />

as noise- free signals but enhanced amplitude. This was found to <strong>in</strong>fluence 20% of the<br />

decisions.<br />

5.4.3 Correlation between simultaneously measured channels<br />

The electric and magnetic fields are related by a transfer function def<strong>in</strong>ed <strong>in</strong> § 3.3.1.2 In<br />

the ideal case the MT signals should be highly correlated and random noise will reduce<br />

the correlation. However, noise can be highly correlated between E and H channels.<br />

Noise from power l<strong>in</strong>es especially near 60/50 Hz is an example. In longer period <strong>data</strong>,<br />

it was found that the correlation coefficients between orthogonal electric and magnetic<br />

fields (E x - H y and E y - H x ) could be a signal discrim<strong>in</strong>ator. Correlation was found to<br />

<strong>in</strong>fluence another 20% of edit<strong>in</strong>g decisions.<br />

As the above-described parameters <strong>in</strong>fluence the bulk of edit<strong>in</strong>g decisions, they were<br />

selected as the basis for automation. An automation scheme was developed <strong>us<strong>in</strong>g</strong> an<br />

artificial neural network for classification/edit<strong>in</strong>g of time series <strong>data</strong>.<br />

5.5 Artificial Neural Networks<br />

5.5.1 Why artificial neural network?<br />

Optimal conventional automation requires statistical characterization of the noise. This<br />

means estimation of certa<strong>in</strong> statistical parameters from the <strong>data</strong>. In other words the<br />

likelihood ratio of signal/noise is replaced by sufficient statistics of <strong>data</strong>. The method is<br />

simple and appeal<strong>in</strong>g. It works very well when the target signal is known and the noise<br />

has a normal distribution. But <strong>in</strong> most cases noise does not have a normal distribution<br />

and the likelihood ratio is a complicated nonl<strong>in</strong>ear function of <strong>in</strong>put <strong>data</strong>. Garth and<br />

Poor [1994] classifies geophysical signals as non-Gaussian, unstructured signals as they<br />

<strong>in</strong>volve least amount of detail. Classification of such signals is best done by label<strong>in</strong>g<br />

them accord<strong>in</strong>g to their signal content and then present<strong>in</strong>g them to a pattern recognition<br />

scheme. Artificial neural networks (ANN) are an emerg<strong>in</strong>g tool that have been applied<br />

<strong>in</strong> many areas of science and eng<strong>in</strong>eer<strong>in</strong>g where pattern recognition is <strong>in</strong>volved, such as<br />

speech and character recognition. The learn<strong>in</strong>g and adaptive capabilities of these models<br />

make them attractive for application to some problems <strong>in</strong> geophysics (Calderon-Macias<br />

et al. [2000]). The most documented application of ANN <strong>in</strong> geophysics has been for the<br />

automation of seismic <strong>data</strong> process<strong>in</strong>g and <strong>in</strong>terpretation (Murat and Rudman [1992],<br />

McCormack et al. [1993], Fish and Kusuma [1994], Dai and MacBeth [1995]). ANN also


72<br />

Figure 5.2: Simple three layer feed forward neural network. The <strong>data</strong> is processed at<br />

each neuron <strong>in</strong> the layers. Each neuron performs a summ<strong>in</strong>g of <strong>in</strong>puts multiplied with a<br />

weight parameter and outputs the <strong>data</strong> through its sigmoid transfer function.<br />

have found other applications <strong>in</strong> geophysics; such as <strong>in</strong> <strong>in</strong>terpretation of well log <strong>data</strong><br />

(W<strong>in</strong>er et al. [1991]) , locat<strong>in</strong>g subsurface targets from electromagnetic field <strong>data</strong> (Poulton<br />

et al. [1992]) prediction of upper atmospheric and ionospheric activities (Lundstedt [1996],<br />

Alt<strong>in</strong>ay et al. [1997], Koons and Gorney [1991]). More recently, artificial neural networks<br />

have been used <strong>in</strong> magnetotelluric <strong>in</strong>version (Zhang and Paulson [1997], Spichak and<br />

Popova [2000]).<br />

5.5.2 ANN theory<br />

An artificial neural network is an <strong>in</strong>formation process<strong>in</strong>g system composed of a large<br />

number of process<strong>in</strong>g elements called neurons, which are modeled on the functions of<br />

neurons <strong>in</strong> the human bra<strong>in</strong>. ANN differ from conventional pattern recognition techniques<br />

<strong>in</strong> their ability to adaptively discrim<strong>in</strong>ate or learn through repeated exposure to examples<br />

and <strong>in</strong> their robustness <strong>in</strong> the presence of high noise levels. ANN do not require a priori<br />

knowledge about the noise distribution of the process under study, as do its statistical<br />

counterparts. Unlike conventional methods, which <strong>in</strong>corporate a fixed algorithm to solve<br />

a particular problem, ANN perform a mapp<strong>in</strong>g, usually non-l<strong>in</strong>ear, between the <strong>in</strong>put<br />

and output <strong>data</strong>, which allows the network to acquire important <strong>in</strong>formation on the<br />

problem be<strong>in</strong>g solved. It is these characteristics of neural networks that motivated us to<br />

<strong>in</strong>vestigate their use <strong>in</strong> MT <strong>data</strong> process<strong>in</strong>g.


73<br />

One of the most widely used types of ANN, the feed forward artificial neural network<br />

(FANN) was used <strong>in</strong> the present study. Its architecture is outl<strong>in</strong>ed <strong>in</strong> Figure 5.2. It<br />

consists of a layer of neurons that accept various <strong>in</strong>puts (<strong>in</strong>put layer). These <strong>in</strong>puts are<br />

fed to further layers of neurons (hidden layers) and ultimately to the output layer, which<br />

produces a response. The aim of the technique is to tra<strong>in</strong> the network such that its<br />

response to a given set of <strong>in</strong>puts is as close as possible to a desired output. A number<br />

of algorithms are available for tra<strong>in</strong><strong>in</strong>g a neural network. Back propagation is the most<br />

popular tra<strong>in</strong><strong>in</strong>g algorithm (Werbos [1990]) and was used <strong>in</strong> the current study. Dur<strong>in</strong>g<br />

FANN tra<strong>in</strong><strong>in</strong>g, each hidden and output neuron process <strong>in</strong>puts by multiply<strong>in</strong>g each <strong>in</strong>put<br />

by its weights. The products are summed and processed <strong>us<strong>in</strong>g</strong> an activation function,<br />

here, a sigmoid function,<br />

f(x) = 1/(1 − e −x ), (5.1)<br />

to produce an output with reasonable discrim<strong>in</strong>at<strong>in</strong>g power. The neural network<br />

learns by modify<strong>in</strong>g the weights of the neurons <strong>in</strong> response to the errors between the<br />

actual and targeted output values. For a given set or vector of N <strong>in</strong>puts (x 1 , x 2 ,....,x n ),<br />

the output of node j is computed as<br />

(∑ )<br />

y j = f Wji x i (5.2)<br />

where W ji is the weight of the connection between the ith and j th neurons. The<br />

learn<strong>in</strong>g rule for the adjustments <strong>in</strong> the weight between neurons i and j is expressed as<br />

∆W ji = ηδ j o i (5.3)<br />

where o i is either the output of node i or an <strong>in</strong>put, η is a positive constant named<br />

the learn<strong>in</strong>g rate and δ j is the error term of node j. Thus,<br />

where<br />

and,<br />

δ j = δE<br />

δ netj<br />

(5.4)<br />

E = 0.5 ∑ (y j − o j ) 2 (5.5)<br />

net j = ∑ W ji o i . (5.6)<br />

Here, y j is the target value for the j th node and o j is the output for the j th node.<br />

The value δ j is computed as,<br />

δ j = o j (1 − o j ) ∑ δ k W kj , (5.7)<br />

if the node is not an output unit. To improve the convergence characteristics, a<br />

momentum ga<strong>in</strong> β is added to the weight correction term which stabilizes oscillations<br />

dur<strong>in</strong>g the learn<strong>in</strong>g process ( Boadu [1998]), i.e.,<br />

∆W ji (n + 1) = ηδ j o i + β∆W ji (n), (5.8)


74<br />

where n is the iteration <strong>in</strong>dex. The tra<strong>in</strong><strong>in</strong>g of the network is complete if the convergence<br />

of weight<strong>in</strong>g coefficients has been achieved. The convergence criterion requires<br />

that the sum square error at the output must be less than a desired tolerable error (Luo<br />

and Unbehauen [1997]).<br />

While tra<strong>in</strong><strong>in</strong>g a network, we may start with arbitrary values for the weights W ji . It<br />

is usual to choose the random numbers <strong>in</strong> the range -1 to 1. Next, the outputs (O) and<br />

errors (E) for that set of weights are calculated, followed by the derivatives of E with<br />

respect to all of the weight (equation 5.7). If <strong>in</strong>creas<strong>in</strong>g a given weight would lead to<br />

more error, the weights are adjusted downwards. If <strong>in</strong>creas<strong>in</strong>g a weight leads to reduced<br />

error, it is adjusted it upwards. After adjust<strong>in</strong>g all the weights up or down, start all over<br />

aga<strong>in</strong> and keep go<strong>in</strong>g through this process until the error is close to zero. The sequence of<br />

present<strong>in</strong>g the entire tra<strong>in</strong><strong>in</strong>g <strong>data</strong>base, calculat<strong>in</strong>g the network response, compar<strong>in</strong>g the<br />

result with the assigned class, propagat<strong>in</strong>g the error backwards and adjust<strong>in</strong>g the weights<br />

is called an epoch. Few thousands of such epochs of tra<strong>in</strong><strong>in</strong>g are usually required by a<br />

neural network to reach zero error. However, if the number of tra<strong>in</strong><strong>in</strong>g patterns exceed<br />

the number of weights <strong>in</strong> the network it may not be possible for the sum squared error<br />

(SSE) to reach zero.<br />

5.6 Data <strong>analysis</strong><br />

5.6.1 Network eng<strong>in</strong>eer<strong>in</strong>g<br />

As a neural network’s massive <strong>in</strong>terconnectivity and <strong>in</strong>herent non-l<strong>in</strong>earity requires significant<br />

comput<strong>in</strong>g resources, dedicated ma<strong>in</strong>frame computers and workstations were<br />

traditionally used for neural network tra<strong>in</strong><strong>in</strong>g. Most of the applications utiliz<strong>in</strong>g ANN<br />

<strong>in</strong> geophysics deal with s<strong>in</strong>gle-channel <strong>data</strong> <strong>us<strong>in</strong>g</strong> a small slid<strong>in</strong>g w<strong>in</strong>dow mov<strong>in</strong>g along<br />

the time series. The current application, where 5 channels of <strong>data</strong>, each conta<strong>in</strong><strong>in</strong>g 256<br />

po<strong>in</strong>ts, were to be classified simultaneously, posed a major challenge. A novel approach<br />

was made to accommodate multi-channel <strong>data</strong> so that all the comput<strong>in</strong>g could be done<br />

on a PC. Figure 5.3 gives a schematic representation of the <strong>data</strong> flow. This method can<br />

be used for any multivariate signal detection scheme.<br />

As the signal shape and pattern controls a major part of the edit<strong>in</strong>g decisions, attention<br />

was focused on this part of the <strong>analysis</strong>. The <strong>in</strong>ter-channel parameters, such<br />

as amplitude ratios and correlation coefficients were also <strong>in</strong>cluded for neural network<br />

tra<strong>in</strong><strong>in</strong>g. The signal detection scheme was divided <strong>in</strong>to two steps.<br />

a) Detection of the patterns of <strong>in</strong>dividual channels: Patterns of <strong>in</strong>dividual channels<br />

were evaluated <strong>us<strong>in</strong>g</strong> a neural network. With<strong>in</strong> each stack (256 po<strong>in</strong>ts) of 5 channels,<br />

each channel’s pattern successively were classified. Thus pattern detection of a s<strong>in</strong>gle<br />

stack resulted <strong>in</strong> 5 values correspond<strong>in</strong>g to the 5 channels.<br />

b) Detection of <strong>in</strong>ter-channel parameters: Amplitude ratios (E x /H y , E y /H x ) and correlation<br />

coefficients between channels were computed. These parameters, along with the<br />

pattern quality <strong>data</strong> from step (a), formed the <strong>in</strong>puts for another neural network. Output<br />

from this network <strong>in</strong>dicates the overall quality of that <strong>data</strong> stack.<br />

This separation provided us with the flexibility and ease of operat<strong>in</strong>g on a small<br />

computer, without hav<strong>in</strong>g to compromise on network performance.


75<br />

Figure 5.3: Data flow through the feed forward artificial neural network (FANN) based<br />

edit<strong>in</strong>g scheme.<br />

5.6.2 Pattern Tra<strong>in</strong><strong>in</strong>g<br />

5.6.2.1 Data used:<br />

Select<strong>in</strong>g an appropriate tra<strong>in</strong><strong>in</strong>g <strong>data</strong> set is one of the most critical steps <strong>in</strong> successful<br />

tra<strong>in</strong><strong>in</strong>g. How many patterns are required <strong>in</strong> tra<strong>in</strong><strong>in</strong>g? How many patterns should be<br />

dom<strong>in</strong>ated by signal and how many by noise? A rule of thumb is that the patterns <strong>in</strong><br />

the tra<strong>in</strong><strong>in</strong>g set should cover the ma<strong>in</strong> categories of signal and noise. The signal pattern<br />

should represent the typical features of a signal with different frequency characteristics.<br />

Obviously there are more noise patterns than signal patterns (Zhao and Takano [1999]).<br />

With an <strong>in</strong>put <strong>data</strong> space of 256, the number of exemplars required to tra<strong>in</strong> a neural<br />

network was large (> 1000 sets). This required us to search for and select a large number<br />

of time series segments. 1200 sets of MT long period time series <strong>data</strong> stacks (256 s.<br />

length, each stack) were collected from different <strong>data</strong> sets giv<strong>in</strong>g 6000 traces.<br />

5.6.2.2 Pre-process<strong>in</strong>g<br />

Each channel was corrected for trend and bias and normalized between 1 and 0. The<br />

<strong>data</strong> segments were then manually classified and assigned a value between 0.9 and 0.1,<br />

depend<strong>in</strong>g on quality. The label varied between 0.1 (bad) and 0.9 (good) depend<strong>in</strong>g on<br />

the pattern quality of the <strong>data</strong>. There is no fixed generic relationship between the label<br />

and quality ( one can as well have labels <strong>in</strong> the reverse order).<br />

Out of the classified time series segments, 5000 extreme cases were selected for tra<strong>in</strong><strong>in</strong>g.<br />

This set conta<strong>in</strong>ed 2500 very good and 2500 very bad <strong>data</strong> samples. Care was taken<br />

to <strong>in</strong>clude all categories of signal and noise patterns generally found <strong>in</strong> MT time series.<br />

The <strong>data</strong>base was shuffled to have a random distribution of good and bad <strong>data</strong> for tra<strong>in</strong><strong>in</strong>g.<br />

From this <strong>data</strong> base 3000 exemplars (tra<strong>in</strong><strong>in</strong>g vectors) were kept for tra<strong>in</strong><strong>in</strong>g and


76<br />

Figure 5.4: Stacked amplitude and phase spectra of the tra<strong>in</strong><strong>in</strong>g <strong>data</strong>base. The spectra<br />

of noisy <strong>data</strong> (class 0.1) clearly different from the signals (class 0.9).<br />

the rest, for test<strong>in</strong>g.<br />

5.6.2.3 FANN Tra<strong>in</strong><strong>in</strong>g<br />

The network was presented with the tra<strong>in</strong><strong>in</strong>g <strong>data</strong>base of 3000 time series segments.<br />

Tra<strong>in</strong><strong>in</strong>g was done with different values of learn<strong>in</strong>g rate (η) and momentum ga<strong>in</strong> (β) for<br />

different epochs (def<strong>in</strong>ed <strong>in</strong> §5.5.2). Tra<strong>in</strong><strong>in</strong>g typically took 60 m<strong>in</strong>utes to complete 1000<br />

epochs on a 500 MHz PC. The longer time for tra<strong>in</strong><strong>in</strong>g hampered effective <strong>in</strong>teraction<br />

with the process and limited an exhaustive search for the optimum network configuration<br />

( i.e.; optimize η, β and the number of hidden neurons). A transform of the <strong>in</strong>put <strong>data</strong><br />

was sought which would preserve the essential <strong>in</strong>formation about the signal while reduc<strong>in</strong>g<br />

the dimensionality of the <strong>in</strong>put space. For this, the time series was Fourier transformed,<br />

after apply<strong>in</strong>g a cos<strong>in</strong>e taper to both ends. To test whether the amplitude spectra alone<br />

could be used as a discrim<strong>in</strong>ator, the spectra of good and bad signals <strong>in</strong> the entire tra<strong>in</strong><strong>in</strong>g<br />

<strong>data</strong>base were stacked separately. As seen <strong>in</strong> Figure 5.4, there is a clear difference between<br />

the two types. Noise generally raises the spectral power at higher frequencies, where as<br />

signal has more power at lower frequencies. Amplitude spectra were used for further<br />

tra<strong>in</strong><strong>in</strong>g. FFT enabled us to reduce the number of <strong>in</strong>put <strong>data</strong> from 256 to 128. Tra<strong>in</strong><strong>in</strong>g<br />

was attempted with different subsets of spectral harmonics by remov<strong>in</strong>g the highest<br />

frequency elements successively. The first 100 harmonics were found to be sufficient<br />

and the tra<strong>in</strong><strong>in</strong>g time was considerably reduced (20 m<strong>in</strong>utes). The sum squared error<br />

(SSE ) as a function of epoch for the last tra<strong>in</strong><strong>in</strong>g is plotted <strong>in</strong> Figure 5.5(a). The SSE<br />

reached a m<strong>in</strong>imum of 33.73 for 2000 tra<strong>in</strong><strong>in</strong>g samples. In neural network tra<strong>in</strong><strong>in</strong>g, more


77<br />

Figure 5.5: Results from the pattern tra<strong>in</strong><strong>in</strong>g. (a) The SSE as a function of epoch. The<br />

error reached the m<strong>in</strong>imum floor after 400 epochs. Stability of convergence is demonstrated<br />

up to 1000 epochs. (b) Deviation between manually classified and network predicted<br />

classes for 500 test time series segments. The scattered po<strong>in</strong>ts shows the major<br />

deviations. The correct pick<strong>in</strong>g constitute 94<br />

importance is given to how the network performs on novel (non tra<strong>in</strong><strong>in</strong>g) <strong>data</strong> than the<br />

SSE of tra<strong>in</strong><strong>in</strong>g itself. FANN on test samples (samples which were not used for tra<strong>in</strong><strong>in</strong>g)<br />

gave 94 % (472/500) correct classification (Figure 5.5(b) with η = 0.09, β = 0.1 and 10<br />

hidden neurons. As observed by Dai and MacBeth [1995], although this solution was<br />

considered optimal for the current application, further architecture optimization could<br />

undoubtedly be achieved by a more exhaustive search procedure on a more powerful<br />

computer.<br />

5.6.2.4 Sensitivity <strong>analysis</strong><br />

The neural network’s sensitivity to the signal to noise ratio was exam<strong>in</strong>ed <strong>us<strong>in</strong>g</strong> the<br />

follow<strong>in</strong>g <strong>analysis</strong>. A time series of 256 po<strong>in</strong>ts with high signal content was mixed with<br />

a normally distributed random noise series. The tra<strong>in</strong>ed neura1 network was assigned to<br />

classify the signal. In each run the signal content was changed with a small <strong>in</strong>crement<br />

so that it covered the range 0% to 100%. The network steadily gave values near to 0<br />

(Figure 5.6) until the signal content reached 60%. The output changes to higher values<br />

as signal content exceeds 60% and asymptotes to 1 after it is 80%. It is now evident that<br />

the network is able to detect signals if the signal content is more than 70%. Another<br />

<strong>in</strong>terest<strong>in</strong>g aspect is the narrow region of high variance (when signal content is between<br />

65% and 75%) <strong>in</strong> the output. A thorough <strong>analysis</strong> with noise of other distributions was<br />

beyond the scope of the present study.


78<br />

Figure 5.6: Network output versus signal content. The network was simulated by <strong>in</strong>puts<br />

with vary<strong>in</strong>g signal content. A narrow region of high variance exists when signal content<br />

is between 65<br />

5.6.3 Inter channel tra<strong>in</strong><strong>in</strong>g<br />

Here, a separate neural network was tra<strong>in</strong>ed with the <strong>in</strong>ter channel parameters (expla<strong>in</strong>ed<br />

<strong>in</strong> §5.6.3). To keep this as the f<strong>in</strong>al step <strong>in</strong> determ<strong>in</strong><strong>in</strong>g the overall quality of the stack,<br />

the 5 pattern quality values predicted by the first step also as <strong>in</strong>put were also <strong>in</strong>cluded.<br />

The <strong>in</strong>puts to the network were,<br />

1. 5 pattern quality values from earlier network<br />

2. 2 amplitude ratios (E x /H y & E y /H x )<br />

3. 2 correlation parameters (E x - H y & E y - H x )<br />

5.6.3.1 The <strong>data</strong><br />

To tra<strong>in</strong> the network, 900 stacks were selected from different <strong>data</strong> sets ( 5 channels,<br />

256 <strong>data</strong> po<strong>in</strong>ts each). Care was taken to <strong>in</strong>clude an equal number of signal and noise<br />

segments. Each stack was manually <strong>in</strong>spected to assign a class vector 0.1 or 0.9 depend<strong>in</strong>g<br />

on its overall signal quality, as expla<strong>in</strong>ed previously. A tra<strong>in</strong><strong>in</strong>g <strong>data</strong>base was made as<br />

follows.<br />

1. The tra<strong>in</strong>ed FANN was used to classify the patterns of the 5 simultaneously measured<br />

channels with<strong>in</strong> each stack. Output of this process<strong>in</strong>g varied between 0.0<br />

(bad) and 1.0 (good). Figure 5.7(a) shows the pattern classes for E x and E y channels<br />

versus the stack numbers. The broken bar below the graph <strong>in</strong>dicates the<br />

manually assigned stack class (Black-good; White-bad). An excellent correlation


79<br />

Figure 5.7: Pattern, amplitude ratios and correlation coefficients of 900 stacks which<br />

form the <strong>data</strong>base for <strong>in</strong>ter channel tra<strong>in</strong><strong>in</strong>g and test<strong>in</strong>g . The thick l<strong>in</strong>e is the runn<strong>in</strong>g<br />

average over 10 po<strong>in</strong>ts. The broken bar <strong>in</strong>dicates the overall stack quality - black good<br />

(0.9) and white bad (0.1). (a) E x (squares) and E y (triangles) pattern quality predicted<br />

as a function of stack number. (b) E x - H y (squares) and E y − H x (triangles) amplitude<br />

ratio. (c) The correlation coefficients of E x to H y (squares) and E y to H x (triangles).<br />

between the manually assigned stack class and the pattern class of <strong>in</strong>dividual channels<br />

computed from FANN process<strong>in</strong>g is evident. This justifies our earlier comment<br />

that signal discrim<strong>in</strong>ation largely depends on the pattern of time series signals.<br />

2. As the amplitude ratios between the orthogonal electric and magnetic field were<br />

found to be another signal discrim<strong>in</strong>ator, the two ratios were <strong>in</strong>cluded <strong>in</strong> the tra<strong>in</strong><strong>in</strong>g<br />

<strong>data</strong>base. Further, <strong>in</strong> order to keep the values with<strong>in</strong> 0 and 1.0 (as necessitated by<br />

the sigmoid function) they were normalized as<br />

A ExHy =<br />

E x<br />

H y<br />

A EyHx =<br />

E x<br />

H y<br />

+ Ey<br />

H x<br />

and (5.9)<br />

E x<br />

H y<br />

E y<br />

H x<br />

+ Ey<br />

H x<br />

(5.10)


80<br />

where, E x , E y , H x and H y are simple ranges of amplitude (maximum - m<strong>in</strong>imum) of<br />

respective channels for a stack. As plotted <strong>in</strong> the Figure 5.7(b) the ratios vary considerably<br />

and a direct correlation with the signal class is impossible. This parameter<br />

was reta<strong>in</strong>ed for tra<strong>in</strong><strong>in</strong>g, as it adds another dimension to the <strong>in</strong>put <strong>data</strong>.<br />

3. Correlation coefficients between orthogonal electric and magnetic fields (E x - H y<br />

and E y - H x ) were calculated for each stack. The correlation coefficient r(τ) (<br />

Molyneux and Schmitt [1999]) between two vectors X and Y ( t = 0,1,2.....n) is<br />

given by<br />

r(τ) =<br />

√<br />

n t=n ∑<br />

t=0<br />

∑<br />

X(t) 2 −<br />

n t=n<br />

t=0<br />

X(t)Y (t + τ) − t=n ∑<br />

( t=n ∑<br />

t=0<br />

)<br />

√<br />

2<br />

X(t)<br />

t=0<br />

n t=n<br />

X(t) t=n ∑<br />

t=0<br />

∑<br />

Y (t + τ) 2 −<br />

t=0<br />

Y (t + τ)<br />

( t=n ∑<br />

t=0<br />

)<br />

(5.11)<br />

2<br />

Y (t + τ)<br />

A value of r =1 <strong>in</strong>dicates perfect positive correlation between two vectors; r = 0<br />

<strong>in</strong>dicates no correlation, mean<strong>in</strong>g the vectors are not similar; r = -1 <strong>in</strong>dicates anticorrelation<br />

(mean<strong>in</strong>g two vectors are of same shape but of opposite polarity). The<br />

correlation coefficients for all the 900 w<strong>in</strong>dows are plotted <strong>in</strong> Figure 5.7(c). The<br />

squares represent E x -H y correlations and diamonds represent the E y -H x correlations.<br />

Most of the E x -H y coefficients are distributed between 0.2 and 0.6 whereas<br />

the E y -H x distribution is between -0.2 and -0.6. It can be clearly seen that the<br />

signal class is = 0.9 (good quality signal) when E x -H y and E y -H x coefficients are<br />

well separated (when they are closer to ± 1).<br />

5.6.3.2 FANN tra<strong>in</strong><strong>in</strong>g<br />

The tra<strong>in</strong><strong>in</strong>g <strong>data</strong>base thus prepared was a 9 x 900 matrix. The rows were shuffled to<br />

produce a random distribution. Around 70% of the <strong>data</strong>base was used for tra<strong>in</strong><strong>in</strong>g the<br />

neural network and the rema<strong>in</strong><strong>in</strong>g 30% was used for test<strong>in</strong>g. As the <strong>in</strong>put vector is only<br />

of length 9, the tra<strong>in</strong><strong>in</strong>g procedure was rather simple compared to the pattern tra<strong>in</strong><strong>in</strong>g. A<br />

three layer feed forward network was tra<strong>in</strong>ed with 9 neurons <strong>in</strong> the <strong>in</strong>put layer, 2 nodes <strong>in</strong><br />

the hidden layer and one node <strong>in</strong> the output layer. The momentum ga<strong>in</strong> β was set to 0.1<br />

and the values of learn<strong>in</strong>g rate (η) and number of hidden layer nodes were changed. Each<br />

tra<strong>in</strong><strong>in</strong>g consisted of 5000 epochs and took 7-10 m<strong>in</strong>utes to complete. As the tra<strong>in</strong><strong>in</strong>g<br />

progressed, the number of hidden layer nodes was reduced to one. The tra<strong>in</strong><strong>in</strong>g stopped<br />

when sum squared error (SSE) was 7.9 for 650 samples with η = 0.09. Figure 5.8(a)<br />

shows SSE as a function of epoch for the f<strong>in</strong>al tra<strong>in</strong><strong>in</strong>g. On simulat<strong>in</strong>g the network <strong>us<strong>in</strong>g</strong><br />

the test <strong>data</strong>, SSE was 3.72 (i.e. 0.1222 per vector ). Figure 5.8((b) shows the deviation<br />

of network-predicted signal class values from the manually classified values. It can be<br />

seen that the network classification was successful.


81<br />

Figure 5.8: Results from <strong>in</strong>ter channel tra<strong>in</strong><strong>in</strong>g. (a) The SSE as a function of tra<strong>in</strong><strong>in</strong>g<br />

epoch. (b) Deviation between manually classified stack quality and network predicted<br />

for 250 stacks. 235 stacks were classified similar to manual classification.<br />

5.6.3.3 Relative significance of <strong>in</strong>put<br />

The network was fully tra<strong>in</strong>ed to classify MT signals. To f<strong>in</strong>d out the relative importance<br />

of the n<strong>in</strong>e <strong>in</strong>puts (5 pattern classes, 2 correlations & 2 amplitude ratios) to the FANN, a<br />

validation tra<strong>in</strong><strong>in</strong>g was carried out. At each run, the particular <strong>in</strong>put of <strong>in</strong>terest was set to<br />

nil all through the run. This resulted <strong>in</strong> a greater error compared to the orig<strong>in</strong>al tra<strong>in</strong><strong>in</strong>g<br />

<strong>us<strong>in</strong>g</strong> non-zero <strong>in</strong>puts. The extent of departure from the previous error is considered<br />

as the relative significance (Alt<strong>in</strong>ay et al. [1997]) of that particular po<strong>in</strong>t for the neural<br />

network. As can be seen <strong>in</strong> Figure 5.9, patterns of E x , H y , E y and H x are dom<strong>in</strong>ant<br />

<strong>in</strong> the neural network response, followed by the correlation parameters. Pattern of H z<br />

and amplitude ratios have the least <strong>in</strong>fluence on the neural network. This roughly agrees<br />

with our earlier classification of the factors <strong>in</strong>fluenc<strong>in</strong>g edit<strong>in</strong>g decisions, but with two<br />

surprises. First is the relative <strong>in</strong>significance of E y as compared to E x . As the <strong>data</strong>base<br />

for tra<strong>in</strong><strong>in</strong>g was equally biased to all the 5 channels, it was not the result of faulty<br />

tra<strong>in</strong><strong>in</strong>g. The excellent performance of the neural network on test <strong>data</strong> also proves this.<br />

Second is the low <strong>in</strong>fluence of amplitude parameters. While formulat<strong>in</strong>g the problem,<br />

the amplitude and correlation were given equal weight, but the tra<strong>in</strong><strong>in</strong>g results disprove<br />

it. As the test<strong>in</strong>g sessions proved the discrim<strong>in</strong>ation capability of ANN, the tra<strong>in</strong><strong>in</strong>g was<br />

stopped.


82<br />

Figure 5.9: Relative significance of various <strong>in</strong>puts to the networks, viz amplitude<br />

ratios (A1 and A2), correlation coefficients (C1 and C2) and five pattern qualities<br />

(E x , E y , H x , H y andH z ). The error deviation aga<strong>in</strong>st each <strong>in</strong>put is a measure of its significance<br />

to the Neural Network.<br />

5.7 Application<br />

The artificial neural network based signal detection scheme was applied to the MT <strong>data</strong><br />

collected from SGT. The neural network approach was successful <strong>in</strong> discrim<strong>in</strong>at<strong>in</strong>g bad<br />

time series segments aga<strong>in</strong>st the good segments <strong>in</strong> majority of the cases. Here the result<br />

of neural network process<strong>in</strong>g from four sites viz, G12, VP12, JN10 and TT8 are presented.<br />

The selections were made to demonstrate the performance of neural network <strong>in</strong> the presence<br />

of vary<strong>in</strong>g degree of noise contam<strong>in</strong>ation. Station TT8 situated <strong>in</strong> the <strong>in</strong>dustrial belt<br />

(§ 4.7) is most affected by noise, followed by JN10 and VP12. G12 is a station occupied<br />

<strong>in</strong> the Western India, with very high signal-to-noise ratio <strong>in</strong> the longer period. This was<br />

<strong>in</strong>cluded to demonstrate neural network’s performance on a very good site as well. The<br />

long period <strong>data</strong> collected <strong>in</strong> the chief session (§ 4.3.1) of each site were subjected to three<br />

type of edit<strong>in</strong>g viz 1) Bl<strong>in</strong>d edit<strong>in</strong>g - selected all the stacks, 2) ANN based edit<strong>in</strong>g scheme<br />

and 3) Manual edit<strong>in</strong>g by a third person. Once edited, the <strong>data</strong> were subjected to a<br />

common process<strong>in</strong>g procedure as detailed below. Auto and cross spectra were computed<br />

for each target frequencies (Figure 3.3) from all the available time segments, after the<br />

edit<strong>in</strong>g. Transfer functions were estimated (§ 3.3.1.3) for each time segments, for all the<br />

target frequencies. Average telluric predicted coherence (equation 4.1) functions were<br />

used to sort the auto and cross spectra. A subset consists of 70% of the highest coherent<br />

spectra sets were the selected and used for estimat<strong>in</strong>g the f<strong>in</strong>al transfer functions. Note<br />

that this method differs from ’Coherency Threshold’ methods, where<strong>in</strong> <strong>data</strong> sets with<br />

predicted coherence below a preset value are rejected for f<strong>in</strong>al transfer function estimation.<br />

This choice of somewhat old-fashioned algorithm is justified as it allowed to show<br />

the efficacy of neural network based edit<strong>in</strong>g. Moreover, the algorithm does not <strong>in</strong>terfere<br />

too much with the <strong>data</strong> selected. Variance of the MT transfer function was computed<br />

accord<strong>in</strong>g to equation 3.41 <strong>in</strong> § 3.3.4. General descriptions of results from each site are<br />

given below.


83<br />

Figure 5.10: Comparison of MT apparent resistivity and phase computed from different<br />

mode of edit<strong>in</strong>g of <strong>data</strong> from site G12 . Filled circles represent xy and diamonds represent<br />

yx components. (a) Us<strong>in</strong>g all stacks available. (b) By neural network edit<strong>in</strong>g. (c) By<br />

manual edit<strong>in</strong>g.<br />

1. G12: The station was located over basaltic prov<strong>in</strong>ce of western India. 192 stacks<br />

were available for process<strong>in</strong>g. The time series was generally noise free with long<br />

period geomagnetic pulsations (s<strong>in</strong>usoids). The occasional spikes were smaller than<br />

the waveforms with<strong>in</strong> which they occur. Figure 5.10(a) shows MT apparent resistivity<br />

and phase computed from all available stacks (192). The curve is quite smooth<br />

and without deviation, suggest<strong>in</strong>g that the time series were relatively noise free.<br />

Neither FANN based edit<strong>in</strong>g (selected 157 stacks) nor manual edit<strong>in</strong>g (selected 150<br />

stacks) improved the curve significantly. Results are given <strong>in</strong> Figures 5.10(b) and<br />

(c). A small number of noisy segments were easily rejected by the coherency-based<br />

estimator, without any need of edit<strong>in</strong>g.<br />

2. VP12: This station was located <strong>in</strong> the granulite prov<strong>in</strong>ce of South India. A total<br />

of 352 stacks were recorded. Almost half of the stacks carried spikes and step<br />

like features orig<strong>in</strong>ated from submersible electric pumps and switch<strong>in</strong>g of power<br />

supplies. The MT apparent resistivity and phase computed from all the stacks are<br />

given <strong>in</strong> Figure 5.11(a). The resistivity (ρ) values, especially between 1.0 and 0.1 Hz<br />

are scattered and phase (φ) is poorly resolved, with both xy and yx modes be<strong>in</strong>g<br />

equally affected. Figure 5.11(b) shows the results of FANN-based edit<strong>in</strong>g ( 140<br />

stacks selected ), where, both apparent resistivity and phase are better resolved.<br />

Improvement is clearly evident <strong>in</strong> the 1.0 to 0.1 Hz range. Manual edit<strong>in</strong>g resulted<br />

<strong>in</strong> select<strong>in</strong>g 127 stacks and the computed apparent resistivity and phase are similar<br />

to FANN edit<strong>in</strong>g (Figure 5.11(c)).<br />

3. JN10: This site was also located <strong>in</strong> the same region as VP12, but with more noise<br />

<strong>in</strong> the <strong>data</strong>. The effect of the noise is evident <strong>in</strong> the apparent resistivity curves


84<br />

Figure 5.11: Comparison of MT apparent resistivity and phase computed from different<br />

mode of edit<strong>in</strong>g of <strong>data</strong> from site VP12 . Filled circles represent xy and diamonds<br />

represent yx components. (a) Us<strong>in</strong>g all stacks available. (b) By neural network edit<strong>in</strong>g.<br />

(c) By manual edit<strong>in</strong>g.<br />

Figure 5.12: Comparison of MT apparent resistivity and phase computed from different<br />

mode of edit<strong>in</strong>g of <strong>data</strong> from site JN10 . Filled circles represent xy and diamonds represent<br />

yx components. (a) Us<strong>in</strong>g all stacks available. (b) By neural network edit<strong>in</strong>g. (c) By<br />

manual edit<strong>in</strong>g.


85<br />

Figure 5.13: Comparison of MT apparent resistivity and phase computed from different<br />

mode of edit<strong>in</strong>g of <strong>data</strong> from site TT08 . Filled circles represent xy and diamonds<br />

represent yx components. (a) Us<strong>in</strong>g all stacks available. (b) By neural network edit<strong>in</strong>g.<br />

(c) By manual edit<strong>in</strong>g.<br />

computed from all the stacks (Figure 5.12(a)). The phase is also affected <strong>in</strong> the<br />

range 1- 0.1 Hz. The neural edit<strong>in</strong>g picked 80 stacks out of 256 available and gave a<br />

better estimate, as shown <strong>in</strong> Figure 5.12(b). A less rigorous manual edit<strong>in</strong>g picked<br />

107 stacks and gave a similar result (Figure 5.12(c)).<br />

4. TT8: This station was the most affected by noise. Both electric channels, especially<br />

E x , were affected by noise orig<strong>in</strong>at<strong>in</strong>g from an <strong>in</strong>dustrial belt nearby. Increased<br />

spike activity was observed <strong>in</strong> all channels. The MT apparent resistivity and phase<br />

computed (Figure 5.13(a)) from all the 256 stacks available gave a very distorted<br />

picture. Both xy and yx components were poorly resolved. Significant improvement<br />

was made by FANN edit<strong>in</strong>g (Figure 5.13(b)) which selected 67 stacks out of 256.<br />

The yx component is now smooth and relatively error free. The xy component is<br />

also improved except near 0.1 Hz. Almost the same result is produced by manual<br />

edit<strong>in</strong>g (69/256) as shown <strong>in</strong> Figure 5.13(c). The stacks picked by neural network<br />

and manual edit<strong>in</strong>g are compared <strong>in</strong> the Figure 5.14. Overall the pick<strong>in</strong>gs match<br />

quite well. Deviation between the two edit<strong>in</strong>g schemes is evident for a few picks<br />

between stack numbers 150 - 180. Between these there are stacks with moderate to<br />

low signal content, which were accepted by manual edit<strong>in</strong>g but rejected by neural<br />

edit<strong>in</strong>g.


86<br />

Figure 5.14: Comparison of manual and neural signal pick<strong>in</strong>g for site TT8. The diamonds<br />

present the neural pick<strong>in</strong>g and crosses, the manual.<br />

5.8 Discussion<br />

The application of ANN based edit<strong>in</strong>g to magnetotelluric time series br<strong>in</strong>gs out some<br />

<strong>in</strong>terest<strong>in</strong>g results. The ANN based noise rejection predicts an overall quality of each<br />

subsection of time series (0 to 1). While this value cannot be treated as errors of the correspond<strong>in</strong>g<br />

impedances, the response is a measure of quality of the time series/ impedances<br />

and help improve the process of edit<strong>in</strong>g of <strong>data</strong> more objectively as compared to manual/<br />

visual edit<strong>in</strong>g. In a low noise environment such as station G12, the network edit<strong>in</strong>g<br />

produces results almost similar to bl<strong>in</strong>d edit<strong>in</strong>g (<strong>us<strong>in</strong>g</strong> all stacks). On such a <strong>data</strong>, a<br />

simple coherency-based estimator can do the signal discrim<strong>in</strong>ation to a certa<strong>in</strong> level of<br />

satisfaction. However, the neural network’s ability to pick out signal from moderate to<br />

high noise environment was evident on the <strong>data</strong> collected from SGT. In such cases it approximates<br />

human <strong>in</strong>telligence - established from the fact that the neural network based<br />

edit<strong>in</strong>g gives a result similar to manual edit<strong>in</strong>g. These results satisfy the objective of the<br />

present chapter, ie, to provide a robust alternative to manual edit<strong>in</strong>g of magnetotelluric<br />

time series. This experiment, brought out ANN’s potential to automate MT time series<br />

edit<strong>in</strong>g, efficiently and reliably. The scheme could be adapted <strong>in</strong>to rout<strong>in</strong>e process<strong>in</strong>g to<br />

save human time and <strong>in</strong>crease reliability <strong>in</strong> edit<strong>in</strong>g. The current scheme performs about<br />

70,000 float<strong>in</strong>g po<strong>in</strong>t operations (flops) per stack for classification. FFT of five channels<br />

alone uses 40,000 flops. If the further computation uses the same spectra, some of<br />

the computational redundancy can be removed. It po<strong>in</strong>ts to the possibility of <strong>in</strong>tegrat<strong>in</strong>g<br />

ANN based edit<strong>in</strong>g <strong>in</strong>to real time process<strong>in</strong>g of MT <strong>data</strong>. The signal and noise characteristics<br />

<strong>in</strong> magnetotelluric <strong>data</strong> are very different <strong>in</strong> different frequency ranges. This is due<br />

to the difference <strong>in</strong> signal and noise source mechanism <strong>in</strong> different part of Earth’s natural<br />

electromagnetic spectrum. Sensor geometry and <strong>in</strong>strumentation also affect pattern of<br />

signal, which the neural network depends on. This necessitates reformulation of MT signal<br />

and noise characteristics for tra<strong>in</strong><strong>in</strong>g of ANN, wherever necessary. However, the extra<br />

computational requirement by re-tra<strong>in</strong><strong>in</strong>g of ANN may not pose a burden on resources,<br />

tak<strong>in</strong>g <strong>in</strong>to consideration the ever-<strong>in</strong>creas<strong>in</strong>g computational power of microprocessors.<br />

The neural network based approach described <strong>in</strong> this chapter see the time vary<strong>in</strong>g


87<br />

processes for a f<strong>in</strong>ite duration as a whole. It does not assume a transfer function relation<br />

between the different channels under preview. Noise processes, which will not affect the<br />

pattern, amplitude and correlation parameters of the time series, but have potential to<br />

badly affect the estimation of MT transfer functions, may still escape from the screen<strong>in</strong>g<br />

of the neural network based edit<strong>in</strong>g. Secondly, once the time series has been selected,<br />

the signal content at all the frequencies from that time segment may not be the same.<br />

This provides for and necessitates a second screen<strong>in</strong>g of MT <strong>data</strong>, now <strong>in</strong> the doma<strong>in</strong><br />

of frequency, after an <strong>in</strong>itial <strong>data</strong> screen<strong>in</strong>g by ANN. One advantage of work<strong>in</strong>g <strong>in</strong> the<br />

frequency doma<strong>in</strong> is that estimation is done <strong>in</strong>dependently at many frequencies. For<br />

stationary processes, the <strong>data</strong> at different frequencies are strictly uncorrelated (Chave<br />

and Thomson [2003]). The robust process<strong>in</strong>g techniques effectively search the frequency<br />

space to estimate MT transfer functions and its superiority above other comparable<br />

methods <strong>in</strong> frequency doma<strong>in</strong> is established (Jones et al. [1989]). However, as stated <strong>in</strong><br />

the <strong>in</strong>troduction, <strong>in</strong> certa<strong>in</strong> cases of noises, robust process<strong>in</strong>g fails. The next chapter<br />

critically analyze and propose improvements to the robust process<strong>in</strong>g techniques for MT<br />

transfer function estimation.


Chapter 6<br />

Estimation of <strong>Magnetotelluric</strong><br />

Transfer Functions: Robust<br />

Statistical Methods<br />

88


89<br />

6.1 Introduction<br />

After the raw time series have been <strong>in</strong>spected by manual or automated noise rejection<br />

methods, the selected segments are used for estimat<strong>in</strong>g MT transfer functions. This <strong>in</strong>volves<br />

estimat<strong>in</strong>g 10 1 -10 2 complex frequency doma<strong>in</strong> transfer function elements Z(ω)<br />

from electric and magnetic field time series E(t) and H(t) (approximately 10 6 real numbers<br />

per site) (Egbert and Livelybrooks [1996]). This <strong>data</strong> reduction, though superficially<br />

simple, can result <strong>in</strong> useless MT transfer functions, <strong>in</strong> the presence of noise <strong>in</strong> measurements.<br />

The classical least-square method of comput<strong>in</strong>g MT transfer functions, allow<strong>in</strong>g<br />

for noise distributed <strong>in</strong> the simplest manner was discussed <strong>in</strong> § 3.3.1. However, the drawbacks<br />

of least-square (LS) method has been widely recognized and documented <strong>in</strong> the<br />

past three decades (Sims et al. [1971], Gamble et al. [1979], Egbert and Booker [1986],<br />

Egbert and Livelybrooks [1996]). The failure of the LS method has been attributed to 1)<br />

presence of noise <strong>in</strong> the ‘<strong>in</strong>put’ channel and 2) violations of Gaussian noise assumptions.<br />

In the first <strong>in</strong>stance, the l<strong>in</strong>ear statistical model, (equation 3.26) through which the<br />

natural electromagnetic fields are related to each other, considers the <strong>in</strong>put as noise free<br />

and the noise is restricted to the output or ‘predicted’ <strong>data</strong>. In MT it is usual to assume<br />

the magnetic fields as <strong>in</strong>put and electrical fields as output. It follows that the noise <strong>in</strong><br />

magnetic fields can down bias the transfer function estimates ( see § 3.3.2). To avoid these<br />

bias errors, Gamble et al. [1979] proposed the measurement of two remote magnetic field<br />

components as references and remote reference process<strong>in</strong>g substantially improved (Jones<br />

et al. [1989], Shalivahan and Bhattacharya [2002]) MT transfer functions, over the s<strong>in</strong>gle<br />

station least squares approach. In the second case, Gaussian distribution of errors is<br />

assumed by the LS estimator. The observed errors <strong>in</strong> magnetotelluric <strong>data</strong> often have a<br />

Gaussian distribution, but with heavier tails due to the presence of outliers (abnormal<br />

<strong>data</strong>). These non-Gaussian noise produces scatter (or low po<strong>in</strong>t to po<strong>in</strong>t cont<strong>in</strong>uity) <strong>in</strong><br />

the processed <strong>data</strong>. If ignored, these outliers can corrupt the estimated magnetotelluric<br />

transfer functions, mak<strong>in</strong>g them useless for geologic <strong>in</strong>terpretation. A number of process<strong>in</strong>g<br />

methods have been proposed which adaptively weights or screen the <strong>data</strong>. Stodt<br />

[1983] showed the usefulness of weight<strong>in</strong>g the subsections of MT <strong>data</strong> accord<strong>in</strong>g to their<br />

predicated coherence (see § 3.3.4). However, such approaches may fail <strong>in</strong> the presence of<br />

correlated noise <strong>in</strong> the <strong>data</strong>. Substantial improvements were bought out by the application<br />

of robust – M (Hampel et al. [1986]) statistical procedures to MT and geomagnetic<br />

time series <strong>analysis</strong> (Egbert and Booker [1986], Chave and Thomson [1989], Chave et al.<br />

[1987], Larsen [1989], Sutarno and Vozoff [1991], Egbert and Livelybrooks [1996], Egbert<br />

[1997], Ritter et al. [1998], Nagarajan [1998], Smirnov [2003], Chave and Thomson<br />

[2003]). The success of robust procedures may be attributed to three factors. First, its<br />

superiority to other <strong>data</strong> process<strong>in</strong>g techniques is established (Jones et al. [1989]). Second,<br />

these procedures can be justified rigorously (Egbert and Livelybrooks [1996]) and<br />

third it can be easily implemented <strong>us<strong>in</strong>g</strong> iterative-weighted LS procedures and extended<br />

to remote reference process<strong>in</strong>g (Chave and Thomson [1989]).<br />

In this chapter, two new approaches are proposed to improve the performance of<br />

robust statistical procedures on MT time series. The basics of the robust M procedures are<br />

discussed <strong>in</strong> § 6.2 and more stress is given to the computational aspects. Non-parametric<br />

estimators such as Jackknife (Efron [1982]) were used to robustly compute the variance


90<br />

of MT transfer functions (Chave and Thomson [1989]). Its use as an effective <strong>in</strong>itial guess<br />

for robust procedures is discussed <strong>in</strong> § 6.2. It is shown that <strong>in</strong> majority of the cases, the<br />

use of Jackknife for <strong>in</strong>itial guess resulted <strong>in</strong> better estimation of MT transfer function as<br />

compared to LS estimations. It is usual <strong>in</strong> MT to sub divide the time series, estimate<br />

the spectral density matrices for each segment <strong>in</strong>dividually and then robustly average<br />

the spectra or transfer functions between the sub segments (section averag<strong>in</strong>g). With<strong>in</strong> a<br />

segment, it is common to use a limited number of target frequencies and obta<strong>in</strong> smooth<br />

spectra by averag<strong>in</strong>g several adjacent Fourier harmonics (frequency band averag<strong>in</strong>g).<br />

The documented researches on robust estimation of MT spectral densities and transfer<br />

functions concentrate on section averag<strong>in</strong>g. This arises from the assumption that, with<strong>in</strong> a<br />

narrow frequency band, the distribution of Fourier coefficients are of Gaussian nature and<br />

a simple average (LS) gives the best estimate. In § 6.3 it is shown that this argument often<br />

fails and the problem of contam<strong>in</strong>ation is applicable to band averag<strong>in</strong>g as well. A robust<br />

weight<strong>in</strong>g approach is proposed for estimation of cross and auto spectral estimation with<strong>in</strong><br />

a band, without mak<strong>in</strong>g specific model assumptions concern<strong>in</strong>g signal or noise. Both these<br />

proposed procedures, while applied on a large volume of MT <strong>data</strong> collected over SGT,<br />

South India, met with moderate to good improvement of MT transfer functions. Majority<br />

of these <strong>data</strong> were collected <strong>in</strong> s<strong>in</strong>gle station mode, except for few stations, where remote<br />

reference <strong>data</strong> were available. A synchronization problem of MT equipment and GPS was<br />

the reason for s<strong>in</strong>gle station record<strong>in</strong>g dur<strong>in</strong>g the study (§ 4.3.1). However, the robust<br />

process<strong>in</strong>g procedures outl<strong>in</strong>ed <strong>in</strong> this thesis can easily be extended to remote reference<br />

process<strong>in</strong>g and such an example is shown for station VP10 (§ 6.3.5). The application of<br />

the proposed robust process<strong>in</strong>g methods are discussed <strong>in</strong> § 6.2.5 and § 6.4.<br />

6.2 Robust estimation of MT transfer functions<br />

The term ‘robust’ was co<strong>in</strong>ed <strong>in</strong> statistics by G.E.P. Box <strong>in</strong> 1953 (Hampel et al. [1986],<br />

Flannery et al. [1992]). Various def<strong>in</strong>itions are possible for a robust procedure. But<br />

<strong>in</strong> general referr<strong>in</strong>g to a statistical estimation like MT transfer function, it means ‘one<br />

which is relatively <strong>in</strong>sensitive to the presence of a moderate amount of bad <strong>data</strong> or to<br />

<strong>in</strong>adequacies <strong>in</strong> the statistical model and that reacts gradually rather than abruptly to<br />

perturbations of either’ (Jones et al. [1989]).<br />

6.2.1 Why robust methods?<br />

LS estimators are best on a <strong>data</strong> with Gaussian (normal) distribution of errors. It is well<br />

known that the break down po<strong>in</strong>t of least square (LS, or L2) estimates is zero. Break down<br />

po<strong>in</strong>t describes the smallest percentage of bad <strong>data</strong> that can corrupt an estimate. This <strong>in</strong><br />

turn means that the presence of few outliers can corrupt an LS estimate. The resistance<br />

of median and other L1 estimates (L1 – m<strong>in</strong>imiz<strong>in</strong>g the first power of residuals) to the<br />

presence of outliers is well - documented <strong>in</strong> the geophysical context (for e.g. Claerbout<br />

and Muir [1973]). Simple median has a break down po<strong>in</strong>t of 50 %. This led to the<br />

suggestion that L1 norm can replace L2 norm <strong>in</strong> many geophysical estimation problems.<br />

However, Chave et al. [1987] po<strong>in</strong>ted out major drawbacks with L1 estimator on practical


91<br />

<strong>data</strong> sets. It was shown that L1 estimator requires about 60% more <strong>data</strong> to achieve the<br />

same parameter uncerta<strong>in</strong>ties as L2 estimator. Also, the natural probability distribution<br />

function for L1 is double exponential (Laplace) which make statistical <strong>in</strong>ferences difficult.<br />

This suggests that it is desirable to treat the outliers with<strong>in</strong> the framework of a Gaussian<br />

model, rather than outright abandonment of that model (Chave et al. [1987]). One<br />

way to achieve this is to identify the outliers and process the rema<strong>in</strong><strong>in</strong>g segments of<br />

<strong>data</strong> with usual LS procedures, one such example is the ANN edit<strong>in</strong>g discussed <strong>in</strong> §5.<br />

However, <strong>in</strong> many situations like time series <strong>analysis</strong>, detection of outliers itself becomes<br />

extremely difficult and it may result <strong>in</strong> rejection of all the <strong>data</strong> as well. As a consequence,<br />

robust methods are developed, which can accommodate outliers and still m<strong>in</strong>imize their<br />

<strong>in</strong>fluence. Without go<strong>in</strong>g <strong>in</strong>to detail, the situation <strong>in</strong> which robust methods are desirable<br />

is shown <strong>in</strong> Figure 6.1. The probability distribution function shown <strong>in</strong> Figure 6.1(a) has<br />

heavier tails than expected for Gaussian distribution. Any fluctuation <strong>in</strong> these tails may<br />

lead to <strong>in</strong>accurate estimate of the location of central peak (Flannery et al. [1992]). The<br />

simple l<strong>in</strong>e-fitt<strong>in</strong>g problem, given <strong>in</strong> Figure 6.1(b) shows the <strong>in</strong>fluence of few outliers on<br />

the estimation of the slope of the l<strong>in</strong>e, constra<strong>in</strong>ed to go through the orig<strong>in</strong>. An estimate<br />

that is robust to the presence of the outliers should <strong>in</strong>stead produce a fit, which satisfies<br />

majority of observations. Thus the need for robust estimates is seen.<br />

6.2.2 Robust M estimators<br />

Statisticians have developed various robust statistical estimators. For MT transfer function<br />

estimation, M – estimators are more relevant (M – stands for maximum likelihood)<br />

and are discussed <strong>in</strong> detail here. A brief discussion of development of robust estimation<br />

<strong>in</strong> MT follows. Let us reproduce the MT transfer function equation 3.26,<br />

E = Z H + r (6.1)<br />

where there are N observations so that E and r are N vectors, H is an N X 2 matrix<br />

and Z is a rank two vector. The last variable <strong>in</strong> equation 6.1, r is the difference between<br />

measured and predicted output field (here, Electric) and is the residual parameter to<br />

be m<strong>in</strong>imized. The classical way of solv<strong>in</strong>g MT equation is the least squares technique<br />

(Swift [1986], Sims et al. [1971]), where the square of the residual power <strong>in</strong> equation<br />

6.1 is m<strong>in</strong>imized to yield a solution for Z. Let r i be the residual of the i th observation,<br />

the difference between observation and prediction. The standard L2 method tries to<br />

m<strong>in</strong>imize Σ r 2 i . The M estimators try to reduce the effect of outliers by replac<strong>in</strong>g the<br />

squared residual r 2 i by another function of residuals, yield<strong>in</strong>g, m<strong>in</strong> Σ ρ(r i ), where ρ is a<br />

symmetric, positive def<strong>in</strong>ed function, called loss function with a unique m<strong>in</strong>imum at zero.<br />

For standard L2, ρ( r i ) = r 2 i /2, while for the L1 estimator ρ( r i ) = | r i |. In general, if<br />

ρ(r) is chosen to be –log f(r), where f(r) is the true probability density function (pdf), of<br />

the residuals, then the M estimator is maximum likelihood (Chave and Thomson [1989]).<br />

However, as it is difficult to obta<strong>in</strong> a pdf from f<strong>in</strong>ite observations, the loss function is<br />

chosen <strong>in</strong> theoretical ground. Perform<strong>in</strong>g m<strong>in</strong>imization,<br />

∑<br />

ψ(r i )z ij = 0, for j = 1, 2. (6.2)<br />

i


Figure 6.1: Examples where robust statistical methods are desirable: (a) A one dimensional<br />

distribution with heavy tails (b) A distribution <strong>in</strong> two dimensions fitted to straight<br />

l<strong>in</strong>es. Adapted from Flannery et al. [1992]<br />

92


93<br />

Type Loss- ρ(r) Influence ψ(r) Weight w(r)<br />

L2<br />

r 2 2<br />

R 1<br />

L1 |r| sign(r)<br />

1<br />

|r|<br />

Huber 1) if |r| < k<br />

x 2<br />

2<br />

R 1<br />

Huber 2) if |r| ≥ k k = (|r| − k 2<br />

sign(r)<br />

k<br />

|r|<br />

Tukey 1) if |r| < k<br />

k 2<br />

6 (1 − [1 − ( r k )2 ] 3 ) r[1 − ( r k )2 ] 2 [1 − ( r k )2 ] 2<br />

Tukey 2) if |r| ≥ k<br />

k 2<br />

6<br />

0 0<br />

Table 6.1: Few commonly used <strong>in</strong>fluence functions. Adapted from Zhang [1996].<br />

Where ψ(r )= ∂ρ( r)/ ∂(r), is called the <strong>in</strong>fluence function and z ij is one component<br />

of the 2 X 2 tensor Z. The equations will reduce to least squares, when ρ( r) = r 2 /2,<br />

ψ(r )=r or to least absolute deviations when ρ( r) =|r|, ψ(r) = sign(r). To normalize<br />

the equations such as 6.2, it is common to divide the equation with a robust estimate of<br />

scale for which median absolute deviation (MAD) is a good choice (Sutarno and Vozoff<br />

[1991]).<br />

d = med|r i − med(r i )|<br />

σ MAD (6.3)<br />

Where σ MAD is the theoretical counter part of an appropriate pdf. Now equation 6.2<br />

becomes,<br />

∑<br />

i<br />

ψ( r i<br />

d )z ij = 0. (6.4)<br />

To solve this equation, it is easiest to write it as a weighted least square, by def<strong>in</strong><strong>in</strong>g<br />

a weight<strong>in</strong>g function, w i = ψ(r i /d)/r i and rewrit<strong>in</strong>g, equation 6.4,<br />

∑<br />

w i r i z ij = 0, for j = 1, 2. (6.5)<br />

i<br />

The weights are computed based on the residual r and scale parameter d, from the previous<br />

iteration and they are <strong>in</strong>itialized <strong>us<strong>in</strong>g</strong> a least square solution (Chave and Thomson<br />

[1989]). The major difference between the solution given above and a normal weighted<br />

LS procedure is that, <strong>in</strong> the first case the weights are computed based on the residuals<br />

and scale estimate from previous iteration, where as <strong>in</strong> later case, weights are computed<br />

based on the <strong>data</strong> itself.<br />

6.2.3 Choice of <strong>in</strong>fluence functions<br />

The <strong>in</strong>fluence function ψ(r) measures the <strong>in</strong>fluence of a datum on the value of the parameter<br />

estimate. We have already seen the <strong>in</strong>fluence function for L1 and L2 estimates. Given<br />

<strong>in</strong> the table 6.1 are four different <strong>in</strong>fluence functions. (Zhang [1996]). Their graphical<br />

representation is given <strong>in</strong> Figure 6.2.<br />

From the table 6.1 and plot (6.2) it may be deduced that,


94<br />

Figure 6.2: Schematic diagram show<strong>in</strong>g loss, <strong>in</strong>fluence and weight functions for Least<br />

Square (LS or L2) , Least Absolute (L1) Huber and Tukey estimators. Values shown <strong>in</strong><br />

y axis are arbitrary See text for discussion. Adapted from Zhang [1996]<br />

1. L2 (least-squares) estimators are not robust because their <strong>in</strong>fluence function is not<br />

bounded. The larger the residual, the heavier weight it gets.<br />

2. L1 (absolute values) estimators are not stable and their weight function at r=0 is<br />

unbounded and solution may become undeterm<strong>in</strong>ed.<br />

3. Huber function is a parabola <strong>in</strong> the vic<strong>in</strong>ity of zero (like L2) and <strong>in</strong>creases l<strong>in</strong>early<br />

at a given level |r| >k (like L1). Where k = 1.5d gives 95% efficiency with Gaussian<br />

<strong>data</strong>. This is most widely used and suitable to residuals drawn from a probability<br />

distribution that is Gaussian <strong>in</strong> the center and Laplacian <strong>in</strong> the tails. This blend<br />

of L2 and L1, is more suitable for the distribution presented <strong>in</strong> Figure 6.1.<br />

4. Tukey’s function is very severe for outliers (Figure 6.2)<br />

However, use of either Huber or Tukey function alone is not advisable. The Huber<br />

weights fall off slowly for large residuals and never descend to zero and thus do not<br />

provide adequate protection aga<strong>in</strong>st large outliers (Chave and Thomson [1989]). Tukey’s<br />

<strong>in</strong>fluence function outputs near zero values for slightly higher residuals. Used alone, it<br />

may result <strong>in</strong> rejection of all <strong>data</strong>, or acceptance of very few. Hence it is advisable to<br />

use a Huber function for the first few iterations and then use Tukey’s function as a f<strong>in</strong>al<br />

weight<strong>in</strong>g to protect aga<strong>in</strong>st large outliers (Egbert and Booker [1986]).<br />

The response of <strong>in</strong>fluence functions to a set of MT <strong>data</strong> is presented <strong>in</strong> Figure 6.3.<br />

The <strong>data</strong> (VP13) were collected <strong>in</strong> the period range 4 sec to 128 sec at a sampl<strong>in</strong>g rate<br />

1Hz. The time series were sub segmented <strong>in</strong>to sections of 2048 <strong>data</strong> po<strong>in</strong>ts each and 105<br />

such overlapped sections were available for process<strong>in</strong>g. Figure 6.3(a) shows the E x -E xp<br />

(def<strong>in</strong>ed <strong>in</strong> § 3.3) residuals at frequency 0.1875 Hz for each stack. Majority of the residuals<br />

assumes small and similar values with a small amount of scatter. However, there are<br />

large outliers present (∼16), especially near stacks 20, 40 and 100. The weight functions<br />

computed accord<strong>in</strong>g to Huber, L1 and Tukey functions for these residuals are presented<br />

<strong>in</strong> Figure 6.3(b). L2 weights are always 1. L1 weights (<strong>in</strong>verted triangles) show low values<br />

for the outliers. However, their weights become very large for very small residuals. Huber<br />

weights (stars) are comb<strong>in</strong>ations of L1 and L2 weights, with a cut off at k = 1.5d. They


95<br />

Figure 6.3: Responses of different <strong>in</strong>fluence functions to a set of residuals from MT <strong>data</strong><br />

process<strong>in</strong>g. Station VP13 Shows the Ex residuals for 0.1875 Hz for all the 105 stacks. (b)<br />

Shows the response of three <strong>in</strong>fluence functions to the residuals. See text for discussion.<br />

are less <strong>in</strong>fluenced by the large outliers (as L1) and can accommodate the fluctuations<br />

with<strong>in</strong> small residuals (near the Gaussian peak, like L2). As seen for values between<br />

stacks 50 and 60, Huber function down weights the large residuals, where the <strong>in</strong>fluence is<br />

same as L1. The steady response of Huber weights is more clearly demonstrated between<br />

stacks 60 and 80. Tukey weights show very low weights (< 10 −4 ) for large outliers. If<br />

used alone, it may result <strong>in</strong> deselect<strong>in</strong>g most of the <strong>data</strong> sets. Severity of Tukey’s weight<br />

dom<strong>in</strong>ate everyth<strong>in</strong>g else, as demonstrated at stacks 50-60, 100 to 105. However, it also<br />

rejects usable <strong>data</strong> especially at stacks 10, 40-50 and 70.<br />

6.2.4 Implementation for MT<br />

The application of robust process<strong>in</strong>g developed for the estimation of magnetotelluric<br />

transfer functions is presented <strong>in</strong> this section. A MATLAB [2001] code ‘robspm.m’ was<br />

realized by <strong>in</strong>tegrat<strong>in</strong>g the algorithms discussed by Egbert and Booker [1986], Chave<br />

and Thomson [1989], Ritter et al. [1998], with the two new approaches <strong>in</strong>troduced <strong>in</strong> this<br />

chapter. All the computations are performed <strong>in</strong> frequency doma<strong>in</strong>. It is assumed that the<br />

time series were collected for sufficient duration of time. It is then divided <strong>in</strong>to segments<br />

(subsets / stacks) of fixed length. Size of subset is chosen based on the lowest frequency of<br />

<strong>in</strong>terest and a target value for degrees of freedom. Each segment was tapered by a Hann<strong>in</strong>g<br />

w<strong>in</strong>dow. The segments may be overlapped, provided corrections to dof are made. After<br />

Fourier transform, each channel was divided by calibration transfer function (§ 1.3.4)<br />

to remove the effect of <strong>in</strong>strument and sensors. A matrix of auto and cross spectra<br />

between channels (5 x 5 for s<strong>in</strong>gle station set up), was made for a particular frequency<br />

by band averag<strong>in</strong>g the adjacent frequency po<strong>in</strong>ts respect<strong>in</strong>g a narrow frequency w<strong>in</strong>dow


96<br />

(this averag<strong>in</strong>g is critically analyzed <strong>in</strong> § 6.3). The preprocess<strong>in</strong>g results <strong>in</strong> generation of<br />

a 4 dimensional matrix for a site, with size L segments, N frequency and 5 x 5 channels.<br />

This forms the basic <strong>data</strong> for robust estimation. A flow chart for the robust process<strong>in</strong>g<br />

implemented <strong>in</strong> this thesis is shown <strong>in</strong> Figure 6.8(a) and discussed towards end of this<br />

section<br />

6.2.4.1 Initial guess of transfer function<br />

One component of equation 6.1 may be written as,<br />

E l x = Z x x l H l x + Z x y l H l y + r l x (6.6)<br />

Where E and H are frequency doma<strong>in</strong> components of electromagnetic field recorded<br />

<strong>in</strong> l = 1,2,. . . .L time segments. Z xx and Z xy are components of Z (equation 6.1). The<br />

algorithm described here will use the above equations. However, the same steps can be<br />

used to estimate the other components of the tensor and also magnetic transfer functions<br />

(equation 3.24). Estimations are done <strong>in</strong>dependently for all the N frequencies, so that<br />

the frequency terms have been omitted. To get an <strong>in</strong>itial guess for transfer functions,<br />

’global’ spectral matrix is obta<strong>in</strong>ed by averag<strong>in</strong>g all the L cross and auto spectra sets for<br />

a particular frequency, <strong>in</strong> the least square sense. For example one element of the global<br />

5 x 5 matrix is obta<strong>in</strong>ed by,<br />

〈<br />

Ex H ∗ y〉<br />

=<br />

1<br />

L<br />

L∑ 〈<br />

Ex Hy〉<br />

∗<br />

l=1<br />

l<br />

(6.7)<br />

Conventional way of gett<strong>in</strong>g an <strong>in</strong>itial estimate for Z is by solv<strong>in</strong>g the equation 6.1<br />

<strong>in</strong> least square sense (see § 3.3.1.3), with the global cross and auto spectra as obta<strong>in</strong>ed<br />

above.<br />

6.2.4.2 Jackknife estimate as <strong>in</strong>itial guess<br />

The idea of <strong>us<strong>in</strong>g</strong> LS as an <strong>in</strong>itial guess is to keep the solution somewhere <strong>in</strong> the convergence<br />

path and then to iterate to a solution with robust re-weight<strong>in</strong>g. If one can keep<br />

the <strong>in</strong>itial solution nearer to the actual/desired one, the convergence can be fast. Here<br />

the use of a simple nonparametric estimator called Jackknife (Efron [1982], Chave and<br />

Thomson [1989], Eisel and Egbert [2001]) as a better replacement for LS as <strong>in</strong>itial guess is<br />

demonstrated. Chave and Thomson [1989] demonstrated its ability to estimate variances<br />

of the MT transfer functions robustly, alleviat<strong>in</strong>g the need to accurately compute the dof<br />

as needed by conventional variance estimate (such as given <strong>in</strong> 6.16). However, Eisel and<br />

Egbert [2001], while discuss<strong>in</strong>g the results from process<strong>in</strong>g of 2 years of cont<strong>in</strong>uous MT<br />

<strong>data</strong>, commented that they are systematically too large for most of the periods. The<br />

jackknife estimator was used to compute an <strong>in</strong>itial MT transfer function (not variance)<br />

<strong>in</strong> the robust process<strong>in</strong>g rout<strong>in</strong>e. This is shown to be resistant to outliers and thus may<br />

produce an <strong>in</strong>itial guess, nearer to the true one. One advantage of Jackknife method is<br />

its computational simplicity. This is important, as these rout<strong>in</strong>es will be regularly called<br />

<strong>in</strong> robust process<strong>in</strong>g. Consider Z as any one component of the transfer function derived


97<br />

by locally solv<strong>in</strong>g the equation (6.1). From all the time segments we have L number of<br />

estimates of Z. Let Z mean be the mean based on all the <strong>data</strong>. The <strong>data</strong> are then divided<br />

<strong>in</strong>to L groups of size L-1 each by delet<strong>in</strong>g an entry <strong>in</strong> turn from the whole set. Let the<br />

estimate of Z based on i th subset, when the i th datum has been removed be Z −i (Chave<br />

and Thomson [1989]). The jackknife mean is given as,<br />

Z jackknife = LZ mean − L − 1<br />

L<br />

L∑<br />

Z −i (6.8)<br />

The quantity <strong>in</strong> the above equation was orig<strong>in</strong>ally <strong>in</strong>troduced as a low bias replacement<br />

for regular mean (Chave & Thomson, 1989, Effron, 1982). We may construct a difference<br />

vector,<br />

∣<br />

Z diff<br />

−i =<br />

∣ Z −i −<br />

The Jackknife variance is then given by,<br />

Z var = L − 1<br />

L<br />

i=1<br />

∣<br />

L∑ ∣∣∣∣<br />

Z −i . (6.9)<br />

i=1<br />

L∑ (<br />

i=1<br />

Z diff<br />

−i<br />

) 2<br />

. (6.10)<br />

Note that jackknife variance is entirely different from the conventional variance which<br />

are ‘parametric’ estimations. (In the sense they depend on number of dof, value of<br />

coherence etc). Advantages and disadvantages of this variance are thoroughly discussed<br />

by Chave and Thomson [1989], Eisel and Egbert [2001]. The aim here is to <strong>in</strong>vestigate<br />

the use of jackknife estimate as an <strong>in</strong>itial guess for robust process<strong>in</strong>g. Iterations can be<br />

performed by successively delet<strong>in</strong>g <strong>data</strong> rows (stacks) which gives maximum difference<br />

as <strong>in</strong> equation 6.9. After each iteration, the variance of current iteration is compared<br />

with the previous one. Iterations are stopped when there is no more improvement to the<br />

variance, or the <strong>data</strong> get exhausted.<br />

An example of the application of Jackknife estimation of MT transfer function is<br />

plotted <strong>in</strong> the Figure 6.4. From the station VP10, a total of 103 overlapped sections<br />

(length 1024) of time series <strong>data</strong> were available. An equal number of transfer functions (Z)<br />

were generated from these <strong>data</strong> sets. The jackknife difference as calculated by equation<br />

6.9 for the first iteration is shown <strong>in</strong> Figure 6.4(a). The plot shows large differences<br />

especially at the beg<strong>in</strong>n<strong>in</strong>g and end of the stacks. To proceed <strong>data</strong> row (stack) with<br />

highest difference will be deleted. Figure 6.4(b) shows the progressive decrease of variance<br />

as the iteration progresses. In this case, iteration was stopped at 28. Figure 6.4(c) and<br />

(d) compares the output of a least square and jackknife estimations of MT apparent<br />

resistivity and phase values (YX). Figures 6.5(a) to (d) depict the jackknife estimation<br />

of MT transfer functions for site VP13. As can be seen from the plots, the jackknife<br />

gives better estimate of the parameters as compared to the LS. The LS estimations were<br />

distorted by few and large unusual <strong>data</strong> elements. However, the jackknife estimations<br />

are not advised as the f<strong>in</strong>al ones, as the variability of jackknife estimate can be large<br />

for some statistical distributions (Chave and Thomson [1989]). It may be concluded<br />

that jackknife can be good start for robust process<strong>in</strong>g, consider<strong>in</strong>g its computational


Figure 6.4: Comparison of Least Square and Jackknife estimation of MT transfer functions<br />

for station VP10 (a) Jackknife difference for the first iteration. (b) Variance as a<br />

function of iteration number. (c) and (d) comparison of ρ and φ values from LS & JK<br />

process<strong>in</strong>g<br />

98


99<br />

Figure 6.5: Comparison of Least Square and Jackknife estimation of MT transfer functions<br />

for station VP13 (a) Jackknife difference for the first iteration. (b) Variance as a<br />

function of iteration number. (c) and (d) comparison of ρ and φ values from LS & JK<br />

process<strong>in</strong>g<br />

simplicity, non-parametric nature and superior performance over LS.


100<br />

6.2.4.3 Scale estimate<br />

Once the <strong>in</strong>itial guess for transfer function is made, we derive the residuals r l , for all the<br />

time segments for a particular frequency as,<br />

r l2<br />

x<br />

= ∣ E<br />

l<br />

x − Z xx Hx l − Z xy Hy<br />

l ∣ 2 (6.11)<br />

Equation 6.11 describes one way of assess<strong>in</strong>g the quality of a least square solution<br />

to the observations <strong>in</strong> E and H. There are also other possibilities like variance ∆Z,<br />

coherence γ 2 etc to be used as quality parameters (Ernst et al. [2001]). However, the<br />

residuals between predicted and observed fields are popularly used to assess the quality<br />

of LS solution of MT transfer function and are used <strong>in</strong> this thesis as well.<br />

An <strong>in</strong>itial MAD scale estimate for the robust weight<strong>in</strong>g procedures is obta<strong>in</strong>ed as,<br />

d M = 1.483med.(|r l − med.(r l )|). (6.12)<br />

For MT <strong>data</strong> the residuals <strong>in</strong> the above equations are complex <strong>in</strong> nature. One way<br />

of measur<strong>in</strong>g the residual size is by its magnitude and it is preferred, as it is rotationally<br />

<strong>in</strong>variant (Chave and Thomson [1989]). In a simple way, the weight assigned to a complex<br />

value changes both real and imag<strong>in</strong>ary parts <strong>in</strong> the same way such that its phase rema<strong>in</strong>s<br />

the same.<br />

An upper limit to the scale estimate may be given as k M = 1.5d M (see §6.2.3 case 3).<br />

6.2.4.4 Robust transfer function estimation<br />

The Huber weights w l for each stack l is calculated accord<strong>in</strong>g to k M .<br />

{ 1 for r<br />

w l =<br />

l ≤ k M<br />

k Mr<br />

l for r l (6.13)<br />

> k M<br />

Now <strong>us<strong>in</strong>g</strong> the weights, new auto and cross-spectral estimates are given by,<br />

〈 1<br />

Ex Hy〉 ∗ =<br />

L∑<br />

L<br />

∑<br />

w l l=1<br />

l=1<br />

w l 〈 E x H ∗ y〉 l<br />

(6.14)<br />

From the modified spectra sets, new estimates (robust) of transfer function Z are<br />

obta<strong>in</strong>ed. To estimate the new MAD scale estimate, we must take <strong>in</strong>to account the<br />

weight<strong>in</strong>g process.<br />

d 2 H = L L 2 c<br />

L∑<br />

w l (r l ) 2 (6.15)<br />

l=1<br />

where, L c is the number of weighted events with w l = 1 <strong>in</strong> step 6.13. Now by <strong>us<strong>in</strong>g</strong><br />

the new upper limit k H = 1.5 d H , the steps 6.11 and 6.15 are repeated, result<strong>in</strong>g another<br />

estimation of Z xx & Z xy . Though, it is advised <strong>in</strong> the literature to iterate equations 6.11<br />

to 6.15, to our experience the transfer functions hardly changes after second iteration.


101<br />

6.2.4.5 Tukey weights<br />

The extreme outliers <strong>in</strong> the <strong>data</strong> are removed by aga<strong>in</strong> weight<strong>in</strong>g the spectral matrices<br />

with Tukey’s biweight criterion. In order to compute a new scale estimate. Follow<strong>in</strong>g<br />

Ritter et al. [1998],<br />

d 2 T =<br />

1<br />

L<br />

L∑<br />

1<br />

(<br />

1 −<br />

(<br />

1<br />

L<br />

L∑<br />

(w l r l ) 2<br />

l=1<br />

) ) 2<br />

r l<br />

k H<br />

(1 − 5<br />

(<br />

) ) (6.16)<br />

2<br />

r l<br />

k H<br />

With an upper limit k T = 6d T , we obta<strong>in</strong> Tukey weights w l as (reproduc<strong>in</strong>g from<br />

Table 6.1),<br />

{ ( )<br />

w l r<br />

1 −<br />

=<br />

k T<br />

for r l ≤ k T<br />

0 for r l > k T<br />

(6.17)<br />

Spectral matrices are weighted aga<strong>in</strong> with this new weight. This forms the f<strong>in</strong>al step<br />

<strong>in</strong> robust process<strong>in</strong>g. The transfer functions are computed from the weighted spectral<br />

matrices.<br />

6.2.4.6 Comput<strong>in</strong>g the variance<br />

The robust process<strong>in</strong>g procedure modifies the degree of freedom of <strong>in</strong>dividual estimates.<br />

While comput<strong>in</strong>g the variance, this also must be taken <strong>in</strong>to consideration. In the process<strong>in</strong>g<br />

code, it was realized for each weight<strong>in</strong>g step as,<br />

dof new = dof (∑<br />

old w<br />

l)<br />

. (6.18)<br />

L<br />

Now the variance for one element of the transfer function is given by,<br />

(∆Z xy ) 2 = k F (k, 2dof − 4, δ = 0.05)[1 − Coh2 (E x E p x)]E x E ∗ x<br />

2dof − 4 [1 − Coh 2 (H x H y )]H x H ∗ y<br />

(6.19)<br />

Müller [2000]. Where F is the Fischer distribution, with k = 4 (See Bendat and<br />

Piersol [1971]). An <strong>in</strong>direct estimate of the degree of noise level can be obta<strong>in</strong>ed from<br />

the misfit of the f<strong>in</strong>al model (impedances) with the weighted <strong>data</strong> sets (cross and auto<br />

powers) of the f<strong>in</strong>al iteration. However this assumes that, once the outliers are removed,<br />

the residuals (model predictions observations) gets a Gaussian distribution<br />

6.2.4.7 Quantile Quantile plots<br />

In Figure 6.6, comparison of Least Square and the robust process<strong>in</strong>g method (described<br />

above) is demonstrated for station TT08. The E x -E xp residuals from LS estimates of<br />

transfer functions for all the stacks of frequency 0.0791 Hz are plotted <strong>in</strong> Figure 6.6(a) and<br />

b (zoomed version). Inverted triangle represents the un-weighted (LS) residuals and stars<br />

represent the weighted (Robust) residuals. In general, the residuals follow a Gaussian<br />

distribution, but superimposed by few outliers. The Quantile – Quantile (QQ) plot <strong>in</strong>


102<br />

Figure 6.6: Comparison of Least Square and Robust process<strong>in</strong>g of magnetotelluric <strong>data</strong><br />

station TT08 for frequency 0.0791Hz . Triangles represent LS process<strong>in</strong>g, and stars<br />

represent robust (RB) process<strong>in</strong>g (a) and (b) time series of Ex residuals. C) Quantile<br />

Quantile plot of Ex residuals d) MT apparent resistivity and phase values from LS and<br />

robust process<strong>in</strong>g. See text for discussion.<br />

Figure 6.6(c) better expla<strong>in</strong>s this statement. QQ plot describes the departure of observed<br />

<strong>data</strong> set from a particular distribution, here Gaussian distribution (§ 3.2.1.1). If the<br />

residuals are drawn from a Gaussian distribution, the QQ plot will be an approximately<br />

straight l<strong>in</strong>e (Chave et al. [1987]). As can be seen from the Figure 6.6(c), up to 1.5σ<br />

(standard normal quantile) the residuals (LS – <strong>in</strong>verted triangles) follow a straight l<strong>in</strong>e.<br />

However, the departure from Gaussian distribution is evident above 1.5σ. The weighted<br />

residuals (Robust – stars) closely follow the straight l<strong>in</strong>e, <strong>in</strong>dicat<strong>in</strong>g the effectiveness of<br />

robust process<strong>in</strong>g methods to elim<strong>in</strong>ate non-Gaussian errors. The effect of these outliers<br />

on computed MT apparent resistivity (ρ xy ) and phases (φ xy ) are shown <strong>in</strong> Figure 6.6(d).<br />

The highly irregular nature of the LS estimate (triangles) are the results of a few outliers<br />

<strong>in</strong> the <strong>data</strong> <strong>in</strong>fluenc<strong>in</strong>g the transfer function estimate. The robust process<strong>in</strong>g resulted <strong>in</strong><br />

better estimation of both ρ xy and φ xy for all the frequencies.<br />

6.2.5 Application<br />

6.2.5.1 Flowchart<br />

A flow chart for the robust process<strong>in</strong>g scheme developed <strong>in</strong> l<strong>in</strong>e with the discussions <strong>in</strong><br />

the above sections is given <strong>in</strong> Figure 6.7(a). Robust process<strong>in</strong>g scheme receives a spectral


103<br />

Figure 6.7: Flow chart for robust process<strong>in</strong>g scheme. The gray area represents the proposed<br />

<strong>in</strong>itialization of the transfer functions <strong>us<strong>in</strong>g</strong> Jackknife. This rout<strong>in</strong>e, concentrates<br />

on the section averag<strong>in</strong>g of MT <strong>data</strong>. See text for discussion.


104<br />

Figure 6.8: Data flow through the flow chart represent<strong>in</strong>g robust process<strong>in</strong>g of MT <strong>data</strong>.<br />

The gray area represents the proposed Jackknife <strong>in</strong>itialization. (a) Process A uses Jackknife<br />

(JK) as <strong>in</strong>itial guess and (b) process B uses Least Square (LS) as <strong>in</strong>itial guess. See<br />

text for discussion.<br />

matrix (4 dimensional) as <strong>in</strong>put. The process starts with an option of least square (LS)<br />

or Jackknife (JK) solution as <strong>in</strong>itial guess. Then the <strong>data</strong> successively go through the<br />

weight<strong>in</strong>g procedures. In majority of cases, more than two iterations may not be needed<br />

and one can apply the Tukey’s weights for f<strong>in</strong>al estimation of spectra sets. F<strong>in</strong>ally the 4D<br />

(L stacks, N frequencies n x n channel) spectral matrix gets reduced to 3D (N frequencies<br />

n x n channel), which is the f<strong>in</strong>al output of the robust algorithm. This rout<strong>in</strong>e, written<br />

<strong>in</strong> MATLAB [2001] takes 10-20 seconds on a 500 MHz PC, with a typical 5 channel MT<br />

time series with a total of ∼10 6 values. Two processes are def<strong>in</strong>ed to demonstrate the<br />

superiority of robust process<strong>in</strong>g with Jackknife as <strong>in</strong>itial guess. The <strong>data</strong> flow for the two<br />

processes are plotted on the flow chart <strong>in</strong> Figure 6.8(a) & (b). Process A uses robust<br />

process<strong>in</strong>g with Jackknife (JK) as <strong>in</strong>itial guess, where as Process B uses least square<br />

(LS) for the same. The two process<strong>in</strong>g schemes are compared <strong>in</strong> the follow<strong>in</strong>g section,<br />

accord<strong>in</strong>g to their performance on application to 8 MT stations from SGT.<br />

6.2.5.2 Comparison of robust process<strong>in</strong>g schemes<br />

The eight stations chosen for the demonstration are evenly distributed <strong>in</strong> the measurement<br />

corridor (Figure 4.1) and the results are presented <strong>in</strong> Figures 6.9 and 6.10. In all<br />

the plots, the solid symbols represent Process A and open symbols represent Process B.<br />

The station JN12 (Figure 6.9(a)) located <strong>in</strong> the northern half of the corridor was affected<br />

by a large number of spikes <strong>in</strong> the longer periods and by power l<strong>in</strong>e harmonics <strong>in</strong> the<br />

range 10 to 100 Hz. This is evident as the ρ xy <strong>in</strong> this frequency range exhibits grossly<br />

different values from Process B. Robust process<strong>in</strong>g with Jackknife (Process A) improved


105<br />

the estimates of ρ xy <strong>in</strong> the frequency range 10 to 100 Hz. In the longer period, Process A<br />

resulted <strong>in</strong> smoother ρ and φ values. In station VP16 (Figure 6.9(b)), no major change<br />

is observed between results of processes A and B. Perhaps as the station is relatively less<br />

affected with noise, the <strong>in</strong>itial LS solution itself is comparable to that of Jackknife (A).<br />

Even then Process A resulted <strong>in</strong> better estimation of φ values <strong>in</strong> the frequency range 10<br />

Hz to 1Hz. The results from OK16 (Figure 6.9(c)) show that process A (JK) <strong>in</strong>itialization<br />

created a down bias <strong>in</strong> ρ xy values as compared to process ‘B’. However, the process<br />

A resulted <strong>in</strong> smoother φ values as compared to process B. In a later discussion it will<br />

be shown that how this bias may be removed from Process A. The Process A resulted<br />

<strong>in</strong> smoother estimates of φ values for station VP12 (Figure 6.9(d)) as compared to the<br />

output of process B. However, the sharp drop of ρ values computed through Process A<br />

near 6 seconds seems to be unreal.<br />

The dramatic improvement of φ values (especially for periods > 1 Sec) at station<br />

VP14 (Figure 6.10(a)) by Process A as compared to Process B gives another evidence<br />

that robust process<strong>in</strong>g results <strong>in</strong> better estimation, if started with a good <strong>in</strong>itialisation.<br />

It is also seen from the Figure 6.10(c) that the ρ values from process A are less scattered<br />

<strong>in</strong> the same range of period. Station TT08 is located near the <strong>in</strong>dustrial zone, along<br />

the Cauvery river (Figure 4.1, see § 4 for discussion on this). It was written earlier<br />

that the long period time series from this station is affected with high magnitude spike<br />

activity. The Jackknife estimates (A) resulted <strong>in</strong> better estimates of φ yx values especially<br />

by def<strong>in</strong><strong>in</strong>g a smooth curve between 1 and 10 sec (Figure 6.10(b)). Though Process A,<br />

could not produce better results for φ xy , it failed gracefully by predict<strong>in</strong>g a trend for<br />

the φ xy , <strong>in</strong> the period range 1 to 10 sec, which was absent from the Process B. The E x<br />

channel of station TT04 was severely affected by near source noise (§ 2.4) <strong>in</strong> the frequency<br />

range 0.1 Hz to 10Hz, as manifested <strong>in</strong> a 45 0 raise <strong>in</strong> ρ xy values and small values for φ xy<br />

values. The conventional robust process<strong>in</strong>g, which was biased by the LS <strong>in</strong>itialisation,<br />

failed <strong>in</strong> the presence of majority of noisy <strong>data</strong>. Jackknife produced a better <strong>in</strong>itial guess<br />

and when followed by the ma<strong>in</strong> robust scheme, dramatically improved the ρ xy and φ xy<br />

values as shown <strong>in</strong> Figure 6.10(c). However, <strong>in</strong> the longer period the ρ yx values seem to<br />

be biased down compared to conventional process<strong>in</strong>g (Process B). At station OK18, the<br />

ρ values from both process<strong>in</strong>g do not differ much. Still, the φ xy values from Process B,<br />

seems to be scattered for periods > 10 sec. Also the steep decrease <strong>in</strong> φ xy values with<br />

<strong>in</strong>crease <strong>in</strong> frequency from1 Hz to 10 Hz seem to be unreal and <strong>in</strong>consistent with ρ xy<br />

values. In both the frequency ranges Process A produced a better estimates as shown <strong>in</strong><br />

Figure 6.10(d).<br />

6.3 Robust Band averag<strong>in</strong>g<br />

6.3.1 Introduction<br />

Role of section and band averag<strong>in</strong>g of auto and cross spectra <strong>in</strong> the context of magnetotelluric<br />

signals was discussed <strong>in</strong> § 3.2.3.4. It was shown by Jenk<strong>in</strong>s and Watts [1968]<br />

that subdivid<strong>in</strong>g a time series <strong>in</strong>to sections of length M and form<strong>in</strong>g smoothed spectra<br />

is equivalent to smooth<strong>in</strong>g the spectrum of undivided time series by a s<strong>in</strong>c (s<strong>in</strong>(x)/x)


106<br />

Figure 6.9: Comparison of robust process<strong>in</strong>g results <strong>us<strong>in</strong>g</strong> Least Square (LS) and Jackknife<br />

(JK) <strong>in</strong>itialization for stations JN12, VP16, OK16 and VP12. Process A refers to<br />

robust process<strong>in</strong>g with JK <strong>in</strong>itialization, where as Process B refers robust process<strong>in</strong>g with<br />

LS <strong>in</strong>itialization See legend for symbol identification


107<br />

Figure 6.10: Comparison of robust process<strong>in</strong>g results <strong>us<strong>in</strong>g</strong> Least Square (LS) and Jackknife<br />

(JK) <strong>in</strong>itialization for stations VP14,TT08,TT04 and OK18. Process A refers to<br />

robust process<strong>in</strong>g with JK <strong>in</strong>itialization, where as Process B refers robust process<strong>in</strong>g with<br />

LS <strong>in</strong>itialization. See legend for symbol identification


108<br />

function. In magnetotellurics, it is a common practice to use the comb<strong>in</strong>ation of section<br />

and band smooth<strong>in</strong>g, due to various reasons.<br />

1. With magnetotelluric <strong>data</strong>, one seeks to estimate a time stationary quantity (transfer<br />

functions), which varies smoothly over frequency – Band averag<strong>in</strong>g over a spectral<br />

w<strong>in</strong>dow facilitates this.<br />

2. The measured time series may not be a contiguous sequence as there may be gaps<br />

<strong>in</strong> time series record<strong>in</strong>g due to <strong>in</strong>strumental reasons – This necessitates the use of<br />

section averag<strong>in</strong>g <strong>in</strong> addition to band averag<strong>in</strong>g<br />

3. A short duration noise with high power, may corrupt larger number of FC’s if band<br />

averag<strong>in</strong>g only used, than the case where spectra is estimated from sub-segmented<br />

time series sections (section averag<strong>in</strong>g) (Egbert and Booker [1986]) - This make it<br />

necessary to reject parts of time series affected by noise.<br />

The robust statistical procedures, as the one presented <strong>in</strong> § 6.2, concentrates only<br />

the section averag<strong>in</strong>g. Sims et al. [1971] commented that the average over adjacent<br />

frequencies would facilitate the estimation of spectra and might give sufficient dof to<br />

reduce noise powers <strong>in</strong> cross spectra. The mention of band averag<strong>in</strong>g can also be found<br />

elsewhere <strong>in</strong> context of magnetotelluric time series process<strong>in</strong>g (Swift [1986], Gamble et al.<br />

[1979]). While apply<strong>in</strong>g the robust procedures on MT transfer function estimates, Chave<br />

et al. [1987], Chave and Thomson [1989] op<strong>in</strong>ed that the same approach might not be<br />

amenable to band averag<strong>in</strong>g. Egbert [1997] discussed robust estimation of spectral density<br />

matrices (SDM) for multi-station MT process<strong>in</strong>g. However, he limited himself to the<br />

problem of estimat<strong>in</strong>g SDMs from a set of sections, which are <strong>in</strong>dividually averaged<br />

over a frequency band. Recogniz<strong>in</strong>g the need for better estimation of cross and auto<br />

spectra before the actual estimation of transfer functions, Ritter et al. [1998] suggested<br />

procedures to predict s<strong>in</strong>gle event spectra by compar<strong>in</strong>g with a global (estimate from all<br />

the stacks/sections) average. While compar<strong>in</strong>g two process<strong>in</strong>g techniques for synthetic<br />

MT time series process<strong>in</strong>g, Ernst et al. [2001], commented that the use of lesser number<br />

of frequency po<strong>in</strong>ts <strong>in</strong> the vic<strong>in</strong>ity of target frequency to average the spectra might help<br />

to prevent over smooth<strong>in</strong>g of long period estimates. Recently Smirnov [2003] computed<br />

MT transfer functions for all the Fourier coefficient pairs and used a repeated median<br />

estimator to arrive at a better estimate of MT transfer function. However, such an<br />

approach may fail at the <strong>in</strong>stance of strongly polarized fields (either due to the Geology or<br />

to coherent noise), as it may not give enough degree of freedom for a practical estimation.<br />

The effect of few outliers on MT <strong>data</strong> process<strong>in</strong>g is well recognized and it is demonstrated<br />

<strong>in</strong> the previous section. However, effect of outliers <strong>in</strong> band averag<strong>in</strong>g <strong>in</strong> the vic<strong>in</strong>ity of<br />

a target frequency has not been estimated. It is usual <strong>in</strong> MT, to apply a small w<strong>in</strong>dow<br />

over the target frequency (f ) (Figure 3.3) and average the cross and auto spectra with<br />

respect to this w<strong>in</strong>dow. For example the cross spectra between E x and H y at the target<br />

frequency f from a s<strong>in</strong>gle event spectra is obta<strong>in</strong>ed by,<br />

E x H ∗ y(f) = 1 M<br />

M/2<br />

∑<br />

k=−M/2<br />

E x (f + k)H ∗ y(f + k).w<strong>in</strong>(k) (6.20)


109<br />

To give lesser weights to the harmonics farther from the central frequency, it is usual<br />

to average over a w<strong>in</strong>dow function. The number of harmonics to be averaged depends<br />

upon the position of target frequency with<strong>in</strong> the spectrum and the length of the time<br />

series segment. If the time series were segmented <strong>in</strong>to smaller sections, the number of<br />

harmonics to be averaged over a particular frequency reduces. However, this will be<br />

compensated by the <strong>in</strong>creased amount of <strong>data</strong> due to larger number of sections. The<br />

robust process<strong>in</strong>g methods thus advocated the need for shorter time w<strong>in</strong>dows (Egbert<br />

and Booker [1986]) . Even if the time series does not arise from a Gaussian process, it<br />

is assumed that the Fourier coefficients (FC) that result from the summarization of time<br />

series, follow a Gaussian distribution, (from Central Limit Theorem) (Chave et al. [1987],<br />

Menke [1984]). So with<strong>in</strong> a band it was felt reasonable to assume a complex Gaussian<br />

distribution for the FCs and LS averages were used to estimate auto and cross spectra.<br />

However, it will be shown <strong>in</strong> the next section that this assumption often fails due to the<br />

presence of some unusual <strong>data</strong>. A weighted robust averag<strong>in</strong>g is essential to remove the<br />

outliers with<strong>in</strong> a frequency band well.<br />

6.3.2 Effect of frequency band width<br />

In Figure 6.11 the concept of frequency band averag<strong>in</strong>g is presented for vary<strong>in</strong>g w<strong>in</strong>dow<br />

lengths, effectively <strong>us<strong>in</strong>g</strong> more FCs <strong>in</strong> each w<strong>in</strong>dow <strong>in</strong> Figure 6.11 (a) to (d). The term<br />

‘frequency band’ refers to the narrow band around a target frequency, whereas the term<br />

band 1, band 2, etc represents the measurement band. ’Parzen’ w<strong>in</strong>dow (§3.2.3.5; also see<br />

Jenk<strong>in</strong>s and Watts [1968]) is applied <strong>in</strong> frequency doma<strong>in</strong>, whereas ’Hann<strong>in</strong>g’ w<strong>in</strong>dow is<br />

applied <strong>in</strong> time doma<strong>in</strong> (Bendat and Piersol [1971]). The upper four plots (a to d) show<br />

the Parzen w<strong>in</strong>dow with different sizes. Parzen w<strong>in</strong>dow is preferable over a boxcar, as 1)<br />

it gives less and less importance to farther harmonics from the target frequency and 2)<br />

Boxcar (§ 3.2.3.3) <strong>in</strong> frequency doma<strong>in</strong> has a s<strong>in</strong>c (s<strong>in</strong>(x)/x) ‘impulse response’, which<br />

is not desirable <strong>in</strong> spectrum estimation. The vertical broken l<strong>in</strong>es are drawn through<br />

each target frequency at which MT transfer functions are to be computed. The bottom<br />

plot show a sample E x E x spectra for <strong>data</strong> collected <strong>in</strong> the frequency range 8 Hz to 0.25<br />

Hz, with a sampl<strong>in</strong>g <strong>in</strong>terval 32 Hz. The spectrum at each target frequency is usually<br />

estimated by averag<strong>in</strong>g the adjacent harmonics with respect to the w<strong>in</strong>dow. Though<br />

the w<strong>in</strong>dow width seems to <strong>in</strong>crease as the frequency decreases, the effective bandwidth<br />

rema<strong>in</strong>s the same, consider<strong>in</strong>g the distribution of FC on a logscale (Wight and Bostick<br />

[1980]).<br />

In order to study the effect of frequency band averag<strong>in</strong>g on MT transfer functions, the<br />

ρ xy and ρ yx for a particular frequency were computed <strong>us<strong>in</strong>g</strong> different frequency w<strong>in</strong>dow<br />

lengths. Figure 6.12 show ρ xy and ρ yx values at 6Hz plotted as a function of number of harmonics<br />

used for the station VP13. The time series were subdivided <strong>in</strong>to sections of 2048<br />

length and robust procedure described <strong>in</strong> § 6.2 is used to estimate the transfer functions.<br />

Variance of each was estimated accord<strong>in</strong>g to equation 6.17, tak<strong>in</strong>g <strong>in</strong>to consideration the<br />

modification to dof by the robust procedures. Two results from this computation are: 1)<br />

As the estimate is made over longer w<strong>in</strong>dow sizes, the expected variance decreases. This<br />

can be seen <strong>in</strong> the form of reduced error bars for the estimates with larger number of<br />

harmonics. 2) However, the reduction <strong>in</strong> variance comes at the cost of bias error. As the


110<br />

Figure 6.11: Spectrum estimation <strong>in</strong> MT <strong>us<strong>in</strong>g</strong> band averag<strong>in</strong>g. (a) to (d) w<strong>in</strong>dows <strong>in</strong><br />

frequency with different radii. Though the w<strong>in</strong>dow radius seems to <strong>in</strong>crease as period<br />

<strong>in</strong>creases, effective bandwidth rema<strong>in</strong>s the same, as spectrum <strong>in</strong> long period conta<strong>in</strong>s fewer<br />

Fourier harmonics. (e) Sample E x E ∗ x spectrum <strong>in</strong> the band 4, with target frequencies<br />

projected as dotted l<strong>in</strong>es. It is common to have 10 12 target frequencies per band.


111<br />

Figure 6.12: Apparent resistivities as a function of frequency w<strong>in</strong>dow length. Triangles<br />

represent ρ xy and circles ρ yx . Error bars represents 95% confidence <strong>in</strong>terval. See text for<br />

discussion.<br />

w<strong>in</strong>dow size <strong>in</strong>creases the computed ρ xy and ρ yx values get biased down. The ρ xy values,<br />

which are around 7000 Ohm m at a frequency w<strong>in</strong>dow with less than 100 harmonics, get<br />

biased to ∼5000 Ohm m with number of harmonics > 250. This dependence of resistivity<br />

values on frequency bandwidth has noth<strong>in</strong>g to do with the geology and are clearly due to<br />

noise. To be more specific, the noise power <strong>in</strong> the magnetic field auto spectra may not get<br />

averaged out as the number of <strong>data</strong> po<strong>in</strong>t <strong>in</strong>crease as they are squared and positive and<br />

any outliers may worsen the situation. It is worthwhile to exam<strong>in</strong>e the distribution of<br />

these residuals to check the assumption of Gaussian distribution for the spectra with<strong>in</strong> a<br />

band. Figure 6.13(a). shows the magnitude of a cross spectrum obta<strong>in</strong>ed by multiply<strong>in</strong>g<br />

H x with H y ∗ respect<strong>in</strong>g the Parzen w<strong>in</strong>dow <strong>in</strong> frequency doma<strong>in</strong>, at frequency 6 Hz for<br />

station VP13. Few frequency po<strong>in</strong>ts with grossly different spectral amplitude than the<br />

majority are seen. A Quantile-Quantile plot of spectra of real part of the cross spectrum<br />

(Figure 6.13(b)) shows (<strong>in</strong>verted triangles - unweighted) large deviation from l<strong>in</strong>ear function<br />

at the higher qu<strong>in</strong>tiles, <strong>in</strong>dicat<strong>in</strong>g the departure from a Gaussian distribution. This<br />

shows that the <strong>in</strong>dividual spectral coefficients with<strong>in</strong> a frequency band might not follow<br />

a Gaussian distribution and an LS average is not desirable <strong>in</strong> such cases.<br />

6.3.3 Robust estimation of spectra<br />

Here, a robust weight<strong>in</strong>g procedure is proposed to m<strong>in</strong>imize the effect of outliers <strong>in</strong> the<br />

estimation of cross and auto-spectra from a frequency band. Once spectra are estimated<br />

<strong>in</strong> a robust sense, any type of section averag<strong>in</strong>g procedures may be adopted for the estimation<br />

of MT transfer functions. Though theoretically possible, computation of transfer<br />

functions <strong>us<strong>in</strong>g</strong> a pair of spectra sets from two adjacent FC’s (Smirnov [2003]) is unstable


112<br />

Figure 6.13: Concepts of robust band averag<strong>in</strong>g. (a) Magnitude of cross spectrum between<br />

H x and H y for station VP13, around a target frequency 6 Hz. The spectra are<br />

multiplied by a Parzen w<strong>in</strong>dow. LS least square, RB Robust. (b) Quantile Quantile<br />

plot for real part of the same cross spectrum. Inverted triangles unweighted, Circles<br />

robust weighted. See text for discussion.


113<br />

because of the chances of hav<strong>in</strong>g same polarization. So m<strong>in</strong>imization of residual noise<br />

from an LS transfer function estimate is undesirable with<strong>in</strong> a frequency band. Without<br />

assum<strong>in</strong>g any signal & noise components, a Huber weight<strong>in</strong>g procedure was used, assum<strong>in</strong>g<br />

that the spectra are random variables. For complex <strong>data</strong>, we used the magnitude to<br />

def<strong>in</strong>e the weights (a discussion <strong>in</strong> this regard is given by Chave and Thomson [1989]. The<br />

algorithm devised for robust spectrum estimation is as follows. At each target frequency<br />

we have M number of FC’s from two channels, <strong>in</strong> the vic<strong>in</strong>ity of the target frequency, f.<br />

We would like to compute the cross spectra between two channels, say Ex and H y from<br />

M realizations of E x H y *, equally distributed around the target frequency, f.<br />

S f ExHy = 〈 〉<br />

E x Hy<br />

∗k<br />

M<br />

∗ w<strong>in</strong>(k), where k = −<br />

2 ...f...M (6.21)<br />

2<br />

Where function w<strong>in</strong> is a Parzen w<strong>in</strong>dow (§ 3.2.3.5) and equation 3.22) with radius<br />

M/2. Residuals r k were calculated by subtract<strong>in</strong>g median (|S|) from each value <strong>in</strong> |S|. A<br />

MAD scale estimate σ is obta<strong>in</strong>ed as <strong>in</strong> equation 6.12 as,<br />

σ = med. (∣ ∣ r<br />

k ∣ ∣ − med.<br />

∣ ∣r k ∣ ∣ ) (6.22)<br />

Then the Huber weights w k are used to down weight the <strong>data</strong> that exceeds r i<br />

{ 1 if rk ≤ σ<br />

w k = σ<br />

r k<br />

if r k > σ<br />

σ .<br />

(6.23)<br />

If the outliers have been elim<strong>in</strong>ated the spectra are the sum of almost normally distributed<br />

cross products. The robust estimate of spectra (here the cross spectra between<br />

E x and H y ) are obta<strong>in</strong>ed as,<br />

E x H ∗ y(f) = 1 M<br />

M/2<br />

∑<br />

k=−M/2<br />

E x (f + k)H ∗ y(f + k).w k .W <strong>in</strong>(k) (6.24)<br />

It may be followed by iteration with Tukey’s weight<strong>in</strong>g (§ 6.2.4.4) to reduce the<br />

<strong>in</strong>fluence of large outliers. However, the experience shows that this may not be required<br />

<strong>in</strong> majority of the cases. Figure 6.13 shows results of robust band averag<strong>in</strong>g. Figure<br />

6.13(b) shows Q-Q plot of robust weighted H x H y cross spectrum. Represented with<br />

circles, the robust weight<strong>in</strong>g clearly reduced the outliers at higher quantiles compared<br />

to un-weighted (or LS) spectra (triangles). The LS estimate gave a value of 2.1563e-010<br />

-3.9535e-011i nT 2 /Hz for the cross spectrum, while the robust estimate gave 1.4426e-010<br />

-2.4245e-011i nT 2 /Hz, both at 6 Hz. This means the LS estimate gave a value ∼60%<br />

higher than that of robust weighted estimate of spectrum at 6Hz. This reduction <strong>in</strong><br />

spectral amplitude (solid horizontal l<strong>in</strong>e, Figure 6.13(a)) by robust weight<strong>in</strong>g could be<br />

attributed to the elim<strong>in</strong>ation of few large outliers, which biased the LS estimate (broken<br />

l<strong>in</strong>e, Figure 6.13(a)) up. Figure 6.14 shows a series of band averag<strong>in</strong>g <strong>in</strong>stances from the<br />

MT <strong>data</strong> at station VP13. All the eight plots show cross spectrum estimation of between<br />

H x and H y , for different target frequencies and different Parzen (Figure 3.3) radius sizes.<br />

Note that X axis do not represent the frequency itself, but the wave number from FFT.<br />

Y axis shows the magnitude of the cross spectrum <strong>in</strong> nT 2 /(Hz). The two horizontal bars


114<br />

show the averages from LS (broken) and robust (solid) weight<strong>in</strong>g scheme. Column (a)<br />

band averag<strong>in</strong>g for four target frequencies and column (b) shows the band averag<strong>in</strong>g at<br />

s<strong>in</strong>gle target frequency (6Hz) with different radius sizes. From these results it can be<br />

seen that there exists a consistent bias error <strong>in</strong> spectrum estimation, which <strong>in</strong> turn may<br />

corrupt the transfer function computed out of it as well. As the magnitude of the spectra<br />

is a positive function, outlier contam<strong>in</strong>ation can only bias it upwards and that is exactly<br />

seen <strong>in</strong> these plots, where robust weight<strong>in</strong>g procedures always resulted <strong>in</strong> smaller values<br />

as compared to LS. The robust band average always resulted <strong>in</strong> smaller magnitudes for<br />

spectra. Therefore the need for robust band averag<strong>in</strong>g is seen.<br />

6.3.4 Flow chart<br />

Flow chart for the proposed robust band averag<strong>in</strong>g to estimate cross and auto spectral<br />

matrices from time series is presented <strong>in</strong> Figure 6.15(a). The entire codes were written<br />

<strong>in</strong> MATLAB [2001]. The <strong>in</strong>put to the rout<strong>in</strong>e can be either standard 5 Channel s<strong>in</strong>gle<br />

station MT time series or 7 channel with remote reference channels. Pre-process<strong>in</strong>g of<br />

<strong>data</strong> <strong>in</strong>volve removal of trend and bias (§ 3.2.3.1) , sub-segmentation, Fourier transform<br />

(§ 3.2.1.5) and calibration (§ 1.3.4. There are two options for band averag<strong>in</strong>g, viz. least<br />

square (LS) and robust (RB). While LS perform conventional frequency band averag<strong>in</strong>g<br />

respect<strong>in</strong>g a w<strong>in</strong>dow around target frequency (§ 3.2.3.5), RB will perform a weighted<br />

averag<strong>in</strong>g approach as presented <strong>in</strong> the above section (equation 6.24). The averag<strong>in</strong>g<br />

rout<strong>in</strong>es are called <strong>in</strong> conf<strong>in</strong>ed loops (L, N and n x n). With a 5 channel MT time series<br />

of 80 stacks of 1024 <strong>data</strong> po<strong>in</strong>ts each, the averag<strong>in</strong>g rout<strong>in</strong>e will be called 24,000 times<br />

to estimate spectra at 12 target frequencies. This necessitates the optimization of the<br />

processes with<strong>in</strong> the averag<strong>in</strong>g rout<strong>in</strong>e. For this reason, the robust weight<strong>in</strong>g procedure<br />

was designed as simple as possible and the f<strong>in</strong>al iteration by Tukey’s weights was avoided.<br />

As a result, the process<strong>in</strong>g of a time series of comparable dimension as earlier stated, takes<br />

∼1 m<strong>in</strong>ute on a 500 MHz Pentium PC. Further reduction <strong>in</strong> process<strong>in</strong>g speed is possible<br />

by rewrit<strong>in</strong>g the codes <strong>in</strong> C or C++. The rout<strong>in</strong>e will deliver a 4 dimensional spectral<br />

matrix as output, which will form <strong>in</strong>put to any section average process<strong>in</strong>g, <strong>in</strong>clud<strong>in</strong>g<br />

robust process<strong>in</strong>g.<br />

6.3.5 Validation<br />

Four more process<strong>in</strong>g schemes are def<strong>in</strong>ed based on this flow chart to validate the results<br />

of robust band averag<strong>in</strong>g. These are 1) s<strong>in</strong>gle station LS band averag<strong>in</strong>g (Process C),<br />

2) s<strong>in</strong>gle station robust band averag<strong>in</strong>g, (Process D) 3) remote reference and LS band<br />

averag<strong>in</strong>g (Process E) and 4) remote reference and robust band averag<strong>in</strong>g (Process F).<br />

They are represented <strong>in</strong> the Figure 6.15(b) to (e) respectively. These processes deliver<br />

the 4D spectral matrix, from the MT time series <strong>data</strong>. The process A, def<strong>in</strong>ed <strong>in</strong> § 6.2.5,<br />

which is a robust section averag<strong>in</strong>g method employ<strong>in</strong>g a Jackknife <strong>in</strong>itial guess will be<br />

used to estimate MT transfer functions for all the four processes def<strong>in</strong>ed above. In order<br />

to demonstrate the efficacy of the robust band averag<strong>in</strong>g methods, station VP10 was<br />

chosen, where, <strong>in</strong> addition to the s<strong>in</strong>gle station <strong>data</strong> for all the bands, one session of<br />

remote reference <strong>data</strong> were also available from station VP13. This enabled us to exam-


115<br />

<strong>in</strong>e the performance of the proposed robust process<strong>in</strong>g procedure with remote reference<br />

process<strong>in</strong>g as well. The results from four process<strong>in</strong>g methods on the <strong>data</strong> set are presented<br />

<strong>in</strong> Figure 6.16. Figure 6.16(a) shows the ρ xy component, where as ρ yx component<br />

is plotted <strong>in</strong> Figure 6.16(b). The φ values are not shown, as they are not much varied<br />

by the processes described above. However, the φ values are <strong>in</strong>cluded at a later stage<br />

when the applications of the procedure is discussed. The quantity of remote reference<br />

<strong>data</strong> for period range 4 sec to 128 sec (band 4) was <strong>in</strong>sufficient for a through estimation.<br />

However, this was <strong>in</strong>cluded to get a complete evaluation. It can be seen from the figures<br />

that the best and unbiased estimate for ρ values are derived from process F (stars). The<br />

ρ values are clearly higher and less smoother than the other estimates, except for a few<br />

<strong>data</strong> po<strong>in</strong>t between 10 and 100 Hz. The conventional robust remote reference process<strong>in</strong>g<br />

(E -solid l<strong>in</strong>e) , are similar to the results from process F, except for few <strong>data</strong> po<strong>in</strong>ts<br />

<strong>in</strong> ‘dead’ band, where they are biased down compared to process F. The usefulness of<br />

robust band averag<strong>in</strong>g results are more evident <strong>in</strong> s<strong>in</strong>gle station estimates (Process C &<br />

D). Clearly both these process output ρ values that are down biased compared to remote<br />

reference <strong>data</strong>. However, the robust band averag<strong>in</strong>g (Process D) improves the situation,<br />

with the ρ values that are nearer the their remote reference counter parts. This clearly<br />

shows the effect of robustly averag<strong>in</strong>g the spectra sets with<strong>in</strong> a band. On s<strong>in</strong>gle station<br />

estimates, the robust band averag<strong>in</strong>g reduces the down bias as compared to LS band<br />

average. Even remote reference process<strong>in</strong>g can benefit from robust band averag<strong>in</strong>g, as<br />

evidenced from the improved RR estimate of ρ xy and ρ yx values by robust band average<br />

<strong>in</strong> the ‘dead’ band range (1 to 10 seconds). This <strong>analysis</strong> could not be extended to all<br />

other <strong>data</strong> sets, as reference <strong>data</strong> sets were not recorded at majority of stations due to<br />

<strong>in</strong>strument problems.<br />

In the second approach to validate the results of robust band averag<strong>in</strong>g, the estimates<br />

of ρ values from robust band averag<strong>in</strong>g were compared with E-referenced and H-referenced<br />

estimators on spectra generated from conventional robust process<strong>in</strong>g. Standard LS estimators<br />

have the property that E-reference (transfer function from admittance <strong>analysis</strong>)<br />

are biased upward by un-correlated noise on the electric field components and the H-<br />

reference estimates are biased downwards by the noise <strong>in</strong> un-correlated noise <strong>in</strong> magnetic<br />

field. (See § 3 for more discussion). Obviously downward and upward biased estimates<br />

give an envelope with<strong>in</strong> which the true transfer function should lie (Jones et al. [1989]).<br />

The ’up’ and ’down’ (§ 3.3.1.3) biased estimates were computed from process C and<br />

compared it with the result of Process D. Both these processes are done <strong>in</strong> s<strong>in</strong>gle station<br />

mode and the results from four stations are plotted <strong>in</strong> the figures 6.17 to 6.20. Figure<br />

6.17 shows the <strong>data</strong> from station JN12. The up biased estimates from Process C shows<br />

larger deviation near the dead band. This is common, as the s/n ratio at this frequency<br />

range is low compared to other ranges. As clearly seen from Figure 6.17, the ρ xy and<br />

ρ yx components are less biased from Process D. This is more pronounce near the ’dead’<br />

band (between 1 and 10 seconds). Near the period of 100 seconds, the ρ xy (top panel)<br />

values from Process C seems to be heavily biased, which to an extend restored by Process<br />

D. Station VP16 is relatively noise free and ρ xy and ρ yx values are well def<strong>in</strong>ed over the<br />

full bandwidth by conventional process<strong>in</strong>g. Process<strong>in</strong>g D results <strong>in</strong> an ’<strong>in</strong>ner envelop’ of<br />

’up’ and ’down’ biased estimate with<strong>in</strong> the same estimates by Process C (Figure 6.18).<br />

The results from station KG02 (Figure 6.19) shows large deviations between the biased


116<br />

estimates for periods > 1 seconds. Here also the robust band average results <strong>in</strong> reduced<br />

split between both estimations. Perhaps the use of robust band averag<strong>in</strong>g (Process D) is<br />

well expla<strong>in</strong>ed by the results from the process<strong>in</strong>g of long period <strong>data</strong> from station OK18<br />

(Figure 6.20). The ρ xy values at the periods > 4 seconds are obviously affected by poor<br />

s/n ratio as evidenced by the large split between the ’up’ and ’down’ biased estimates<br />

of Process C, especially for periods > 10 seconds. The ρ yx vales <strong>in</strong> the same period<br />

range are not that affected with noise, as the up and down biased estimates follow same<br />

curve. However, is evident <strong>in</strong> ρ yx values between 0.1 to 10 Hz. Here aga<strong>in</strong>, the robust<br />

band averag<strong>in</strong>g (Process D) resulted <strong>in</strong> reduced split between the up and down biased<br />

estimate.<br />

6.4 Results and discussion<br />

From the validation experiments carried out <strong>in</strong> the earlier section, it is established that<br />

the robust band averag<strong>in</strong>g reduces bias <strong>in</strong> the estimation of ρ values and may even improve<br />

the remote referenced solution for transfer functions. In this section a comb<strong>in</strong>ation<br />

of robust band averag<strong>in</strong>g (to produce robust spectral matrix) and robust section averag<strong>in</strong>g<br />

(to estimate transfer functions) (Process D) were used to the same set of sites<br />

used for validation. For discussion the ρ and φ values and their associated telluric predicted<br />

coherencies (positive square root of coherence function, γ 2 ) derived from the above<br />

comb<strong>in</strong>ation were compared with that of conventional spectrum estimation with robust<br />

transfer function estimation (Process C). The long period MT <strong>data</strong> were subjected to<br />

ANN edit<strong>in</strong>g as detailed <strong>in</strong> § 5. The results are shown <strong>in</strong> Figures 6.21 to 6.24. The top<br />

panel <strong>in</strong> each Figure compares the coherencies, the middle panel compares the apparent<br />

resistivities and the bottom panel compares the phase values.<br />

6.4.1 VP14<br />

This site was occupied <strong>in</strong> the southern part of the measurement area where the ma<strong>in</strong><br />

rock type exposed is gneiss ( see § 4.2). The predicated coherencies from process D are<br />

consistently higher than that process C. The <strong>in</strong>crease <strong>in</strong> predicted coherencies are more<br />

clear <strong>in</strong> the frequency range 0.1 Hz to 10 Hz. This is also attested <strong>in</strong> the form of improved<br />

estimates of ρ and φ values <strong>in</strong> the bottom panels. However, <strong>in</strong> the xy coherencies from<br />

process D for periods > 1 seconds are less than compared to Process C. (Figure 6.21(a))<br />

6.4.2 TT08<br />

The noise environment for the station TT08 was described <strong>in</strong> §5, <strong>in</strong> the context of application<br />

of ANN. The improvement <strong>in</strong> the estimation of ρ and φ values by Process D<br />

are evident <strong>in</strong> the middle and bottom panel of Figure 6.21(b). Frequency to frequency<br />

variability of φ values were remarkably reduced for the periods > 1 sec. The Process<br />

D also <strong>in</strong>creased the values of predicted coherencies as shown <strong>in</strong> the top panel. The<br />

improvement <strong>in</strong> φ values <strong>in</strong>dicate a better estimate of cross spectra sets, where as the<br />

reduction <strong>in</strong> bias may be attributed to the better estimation of auto spectra. This is


117<br />

stated to emphasize that robust band average give cleaner spectral matrices, from which<br />

the robust section average can further, improve upon.<br />

6.4.3 TT04<br />

This station, also falls <strong>in</strong> the <strong>in</strong>dustrial belt near the Cauvery river. The xy components<br />

(ρ and φ) between 0.1 second and 10 seconds were seriously affected by a coherent noise<br />

source, which distorted these curves as shown <strong>in</strong> Figure 6.22(a). While discuss<strong>in</strong>g the<br />

results of robust process<strong>in</strong>g with jackknife <strong>in</strong>itialization (Process A) (§ 6.7), it was seen<br />

that Process A resulted <strong>in</strong> recover<strong>in</strong>g the heavily distorted ρ values of this station. However,<br />

Process A <strong>in</strong>duced a bias <strong>in</strong> long period ρ yx values. The robust process<strong>in</strong>g with<br />

robust band averag<strong>in</strong>g and jackknife <strong>in</strong>itial guess substantially reduced this bias as seen<br />

<strong>in</strong> the middle panel of Figure 6.22(a). The improvement <strong>in</strong> predicted coherencies are<br />

evident, perhaps the difference is less compared to TT08 or VP14<br />

6.4.4 OK18<br />

The station lies <strong>in</strong> the south end of the measurement corridor (see Figure 4.1). Electric<br />

water pumps for domestic and small scale agricultural purposes were operational <strong>in</strong> the<br />

area, dur<strong>in</strong>g the measurements. This resulted <strong>in</strong> <strong>in</strong>creased noise activity <strong>in</strong> the electric<br />

fields, ma<strong>in</strong>ly for long period <strong>data</strong>. The high frequency <strong>data</strong> were <strong>in</strong>fluenced by power<br />

l<strong>in</strong>e related noise. It can be seen from the plot (Figure 6.22(b)) that the xy component of<br />

the <strong>data</strong> are more affected than yx components. The heavy bias for ρ xy values at periods<br />

> 1 second are associated by a low <strong>in</strong> predicted coherency values confirm this. Process<br />

D gave reduced bias <strong>in</strong> ρ values and smoother φ values, which are consistent with their<br />

orthogonal counterparts. The <strong>in</strong>crease <strong>in</strong> predicted coherencies brought by Process D is<br />

also evident <strong>in</strong> Figure 6.22(b).<br />

6.4.5 JN12<br />

The MT station JN12, lies <strong>in</strong> the northern part of the measurement corridor, near the<br />

town Dharmapuri. The area has as average long period noise/(signal + noise) ratio of<br />

0.35 (Figure 4.8) and relatively less noisy. The resistivity values monotonously reduce<br />

with <strong>in</strong>crease <strong>in</strong> period (Figure 6.23(a)). The Process C poorly estimated the ρ xy values<br />

especially <strong>in</strong> the frequency range 100 Hz and 10Hz. This resulted from power related<br />

noise that biased up the ρ xy values near 50 Hz. The robust band and section averag<strong>in</strong>g<br />

with jackknife <strong>in</strong>itial guess (process D) gave a better estimation of all but two ρ xy values<br />

<strong>in</strong> the band. However, there was not much improvement obta<strong>in</strong>ed for long period φ values<br />

from robust process<strong>in</strong>g. For ρ xy values at periods > 4 seconds, there small but perceptible<br />

reduction <strong>in</strong> bias by Process D. This is also attested by the <strong>in</strong>creased predicted coherencies<br />

values as shown <strong>in</strong> the top panel of Figure 6.23(a).


118<br />

6.4.6 VP16<br />

The site falls <strong>in</strong> the southern portion of the measurement corridor, over exposed gneisses<br />

(Figure 4.1). The moderate geomagnetic <strong>in</strong>dices (Figure 4.6) for the period of survey<br />

were favorable for the occurance of MT signals <strong>in</strong> the long period. Average noise ratio<br />

(Figure 4.8) of 0.4 for the long period <strong>data</strong> also <strong>in</strong>dicate that the site is relatively noise<br />

free. The Process C resulted <strong>in</strong> almost same estimate as Process D as shown <strong>in</strong> the middle<br />

and bottom panel of Figure 6.23(b). The predicated coherencies for the estimates from<br />

process D do have higher values as compared to the that of process D. However, near<br />

10 seconds, the robust band averag<strong>in</strong>g resulted <strong>in</strong> a small k<strong>in</strong>k (upward) which was not<br />

there <strong>in</strong> the estimates from Process C.<br />

6.4.7 OK16<br />

This station lies <strong>in</strong> the southern end of the measurement corridor. The site is away from<br />

the major cultural noise sources as shown <strong>in</strong> Figure 4.8. However, the effect of noise<br />

from other sources are evident <strong>in</strong> the ρ and φ values (Figure 6.24(a)) from Process C.<br />

Remarkable improvement <strong>in</strong> the smoothness of φ values resulted from the application<br />

of process D. The improvement is also evident <strong>in</strong> ρ xy values , as Process D resulted <strong>in</strong><br />

smoother and less biased estimates especially <strong>in</strong> the frequency range 1 Hz to 10 Hz. This<br />

is also attested by the <strong>in</strong>creased telluric predicted coherencies as shown <strong>in</strong> the top panel<br />

of Figure 6.24(a).<br />

6.4.8 VP12<br />

Situated far south of the major <strong>in</strong>dustrial zone along the Cauvery river, the station VP12<br />

is relatively free from cultural noise sources. The ρ and φ values from Process C and<br />

D are almost same, except for four frequencies <strong>in</strong> ’dead’ band (period 1 second and 10<br />

seconds) as shown <strong>in</strong> Figure 6.24(b). In this frequency range, the Process D produced<br />

ρ values that are higher than that of Process C. However, as there are no appreciable<br />

<strong>in</strong>crease <strong>in</strong> predicted coherencies from Process D <strong>in</strong> this period range, <strong>in</strong>creased estimates<br />

<strong>in</strong> ρ values cannot be confirmed as an improvement. Between 10 and 100 Hz, the Process<br />

D resulted <strong>in</strong> better estimation of ρ xy values as compared to that of Process C. The<br />

predicted coherencies from Process D are slightly higher than that of Process D <strong>in</strong> the<br />

frequency range 10 Hz to 0.1 Hz.<br />

6.4.9 Comparison of results from the Vellar - Palani profile <strong>in</strong><br />

SGT<br />

The results presented <strong>in</strong> Figure 6.25 compare the MT <strong>data</strong> sections along the Vellar -<br />

Palani profile (the broken l<strong>in</strong>e <strong>in</strong> the Figure 4.1) over the Southern Granulite Terra<strong>in</strong>.<br />

The left panels show the ρ xy and φ xy sections from ”ProcMT”, the MT <strong>data</strong> process<strong>in</strong>g<br />

software of Metronix (Ell<strong>in</strong>ghaus [1997]) and the right panels show the results from the<br />

proposed approach. The <strong>data</strong> are not smoothed or contoured to better represent the<br />

reality. The <strong>data</strong> that are miss<strong>in</strong>g/ out of range are represented as white patches. The


119<br />

length of the FFT and parzen w<strong>in</strong>dows were kept identical for both schemes. For ProcMT,<br />

the available time series were manually edited and the MT impedances were estimated<br />

by Robust-M process<strong>in</strong>g. While <strong>us<strong>in</strong>g</strong> the scheme proposed here, the time series were<br />

not manually edited. The long period <strong>data</strong> were subjected to ANN edit<strong>in</strong>g. The cross<br />

and auto spectra were computed by robust band averag<strong>in</strong>g. The MT impedance were<br />

estimated by robust process<strong>in</strong>g with jackknife <strong>in</strong>itialization.<br />

The resistivity <strong>data</strong> ranges from 10 1 to 10 5 Ohm-m, typical of the highly resistive crust.<br />

The high frequency <strong>data</strong> (>1 Hz) shows a relatively resistive crust <strong>in</strong> the northern part<br />

of the profile (especially north of station VP10). However, po<strong>in</strong>t to po<strong>in</strong>t discont<strong>in</strong>uities<br />

are evident along majority of the stations for the <strong>data</strong> processed by ProcMT (Figure<br />

6.25a). Figure 6.25b shows the ρ xy section derived from the scheme proposed. A greater<br />

smoothness for <strong>data</strong> are now evident between stations, and the two major resistivity<br />

blocks are clearly visible. In addition, the smaller resistive structures with<strong>in</strong> the high<br />

frequency <strong>data</strong> below the southern part of the profile are more evident. The improvements<br />

are not so evident for the low frequency <strong>data</strong> (


120<br />

to be accounted <strong>in</strong> frequency band averag<strong>in</strong>g as well. The bias effect <strong>in</strong> ρ values caused<br />

by ignor<strong>in</strong>g this factor is demonstrated on MT <strong>data</strong> collected from SGT. The robust<br />

weight<strong>in</strong>g approach proposed <strong>in</strong> this chapter can alleviate this problem to an extent as<br />

demonstrated by its application to the field <strong>data</strong>. The robust weight<strong>in</strong>g procedures for<br />

spectral estimation may easily extended to robust process<strong>in</strong>g as well.<br />

In summary, the Robust band averag<strong>in</strong>g deals with the averag<strong>in</strong>g of the adjacent<br />

Fourier coefficients around a target frequency while estimat<strong>in</strong>g the auto and cross powers<br />

(or power spectral densities PSD). As shown <strong>in</strong> Figure 6.12, the apparent resistivity<br />

value has a systematic dependency on the number of harmonics averaged. This means<br />

the conventional way of comput<strong>in</strong>g PSD by band averag<strong>in</strong>g leads to biased estimates<br />

of apparent resistivity values. Robust band averag<strong>in</strong>g, proposed , here accounts for the<br />

abnormal Fourier coefficients while averag<strong>in</strong>g around a target frequency. This clearly<br />

results <strong>in</strong><br />

1. M<strong>in</strong>imiz<strong>in</strong>g the effect due to non Gaussian distribution of Fourier coefficients. Once<br />

the spectral coefficients are robustly weighted, they are more appropriate for a LS<br />

averag<strong>in</strong>g (Figure 6.13)<br />

2. The robust band averag<strong>in</strong>g results <strong>in</strong> improved (bias less) estimates of power spectra<br />

(Figure 6.14)<br />

3. F<strong>in</strong>ally the MT transfer functions computed from the above PSDs have low bias<br />

compared to the conventional process<strong>in</strong>g (Figure 6.16 and §6.3.5)<br />

It may be concluded that the two new approaches made <strong>in</strong> the robust process<strong>in</strong>g of<br />

magnetotelluric <strong>data</strong> clearly showed it effectiveness while applied to a large volume of<br />

<strong>data</strong> collected from SGT and are easily adaptable to the conventional MT time series<br />

process<strong>in</strong>g rout<strong>in</strong>e.


121<br />

Figure 6.14: Comparison of robust and least square band averag<strong>in</strong>g for different frequencies<br />

and band (Parzen) radius for station VP13. Magnitude of H x Hy<br />

∗ (nT 2 /Hz) cross<br />

spectra are plotted for all the cases. X-axis show the frequency b<strong>in</strong>s. Solid l<strong>in</strong>e represents<br />

robust average and broken l<strong>in</strong>e represents Least Square average.The left column represent<br />

the the band averag<strong>in</strong>g for different target frequencies, where as the right column<br />

represents the band averag<strong>in</strong>g with different radius length for the same target frequency.<br />

See text for discussion


122<br />

Figure 6.15: Flow chart represent<strong>in</strong>g robust band averag<strong>in</strong>g. Gray area represents the<br />

proposed process<strong>in</strong>g rout<strong>in</strong>e (a) Shows the different steps <strong>in</strong> robustly estimat<strong>in</strong>g the<br />

spectral matrix from raw time series Schematic diagrams (b) to (e) shows four process<strong>in</strong>g<br />

schemes def<strong>in</strong>ed. SS s<strong>in</strong>gle station, RR remote reference, RB robust band averag<strong>in</strong>g,<br />

LS Least square


123<br />

Figure 6.16: Results of s<strong>in</strong>gle station (SS) and remote reference (RR) process<strong>in</strong>g for<br />

station VP10. Process C (SS) and E (RR) use least squares band averag<strong>in</strong>g, whereas<br />

process D (SS) and F (RR) use robust band averag<strong>in</strong>g. The MT transfer functions were<br />

derived from a robust section averag<strong>in</strong>g §6.2.4 from the spectra sets produced by Process<br />

C to F. See text for discussion.<br />

Figure 6.17: Comparison of upward and downward biased estimates of apparent resistivities<br />

for JN12 for processes C and D (a) XY component and (b) YX component. UP - up<br />

biased (E- reference) and DN down biased (H-references). Symbols are same for both<br />

plots. See text for discussion.


124<br />

Figure 6.18: Comparison of upward and downward biased estimates of apparent resistivities<br />

for VP16 for processes C and D (a) XY component and (b) YX component. UP<br />

- up biased (E- reference) and DN down biased (H-references). Symbols are same for<br />

both plots. See text for discussion.<br />

Figure 6.19: Comparison of upward and downward biased estimates of apparent resistivities<br />

for KG02 for processes C and D (a) XY component and (b) YX component. UP<br />

- up biased (E- reference) and DN down biased (H-references). Symbols are same for<br />

both plots. See text for discussion.


125<br />

Figure 6.20: Comparison of upward and downward biased estimates of apparent resistivities<br />

for OK18 for processes C and D (a) XY component and (b) YX component. UP<br />

- up biased (E- reference) and DN down biased (H-references). Symbols are same for<br />

both plots. See text for discussion.


126<br />

Figure 6.21: Comparison of telluric predicted coherency, apparent resistivity and phase<br />

values from robust process<strong>in</strong>g with and without robust band averag<strong>in</strong>g for stations VP14<br />

and TT08. The explanation for symbols are given <strong>in</strong> legend. See text for discussion.


127<br />

Figure 6.22: Comparison of telluric predicted coherency, apparent resistivity and phase<br />

values from robust process<strong>in</strong>g with and without robust band averag<strong>in</strong>g for stations TT04<br />

and OK18. The explanation for symbols are given <strong>in</strong> legend. See text for discussion.


128<br />

Figure 6.23: Comparison of telluric predicted coherency, apparent resistivity and phase<br />

values from robust process<strong>in</strong>g with and without robust band averag<strong>in</strong>g for stations JN12<br />

and VP16. The explanation for symbols are given <strong>in</strong> legend. See text for discussion.


129<br />

Figure 6.24: Comparison of telluric predicted coherency, apparent resistivity and phase<br />

values from robust process<strong>in</strong>g with and without robust band averag<strong>in</strong>g for stations OK16<br />

and VP12. The explanation for symbols are given <strong>in</strong> legend. See text for discussion.


130<br />

a) b)<br />

c) d)<br />

Figure 6.25: Comparison of results from two different process<strong>in</strong>g scheme for MT <strong>data</strong>:<br />

Left panel shows the results from ProcMT and the right panel shows the results from the<br />

robust process<strong>in</strong>g and ANN edit<strong>in</strong>g.


Chapter 7<br />

Discussion and conclusions<br />

131


132<br />

7.1 Introduction<br />

<strong>Magnetotelluric</strong>s has been established as a versatile exploration tool, serv<strong>in</strong>g the needs<br />

of shallow – resource exploration as well as deeper crustal and lithospheric studies (Vozoff<br />

[1991], Jones [1992]) Cont<strong>in</strong>u<strong>in</strong>g improvements <strong>in</strong> <strong>in</strong>strumentation, <strong>data</strong> acquisition<br />

procedures, <strong>data</strong> process<strong>in</strong>g and modell<strong>in</strong>g of MT measurements provide a fertile environment<br />

for experimentation and application. In order to determ<strong>in</strong>e the sub-surface<br />

electrical conductivity, it would be necessary to obta<strong>in</strong> the ‘best’ possible transfer functions<br />

from the <strong>data</strong> collected. Developments <strong>in</strong> this aspect of MT have been outl<strong>in</strong>ed <strong>in</strong><br />

Chapters 3, 4, 5 and 6. Just as <strong>in</strong> the preparations for a MT experiment it is possible<br />

to consider several scenarios, models and cont<strong>in</strong>gencies and attempt to plan the optimal<br />

survey. Similarly it is observed <strong>in</strong> this work that while process<strong>in</strong>g the time series, it is<br />

necessary to study all the characteristics possible, <strong>in</strong> space, time and frequency doma<strong>in</strong><br />

and attempt through a suite of process<strong>in</strong>g rout<strong>in</strong>es to obta<strong>in</strong> the ‘best’ or optimal curves.<br />

It is concluded here that the study of <strong>data</strong> characteristics after the survey also shed light<br />

on the most suitable methods of process<strong>in</strong>g. It is clearly demonstrated <strong>in</strong> Chapters 5 and<br />

6 that the same rout<strong>in</strong>e may not be optimal for different <strong>data</strong> sets with different noise<br />

characterization (Egbert and Livelybrooks [1996]). Further, for the same measurement,<br />

one process would elim<strong>in</strong>ate some types of noise and another process the other.<br />

Significant developments for MT transfer function estimation <strong>in</strong> magnetotellurics, <strong>in</strong><br />

the past three decades were the <strong>in</strong>troduction of remote reference process<strong>in</strong>g (Gamble et al.<br />

[1979]) and robust process<strong>in</strong>g (Egbert and Booker [1986]). This was complementary to the<br />

requirement of more precise MT estimates by the new and sophisticated <strong>data</strong> model<strong>in</strong>g<br />

algorithms developed simultaneously. Advances <strong>in</strong> both these fields accelerated <strong>in</strong> 1980’s<br />

with the availability of cheap and fast microprocessors. Thus the tremendous development<br />

<strong>in</strong> MT methodologies <strong>in</strong> the past three decades, greatly owe to the <strong>advances</strong> <strong>in</strong> digital<br />

comput<strong>in</strong>g. The effective depth of MT prob<strong>in</strong>g ranges from few hundred meters typical<br />

for geo-technical <strong>in</strong>vestigation (Garcia and Jones [2002]) to more that hundred kilometers<br />

for deep crustal studies (Chen et al. [1996]). This greater range of depths demands wider<br />

bandwidth of measurements, which <strong>in</strong> turn demands better <strong>in</strong>struments/sensors. To<br />

estimate the transfer function for full bandwidth, as precisely as possible, <strong>data</strong> process<strong>in</strong>g<br />

algorithms need to be improved as well.<br />

7.2 Summary<br />

The time series process<strong>in</strong>g steps <strong>in</strong>troduced <strong>in</strong> this thesis were applied to the follow<strong>in</strong>g:<br />

1. Identify<strong>in</strong>g the noise sources <strong>in</strong> space doma<strong>in</strong> (§ 4),<br />

2. Discrim<strong>in</strong>ation of bad segments of time series <strong>data</strong> aga<strong>in</strong>st good segments <strong>in</strong> time<br />

doma<strong>in</strong> (§ 5) and<br />

3. Effectively down weight<strong>in</strong>g the unusual <strong>data</strong> <strong>in</strong> a robust sense <strong>in</strong> time - frequency<br />

doma<strong>in</strong> (§ 6).


133<br />

In the first <strong>in</strong>stance, an attempt was made to characterize the signal and noise components<br />

<strong>in</strong> the MT <strong>data</strong> collected from SGT. The MT <strong>data</strong> were analyzed <strong>in</strong> time, frequency<br />

and space doma<strong>in</strong> <strong>in</strong> this regard. The comparison of <strong>in</strong>dices of geomagnetic activity with<br />

the long period coherence showed their general agreement. The effect of noise from an <strong>in</strong>dustrial<br />

zone was evident <strong>in</strong> the long period <strong>data</strong>, while analyz<strong>in</strong>g the noise/(signal+noise)<br />

for all the sites <strong>in</strong> relation with the known cultural noise sources.<br />

The second approach was made, <strong>in</strong> response to another requirement of MT exploration<br />

- an automated method of process<strong>in</strong>g. For example, the <strong>data</strong> acquired <strong>in</strong> SGT, amounted<br />

to ∼10 6 real numbers of <strong>data</strong> per site, and with 80 such stations, the magnitude of <strong>data</strong><br />

reduction <strong>in</strong> MT is a major task. In such scenario, the edit<strong>in</strong>g of MT time series becomes<br />

challeng<strong>in</strong>g and may consume more that 80 % of the time spent on time series <strong>analysis</strong>.<br />

The artificial neural network approach <strong>in</strong>troduced <strong>in</strong> § 5, is a very promis<strong>in</strong>g solution to<br />

this. It not only proved to be a good replacement to manual <strong>in</strong>spection of time series,<br />

but turns out to be a stable and objective decision-maker as well.<br />

In the third approach made <strong>in</strong> this thesis, the robust process<strong>in</strong>g algorithm was critically<br />

reviewed and two steps were <strong>in</strong>troduced to further improve this type of estimators<br />

<strong>in</strong> the presence of severe noise contam<strong>in</strong>ation as found <strong>in</strong> SGT. The two approaches<br />

namely Jackknife and Robust band averag<strong>in</strong>g were used to replace simple Least Square<br />

procedures, that were still <strong>in</strong> use <strong>in</strong> two steps of robust process<strong>in</strong>g. The effects of a few<br />

outliers on MT <strong>data</strong> are well recognized and robust process<strong>in</strong>g was essentially <strong>in</strong>troduced<br />

to tackle this (Eiesel & Egbert, 2001, Chave et al 1987, Chave & Thomson, 1989). Some<br />

of the computational steps reta<strong>in</strong>ed LS averag<strong>in</strong>g, overlook<strong>in</strong>g its effect. The effect of<br />

unusual <strong>data</strong> at the frequency band averag<strong>in</strong>g level has been evaluated <strong>in</strong> this thesis and<br />

a robust band averag<strong>in</strong>g technique suggested to alleviate this problem. In the same manner,<br />

Jackknife estimates were used to f<strong>in</strong>d an <strong>in</strong>itial guess for robust process<strong>in</strong>g, replac<strong>in</strong>g<br />

the commonly employed Least Squares.<br />

Holistically, the last two approaches may be considered as a natural outgrowth of<br />

MT time series process<strong>in</strong>g procedures, aga<strong>in</strong>st the back drop of <strong>in</strong>creas<strong>in</strong>g computational<br />

power and improv<strong>in</strong>g <strong>in</strong>strumentation. Neural networks, with their computational complexity,<br />

were rarely used <strong>in</strong> geophysics <strong>in</strong> 1980’s. With the availability of fast and cheap<br />

microprocessors, neural networks are widely used <strong>in</strong> a variety of applications <strong>in</strong> geophysics<br />

today. When <strong>in</strong>troduced, robust process<strong>in</strong>g (<strong>in</strong> MT) concentrated on the averag<strong>in</strong>g of<br />

transfer functions. Later, by go<strong>in</strong>g back one step <strong>in</strong> the process<strong>in</strong>g algorithm, Egbert<br />

[1997] proposed robust estimation of global spectral density matrix (SDM) from all the<br />

available SDMs. The approach made <strong>in</strong> this chapter may treated as one more step back<br />

<strong>in</strong> the process<strong>in</strong>g level, where<strong>in</strong>, the <strong>in</strong>dividual SDMs from a s<strong>in</strong>gle time section is also<br />

estimated <strong>in</strong> robust sense.<br />

7.3 Neural Network and Robust process<strong>in</strong>g<br />

Neural network approach <strong>in</strong>troduced <strong>in</strong> Chapter 5 and the robust process<strong>in</strong>g techniques<br />

(Chapter 6) discrim<strong>in</strong>ate signal aga<strong>in</strong>st noise. The fundamental differences between these<br />

two methods are<br />

1. ANN based edit<strong>in</strong>g is performed <strong>in</strong> time doma<strong>in</strong> alone, where as robust process<strong>in</strong>g


134<br />

is done on spectra sets frequency doma<strong>in</strong>, generated from short time segments.<br />

2. ANN based edit<strong>in</strong>g is used to discrim<strong>in</strong>ate signal patterns, correlation structure<br />

and amplitude ratio to search for and select signal elements from the subdivided<br />

time series. Robust process<strong>in</strong>g attempts to down weight <strong>data</strong> that deviate from the<br />

fit with the transfer function (large residuals).<br />

3. Decision of ANN edit<strong>in</strong>g is b<strong>in</strong>ary, i.e. it either accepts or rejects the time segment.<br />

Robust process<strong>in</strong>g algorithm tries to m<strong>in</strong>imize the effect of larger residuals, with<br />

proportional weights and the decision is analog. Tukey weighs at the f<strong>in</strong>al iteration<br />

<strong>in</strong> robust process<strong>in</strong>g may be considered b<strong>in</strong>ary. However, it affects only a part of<br />

the <strong>data</strong>.<br />

4. ANN requires previously saved ’<strong>in</strong>telligence’ (weights) to perform process<strong>in</strong>g, where<br />

as robust process<strong>in</strong>g requires none.<br />

In the figure 7.1 & 7.2, the results of process<strong>in</strong>g MT <strong>data</strong> from two sites are given. The<br />

MT time series collected were subjected to the robust process<strong>in</strong>g algorithm described <strong>in</strong><br />

§ 6 (Process D), with and without prior ANN edit<strong>in</strong>g. The derived apparent resistivity &<br />

phase for stations JN10 and OK18 are plotted <strong>in</strong> Figure 7.1(a) &7.2(a). In Figure 7.1(b)<br />

& 7.2(b), the comparison of neural network pick<strong>in</strong>g and robust weights are compared<br />

for all the stacks. The upper panel <strong>in</strong> Fig Figure 7.1(b) & 7.2(b) shows the neural<br />

network decision as a function of stacks. In the lower panels, the robust weights from<br />

the f<strong>in</strong>al step <strong>in</strong> robust process<strong>in</strong>g (§ 6) are presented as a function of frequency and<br />

stacks. Dark areas <strong>in</strong>dicates good <strong>data</strong>, white <strong>in</strong>dicates bad <strong>data</strong> and the gray levels<br />

<strong>in</strong> between <strong>in</strong>dicate <strong>in</strong>termediate <strong>data</strong>. As can be deduced from these plots, the robust<br />

process<strong>in</strong>g with and without NN edit<strong>in</strong>g at station JN10 seems to yield the same results.<br />

On closer <strong>in</strong>spection of the network pick<strong>in</strong>g and robust weights <strong>in</strong> same figure(Figure<br />

7.1(b) & 7.2(b)), it is seen that from stacks 1 to 15, the neural network consistently<br />

gave zero and the correspond<strong>in</strong>g robust weights also gave smallest values, throughout<br />

time and frequency scale with the exception of stack 3 & 4. In the same way stacks<br />

around 30 also gave similar results for robust and NN process<strong>in</strong>g. It may be concluded<br />

that the similar results from robust process<strong>in</strong>g with and without neural network edit<strong>in</strong>g<br />

is due to fact that the signal discrim<strong>in</strong>ation were similar, though the approaches were<br />

made <strong>in</strong> different doma<strong>in</strong>s. In the second example the process<strong>in</strong>g results from stations<br />

OK18 is presented. The robust process<strong>in</strong>g alone resulted <strong>in</strong> biased estimate of ρ xy values<br />

and scattered φ xy values (open circles) with large error bars. However robust process<strong>in</strong>g<br />

after neural network edit<strong>in</strong>g resulted <strong>in</strong> better estimation of ρ xy and φ xy (stars). (Figure<br />

7.2(a)). On verify<strong>in</strong>g the neural network output and robust weights as a function of stacks<br />

(Fig 7.2(b)), it is observed that the two considerably vary especially between stacks 20<br />

and 60. The neural network rejected around 12 stacks <strong>in</strong> this <strong>in</strong>terval, whereas the robust<br />

weights are near the value 1 for all frequencies. However between stacks 60 and 80, the<br />

robust procedures down weight around 5 stacks, where as the neural network did not<br />

reject any. The reason is that the outliers and/or coherent noises which are obvious <strong>in</strong><br />

time doma<strong>in</strong> between stacks 20 and 60, failed to produce larger outliers and thus were<br />

not down weighted by robust procedures.


135<br />

a)<br />

b)<br />

Figure 7.1: Comparison of robust process<strong>in</strong>g (RB) with and without neural network (NN)<br />

edit<strong>in</strong>g for Station JN12. (a) Apparent resistivity and phase values from both process<strong>in</strong>g<br />

schemes. (b) Comparison of robust and neural network weights.<br />

Those <strong>data</strong>, which the neural network passed on to robust process<strong>in</strong>g between stacks<br />

60 an 80 with noise that was not evident <strong>in</strong> time doma<strong>in</strong>, were down weighted by robust<br />

procedures. This clearly shows the need to comb<strong>in</strong>e the two techniques <strong>in</strong> order to<br />

discrim<strong>in</strong>ate / down weight a majority of noisy <strong>data</strong>.


136<br />

a)<br />

b)<br />

Figure 7.2: Comparison of robust process<strong>in</strong>g (RB) with and without neural network (NN)<br />

edit<strong>in</strong>g for Station OK18. (a) Apparent resistivity and phase values from both process<strong>in</strong>g<br />

schemes. (b) Comparison of robust and neural network weights.<br />

7.4 Conclusions<br />

Major conclusions that can be drawn from the thesis are<br />

1. The magnetotelluric <strong>data</strong> collected from SGT encountered severe <strong>in</strong>terference from<br />

cultural noise sources, and their effects are fully understood only if analyzed <strong>in</strong> a


137<br />

multi- doma<strong>in</strong> approach. Rout<strong>in</strong>e application of any one of the <strong>data</strong> process<strong>in</strong>g<br />

methods is bound to give <strong>in</strong>ferior results <strong>in</strong> such cases.<br />

2. In time doma<strong>in</strong>, the application of artificial neural networks clearly demonstrated<br />

its usefulness as an effective signal discrim<strong>in</strong>ator, even <strong>in</strong> the presence of large noise<br />

contam<strong>in</strong>ation.<br />

3. The prom<strong>in</strong>ent weak po<strong>in</strong>ts <strong>in</strong> robust process<strong>in</strong>g of MT <strong>data</strong> are the uses of least<br />

square estimators <strong>in</strong> two of the steps, namely estimation of spectra and transfer<br />

function <strong>in</strong>itialization.<br />

4. The application of a non-parametric estimator called Jackknife to <strong>in</strong>itialize MT<br />

transfer functions, resulted <strong>in</strong> improved robust estimation of MT transfer functions.<br />

Robust band averag<strong>in</strong>g reduces the uncerta<strong>in</strong>ties <strong>in</strong> estimat<strong>in</strong>g the spectra of<br />

magnetic and electric field components. Robustly estimated spectra always result<br />

<strong>in</strong> better MT transfer functions.<br />

5. On apply<strong>in</strong>g these methods to a large numbers MT stations <strong>in</strong> SGT, the effectiveness<br />

of a comb<strong>in</strong>ed use of neural network and robust process<strong>in</strong>g, <strong>in</strong> elim<strong>in</strong>at<strong>in</strong>g<br />

severe noise was established and demonstrated.<br />

7.5 Suggestions for future work<br />

The natural signal sources for MT <strong>data</strong> >1Hz (AMT) are the world wide thunderstorms<br />

(§ 2). These signals are manifested <strong>in</strong> high frequency MT time series (band1 of table<br />

2.1) as transient bursts. It is well known that the MT apparent resistivity values are<br />

systematically biased down at frequencies > 10 3 Hz. The frequency range 10 3 to 10 4 Hz<br />

conta<strong>in</strong>s the ‘dead band’ of AMT (Garcia and Jones [2002]), where the natural signal<br />

energy is very low compared to the higher and lower frequencies (Figure 2.1). The<br />

transient signal such as we found <strong>in</strong> Fig 2.4, has enough spectral energy <strong>in</strong> the AMT<br />

dead band. However, there are two reasons, why they fail to improve the MT estimates<br />

<strong>in</strong> the ‘dead band’: 1) The MT signals appear as transient envelopes of signals, on the<br />

background of power transmission harmonic noise (and 50 and 150Hz, <strong>in</strong> this case) and<br />

their occurrences are random. 2) It is widely recognized that FFT fail to perform spectral<br />

<strong>analysis</strong> on transient signals, as the amplitude spectra thus derived do not give any<br />

temporal <strong>in</strong>formation. Due to these reasons, the sparse signal activity <strong>in</strong> this frequency<br />

range gets down weighted <strong>in</strong> spectral estimation as well as <strong>in</strong> robust process<strong>in</strong>g. It would<br />

be worthwhile to explore the use of the wavelet transform (Kumar and Foufoula-Georgiou<br />

[1997] to discrim<strong>in</strong>ate transient signals <strong>in</strong> the high frequency MT time series aga<strong>in</strong>st the<br />

majority of background noise. Wavelet transforms have the unique ability to reta<strong>in</strong><br />

temporal characteristics, while giv<strong>in</strong>g the spectral <strong>in</strong>formation. It is thus important to<br />

estimate MT transfer function <strong>in</strong> dilatation – translation doma<strong>in</strong> of wavelet transform<br />

(somewhat equivalent to frequency-time doma<strong>in</strong>), as it allows us to effectively isolate<br />

energetic signal events both <strong>in</strong> frequency and time doma<strong>in</strong>.<br />

Thus it may be seen that cont<strong>in</strong>u<strong>in</strong>g improvement <strong>in</strong> MT <strong>data</strong> acquisition, process<strong>in</strong>g<br />

and model<strong>in</strong>g complement each other with the need for ever-more accurate imag<strong>in</strong>g of the


138<br />

Earth’s sub-surface structure. In the case of MT <strong>data</strong> process<strong>in</strong>g, novel applications of<br />

computational and statistical <strong>advances</strong> <strong>in</strong> signal process<strong>in</strong>g, cont<strong>in</strong>ue to be implemented<br />

as <strong>in</strong> the work outl<strong>in</strong>ed above. Multi-doma<strong>in</strong> <strong>analysis</strong> of signal discrim<strong>in</strong>ation would be<br />

favored as the benefits of such approaches are validated by more <strong>data</strong> <strong>analysis</strong>.


Bibliography<br />

D. Allsopp, M. Burke, D. Rank<strong>in</strong>, and I. Reddy. A wide band magnetotelluric record<strong>in</strong>g<br />

system. Geophys. Prosp., 22:272–278, 1973.<br />

O. Alt<strong>in</strong>ay, E. Tulunay, and Y. Tulunay. Forecast<strong>in</strong>g ionospheric critical frequency <strong>us<strong>in</strong>g</strong><br />

neural networks. Geophys. Res. Lett., 24:1467 – 1470, 1997.<br />

R. Bailey. Inversion of the geomagnetic <strong>in</strong>duction problem. Proc. Roy. Soc. Lond., 315:<br />

185–194, 1970.<br />

M. Balser and C. Wagner. Observations of earth-ionosphere cavity resonance. Nature,<br />

188:638 – 641, 1960.<br />

R. J. Banks. The effects of non-stationary noise on electromagnetic response estimates.<br />

Geophys. J. Int., 135:553–563., 1998.<br />

J. S. Bendat and A. G. Piersol. Random <strong>data</strong>: Analysis and measurement procedures.<br />

John Wiley & Sons, New York., 1971.<br />

F. K. Boadu. Inversion of fracture density from field seismic velocities <strong>us<strong>in</strong>g</strong> artificial<br />

neural networks. Geophysics, 63:534–545., 1998.<br />

L. Cagniard. Basic theory of the magneto-telluric methods of geophysical prospect<strong>in</strong>g.<br />

Geophysics, 18:605–635., 1953.<br />

C. Calderon-Macias, M. K. Sen, and P. Stoffa. Artificial neural networks for parameter<br />

estimation <strong>in</strong> geophysics. Geophys. Prosp., 48:21–47., 2000.<br />

L. Chaize and M. Lavergne. Signal et bruit en magnetotellurique. Geophys. Prosp, 18:<br />

64–87., 1970.<br />

A. D. Chave and D. J. Thomson. Some comments on magnetotelluric response function<br />

estimation. J. Geophys. Res., 94:1421514225., 1989.<br />

A. D. Chave and D. J. Thomson. A bounded <strong>in</strong>fluence regression estimator based on the<br />

statistics of the hat matrix. Appl. Statist., 52(3):307 – 322., 2003.<br />

A. D. Chave, D. J. Thomson, and M. Ander. On the robust estimation of power spectra,<br />

coherences and transfer functions. J. Geophys. Res., 92:633–648, 1987.<br />

139


140<br />

L. Chen, J. Booker, A. Jones, N. Wu, M. Unsworth, W. Wei, and H. Tan. Electrically<br />

conductive crust <strong>in</strong> southern tibet from <strong>in</strong>depth magnetotelluric survey<strong>in</strong>g. Science,<br />

274:1694–1696., 1996.<br />

J. F. Claerbout and F. Muir. Robust model<strong>in</strong>g with erratic <strong>data</strong>. Geophysics, 38(5):<br />

826–844., 1973.<br />

J. Clarke, G. T. D., W. M. Goubau, R. Koch, and R. Miracky. Remote-reference magnetotellurics<br />

equipment and procedures. Geophys. Prosp., 31:149–170., 1983.<br />

S. C. Constable, R. L. Parker, and C. G. Constable. Occam’s <strong>in</strong>version: A practical algorithm<br />

for generat<strong>in</strong>g smooth models from electromagnetic sound<strong>in</strong>g <strong>data</strong>. Geophysics,<br />

52(3):289–300., 1987.<br />

H. Dai and C. MacBeth. Automatic pick<strong>in</strong>g of seismic arrivals <strong>in</strong> local earthquake <strong>data</strong><br />

<strong>us<strong>in</strong>g</strong> an artificial neural network. Geophys. J. Int., 120:758 – 774., 1995.<br />

D. L. Dekker and L. M. Hastie. Sources of error and bias <strong>in</strong> magnetotelluric depth<br />

sound<strong>in</strong>g of the Bown Bas<strong>in</strong>. PEPI, 25:219–225., 1981.<br />

S. A. Drury, N. B. W. Harris, R. W. Holt, G. F. Reeves-Smith, and R. T. Wightman.<br />

Precambrian tectonics and the evolution of South India. J. Geol., 92:3–20., 1984.<br />

J. Dungey. Electrodynamics of the outer atmosphere <strong>in</strong> the physics of the ionosphere.<br />

The physics of the Ionosphere, The Physical Society (London), page 229, 1955.<br />

W. D. E. Mt/emap Data Interchange Standards, Revision. Society of Exploration Geophysicists,<br />

USA., Revision 1.0, 1988.<br />

B. Efron. The jackknife, the bootstrap and other resampl<strong>in</strong>g plans. Society for Industrial<br />

and Applied Mathematics, Philadelphia, 1982.<br />

G. D. Egbert. Noncausality of the discrete-time magnetotelluric impulse response. Geophysics,<br />

57:1354–1358., 1992.<br />

G. D. Egbert. Robust multiple station magnetotelluric <strong>data</strong> process<strong>in</strong>g. Geophys. J. Int.,<br />

130:475–496., 1997.<br />

G. D. Egbert and J. R. Booker. Robust estimation of geomagnetic transfer function.<br />

Geophys. J. Roy. astr. Soc., 87:173–194., 1986.<br />

G. D. Egbert and D. W. Livelybrooks. S<strong>in</strong>gle station magnetotelluric impedance estimation:<br />

Coherence weight<strong>in</strong>g and the regression M-estimate. Geophysics, 61:964–970,<br />

1996.<br />

M. Eisel and G. D. Egbert. On the stability of magnetotelluric transfer function estimates<br />

and the reliability of their variances. Geophys. J. Int., 144(1):65–82., 2001.<br />

A. Ell<strong>in</strong>ghaus. PROCMT - Users guide, Revision 2. Metronix GmbH, Braunschweig,<br />

Germany, 1997.


141<br />

T. Ernst, E. Y. Sokolva, I. M. Varentsov, and N. G. Golubev. Comparison of two techniques<br />

for magnetotelluric <strong>data</strong> process<strong>in</strong>g <strong>us<strong>in</strong>g</strong> synthetic <strong>data</strong> sets. Acta Geophysica<br />

Polonica, XLIX (2):213–243., 2001.<br />

B. C. Fish and T. Kusuma. A neural network approach to automate velocity pick<strong>in</strong>g.<br />

<strong>in</strong> Society of Exploration Geophysicists 64th annual <strong>in</strong>ternational meet<strong>in</strong>g; Technical<br />

program, expanded abstracts with authors’ biographies, 64:185–188., 1994.<br />

B. P. Flannery, W. H. Press, S. A. Teukolsky, and W. T. Vetterl<strong>in</strong>g. Numerical recipes <strong>in</strong><br />

c. The art of scientific comput<strong>in</strong>g, Cambridge University Press, UK, page 1020, 1992.<br />

T. D. Gamble, W. M. Goubau, and J. Clarke. <strong>Magnetotelluric</strong>s with a remote reference.<br />

Geophysics, 44:53–68, 1979.<br />

X. Garcia, A. Chave, and A. Jones. Robust process<strong>in</strong>g of magnetotelluric <strong>data</strong> from the<br />

auroral zone. J. Geomag. Geoelectr, 49:1451–1468., 1997.<br />

X. Garcia and A. Jones. Atmospheric sources for audio-magetotelluric sound<strong>in</strong>g. Geophysics,<br />

67 (2):448 – 458., 2002.<br />

L. Garth and H. Poor. Detection of non - gaussian signals: a paradigm for modern<br />

statistical signal process<strong>in</strong>g. Proc. IEEE, 82:1060 – 1095., 1994.<br />

M. Ghil, M. R. Allen, M. D. Dett<strong>in</strong>ger, K. Ide, D. Kondrashov, M. E. Mann, A. W.<br />

Robertson, A. Saunders, Y. Iian, F. Varadi, and P. Yiou. Advanced spectral methods<br />

for climatic time series. Rev. <strong>in</strong> Geophysics, 40(1):1–41, 2002.<br />

W. M. Goubau, T. D. Gamble, and J. Clarke. <strong>Magnetotelluric</strong> <strong>data</strong> <strong>analysis</strong>: Removal<br />

of bias. Geophysics, 43:1157–1166., 1978.<br />

R. W. Groom and R. C. Bailey. Decomposition of magnetotelluric impedance tensors <strong>in</strong><br />

the presence of local three-dimensional galvanic distortion. J. Geophys. Res., 94(B2):<br />

1913–1925., 1989.<br />

GSI. Geological and M<strong>in</strong>eral map of Tamil Nadu and Pondicherry (scale 1:0.5 million).<br />

Geological Survey of India, Calcutta, India, 1995.<br />

F. Hampel, E. M. Ronchetti, P. J. Rousseeuw, and W. Stahel. Robust statistics: The<br />

approach based on <strong>in</strong>fluence functions. John Wiley & Sons Inc. , New York, USA.,<br />

1986.<br />

T. Har<strong>in</strong>arayana, D. N. Murty, S. P. E. Rao, C. Manoj, K. Veeraswamy, K. Naganjaneyulu,<br />

K. K. A. Azeez, R. S. Sastry, and G. Virupakshi. <strong>Magnetotelluric</strong> Field<br />

Investigations <strong>in</strong> Puga Geothermal Region, Jammu and Kashmir, India: 1-D Model<strong>in</strong>g.<br />

Geothermal Resources Council Transactions, 122:393–397, 2002.<br />

T. Har<strong>in</strong>arayana, K. Naganjaneyulu, C. Manoj, B. P. K. Patro, S. K. Begum, D. N.<br />

Murthy, Madhusudana-Rao, V. T. C. Kumaraswamy, and G. Virupakshi. <strong>Magnetotelluric</strong><br />

<strong>in</strong>vestigation along Kuppam - Palani geotransect, South India - 2-D model<strong>in</strong>g<br />

results. Memoir Geological Society of India, 50:107–124, 2003.


142<br />

M. Hatt<strong>in</strong>g. The use of <strong>data</strong>-adaptive filter<strong>in</strong>g for noise removal on magnetotelluric <strong>data</strong>.<br />

PEPI, 53:239–254., 1989.<br />

W. Hernandez and J. Jacobs. Discussion on ” enhancement of signal-to-noise ratios <strong>in</strong><br />

magnetotelluric <strong>data</strong>”, by kao dw and rank<strong>in</strong> d. Geophysics, 44:1594 – 1596, 1979.<br />

D. B. Hoover, F. Frischknecht, and C. L. Tippens. Audiomagnetotelluric sound<strong>in</strong>g as a<br />

reconnaissance exploration technique. J. of Geophys. Res., 81:801–809., 1976.<br />

G. M. Jenk<strong>in</strong>s and D. G. Watts. Spectral <strong>analysis</strong> and its applications. Holden-Day, San<br />

Francisco, California., 1968.<br />

A. Jones. Electrical conductivity of the cont<strong>in</strong>ental lower crust. <strong>in</strong> D.M Founta<strong>in</strong> et al.<br />

(eds). The cont<strong>in</strong>ental lower crust, Elsevier:81–143., 1992.<br />

A. Jones, A. D. Chave, G. D. Egbert, D. Auld, and K. Bahr. Comparison of techniques for<br />

magnetotelluric response function estimation. J. Geophys. Res., 94(B10):14201–14213.,<br />

1989.<br />

A. G. Jones. Transformed coherence functions for mutivariate studies. IEEE trans. ASSP,<br />

29 (2):317–319., 1981.<br />

A. G. Jones, B. Olafsdottir, and J. Tikka<strong>in</strong>en. Geomagnetic <strong>in</strong>duction studies <strong>in</strong> Scand<strong>in</strong>avia<br />

III : <strong>Magnetotelluric</strong> observations. Jour. Geophys., 54:35–50., 1983.<br />

A. Junge. Characterization of and correction for cultural noise. Surv. Geophys., 17:361<br />

– 391., 1996.<br />

D. W. Kao and D. Rank<strong>in</strong>. Enhancement of signal to noise ratio <strong>in</strong> magnetotelluric <strong>data</strong>.<br />

Geophysics, 42:103–110., 1977.<br />

R. Karmann. Search coil magnetometers with optimum signal-to-noise ratio. <strong>in</strong> Vozoff,<br />

K. (ed) <strong>Magnetotelluric</strong> methods: SEG Geophysics Repr<strong>in</strong>t Series 5, pages 215–218,<br />

1977.<br />

A. A. Kaufman and G. V. Keller. The magnetotelluric sound<strong>in</strong>g method. Elsevier Scientific<br />

Publish<strong>in</strong>g Company, Amsterdam, 1981.<br />

H. C. Koons and D. J. Gorney. A neural network model of the relativistic electron flux<br />

at geosynchronous orbit. J. Geophys. Res., 96(A4):5549–5556., 1991.<br />

P. Kumar and E. Foufoula-Georgiou. Wavelet <strong>analysis</strong> for geophysical applications. Reviews<br />

<strong>in</strong> Geophysics, 35(4):385–412., 1997.<br />

G. Kunetz. Process<strong>in</strong>g and <strong>in</strong>terpretation of magnetotelluric sound<strong>in</strong>gs. Geophysics, 37:<br />

1005–1021., 1972.<br />

J. C. Larsen. Transfer functions: Smooth robust estimates by least squares and remote<br />

reference methods. Geophys. J. Int., 99:645–663, 1989.


143<br />

J. C. Larsen, R. L. Mackie, A. Manzella, A. Fiordelisi, and S. Rieven. Robust smooth<br />

MT transfer functions. Geophys. J. Int., 124:801–819., 1996.<br />

H. Lundstedt. Solar orig<strong>in</strong> of geomagnetic storms and prediction of storms with the use<br />

of neural networks. Surv. Geophys., 17:561–573., 1996.<br />

F. Luo and R. Unbehauen. Applied neural networks for signal process<strong>in</strong>g. Cambridge<br />

University press, Cambridge, UK., page 123, 1997.<br />

J. Macnae, Y. Lamontagne, and W. G. F. Noise process<strong>in</strong>g technique for time doma<strong>in</strong><br />

electromagnetic systems. Geophysics, 49:934–948., 1984.<br />

T. Madden. Spectral, cross spectral and bispectral <strong>analysis</strong> of low frequency electromagnetic<br />

<strong>data</strong>. ed. Bleil D.F., Pentium Press, New York., 1964.<br />

T. Mahadevan. Deep Cont<strong>in</strong>ental Structure of India: A Review. Geol. Soc. India Memoir,<br />

28:569, 1994.<br />

C. Manoj and N. Nagarajan. The application of artificial neural networks to magnetotelluric<br />

time series <strong>analysis</strong>. Geophys. J. Int., 153:409–423., 2003.<br />

M. Mareschal, R. D. Kurtz, and R. C. Bailey. A review of electromagnetic <strong>in</strong>vestigations<br />

<strong>in</strong> the Kapuskas<strong>in</strong>g Uplift and surround<strong>in</strong>g regions; electrical properties of key rocks.<br />

Canadian Journal of Earth Sciences, 31(7):1042–1051., 1994.<br />

MATLAB. User’s manual, version 6, mathworks <strong>in</strong>c. MathWorks Inc., Natick, USA.,<br />

2001.<br />

M. D. McCormack, Z. D. E., and D. D. W. First-break refraction event pick<strong>in</strong>g and<br />

seismic <strong>data</strong> trace edit<strong>in</strong>g <strong>us<strong>in</strong>g</strong> neural networks. Geophysics, 58:67–78., 1993.<br />

M. L. McCraken, K. G. Oristaglio and G. W. Hohman. M<strong>in</strong>imization of noise <strong>in</strong> electromagnetic<br />

exploration systems. Geophysics, 51:819–832., 1986.<br />

G. W. McNeice and A. G. Jones. Multi-site, multi-frequency tensor decomposition of<br />

magnetotelluric <strong>data</strong>. Geophysics, 66:158–173, 2001.<br />

W. Menke. Geophysical <strong>data</strong> <strong>analysis</strong>, discrete <strong>in</strong>verse theory. Academic Press Inc,<br />

Orlando, Florida, USA., 58:67–78, 1984.<br />

Metronix. Geophysical Measurement System (GMS05). Operat<strong>in</strong>g Manual, Revision 3.1,<br />

Metronix GmbH, Germany, page 112, 1997.<br />

D. C. Mishra. Geophysical evidence for a thick crust south of Palghat-Tiruchi gap <strong>in</strong> the<br />

high grade terra<strong>in</strong>s of South India. Jour. Geol. Soc. India, 33:79–81., 1988.<br />

D. C. Mishra and M. Venkatarayudu. Magsat scalar anomaly map of India and part<br />

of Indian ocean - Magnetic crust and tectonic correlation. Geophys. Res. Lett., 12:<br />

781–784., 1985.


144<br />

J. B. Molyneux and D. R. Schmitt. First-break tim<strong>in</strong>g: Arrival onset times by direct<br />

correlation. Geophysics, 64:1492 – 1501., 1999.<br />

A. Müller. A new method to compensate for bias <strong>in</strong> magnetotellurics. Geophy. J. Int.,<br />

142:257–269., 2000.<br />

M. E. Murat and A. J. Rudman. Automated first arrival pick<strong>in</strong>g; a neural network<br />

approach. Geophys. Prosp., 40:587–604., 1992.<br />

N. Nagarajan. Application of robust estimation of transfer function for a magnetovariational<br />

array <strong>in</strong> Eastern India. <strong>in</strong> Deep Electromagnetic Exploration ed. KK. Roy et al,<br />

Narosa Publish<strong>in</strong>g House, India, page New Delhi, 1998.<br />

S. M. Naqvi and J. J. W. Rogers. Precambrian Geology of India. Oxford Monographs <strong>in</strong><br />

Geology and Geophysics No 6, Oxford University press, Oxford., page 223, 1987.<br />

N. Nityananda, A. K. Agarwal, and B. P. S<strong>in</strong>gh. Induction at short period on the<br />

horizontal field variation <strong>in</strong> the Indian Pen<strong>in</strong>sula. PEPI, 15:5–9., 1977.<br />

N. Nityananda and D. Jayakumar. Proposed relation between anomalous geomagnetic<br />

variations and tectonic history of South India. PEPI, 27:223–228., 1981.<br />

G. Oett<strong>in</strong>ger, V. Haak, , and J. C. Larsen. Noise reduction <strong>in</strong> magnetotelluric time-series<br />

with a new siganl-noise separation method and its application to a field experiment <strong>in</strong><br />

the Saxonian Granulite Massif. Geophys. J. Int., 146:659 – 669, 2001.<br />

A. L. Padilha, I. Vitorello, , and L. Rijo. Effects of the Equatorial Electrojet on magnetotelluric<br />

surveys: Field results from Northwest Brazil. Geophys. Res. Lett., 24(1):<br />

89–92., 1997.<br />

J. Park and A. D. Chave. On the estimation of magnetotelluric response functions <strong>us<strong>in</strong>g</strong><br />

the s<strong>in</strong>gular value decomposition. Geophys. Jour. of Roy. astr. Soc., 77:683–709., 1984.<br />

L. Pedersen. Some aspects of magnetotelluric field procedures. Surveys <strong>in</strong> Geophysics, 9:<br />

245–257., 1988.<br />

G. Petiau and A. Dupis. Noise, temperature coefficient and long time stability of electrodes<br />

for telluric observations. Geophys. Prosp., 28:792–804, 1980.<br />

M. Poulton, B. Stenberg, and C. Glass. Location of subsurface targets <strong>in</strong> geophysical<br />

<strong>data</strong> <strong>us<strong>in</strong>g</strong> neural networks. Geophysics, 57:1534 – 1544., 1992.<br />

E. Pulz and O. Ritter. Entwicklung e<strong>in</strong>er kalibriere<strong>in</strong>richtung fuer <strong>in</strong>duktionsspulenmagnetometer<br />

(search coils) am GeoForschungsZentrum Potsdam. Scientific Technical<br />

Report, GeoForschungsZentrum - Potsdam 10/01., 2001.<br />

B. Radhakrishna. Suspect tectono-stratigraphic terra<strong>in</strong> elements <strong>in</strong> the Indian subcont<strong>in</strong>ent.<br />

Jour. Geo. Soc.India, 34:1–24., 1989.


145<br />

S. S. Rai, D. Sr<strong>in</strong>agesh, and V. K. Gaur. Granulite evolution <strong>in</strong> south India - a seismic<br />

tomographic perspective. Mem. Geol.Soc. of India, 25:235 – 265., 1993.<br />

M. Rao. Study of geomagnetic pulsation characteristics and their solar cycle dependence<br />

<strong>in</strong> the equatorial region India? Ph.D. thesis, Osmania University., 2000.<br />

M. Rao, S. V. S. Sarma, N. Nagarajan, T. Har<strong>in</strong>arayana, C. Manoj, K. Naganjaneyulu,<br />

and V. T. C. Kumaraswamy. On the magnetotelluric source field effects <strong>in</strong> the Indian<br />

equatorial region - an experimental study. Presented at 16th EM <strong>in</strong>duction workshop<br />

held at Santa Fe, USA:12–22nd June, 2002.<br />

A. B. Reddi, M. P. Mathew, B. S<strong>in</strong>gh, and P. S. Naidu. Aeromagnetic evidence of crustal<br />

structure <strong>in</strong> the granulite terrane of Tamilnadu - Kerala. Jour. Geol. Soc. India, 32:<br />

368 – 381., 1988.<br />

O. A. Ritter, , and G. J. K. Dawes. New equipment and process<strong>in</strong>g for magnetotelluric<br />

remote refernce observations. Geophys. Jour. Int., 132:535–548., 1998.<br />

W. Rodi and R. L. Mackie. Nonl<strong>in</strong>ear conjugate gradients algorithm for 2-d magnetotelluric<br />

<strong>in</strong>versions. Geophysics, 66:174–187., 2001.<br />

Y. Sarma, T. Sastry, and S. Sarma. Equatorial effects on nighttime Pc3 puslation. Ind.<br />

J. Radio & Space. Phys., 11:29–32, 1982.<br />

T. S. Sastry, Y. S. Sarma, S. V. S. Sarma, and P. V. S. Narayan. Daytime Pi pulsations<br />

at equatorial latitudes. J. Atmos. Terr. Phys., 45:733–741, 1983.<br />

U. Schmucker. Auswerteverfahren gott<strong>in</strong>gen, <strong>in</strong> protokoll 7 kolloquium ’elekromagnetische<br />

tiefenforschung. Niedersachsisches Landesamt, Hannover., pages 163–188, 1978.<br />

Shalivahan and B. B. Bhattacharya. How remote can the far remote reference site for<br />

magnetotelluric measurements be? Jour. Geophys. Res, 107(B6):ETG 1–1, 2002.<br />

W. E. Sims, F. Bostick, and H. W. Smith. The estimation of the magnetotelluric<br />

impedance tensor elements from the measured <strong>data</strong>. Geophysics, 36:938–942., 1971.<br />

W. Siripunvaraporn and G. Egbert. An efficient <strong>data</strong>-subspace <strong>in</strong>version method for<br />

two-dimensional magnetotelluric <strong>data</strong>. Geophysics, 65(3):791–803., 2000.<br />

M. Y. Smirnov. <strong>Magnetotelluric</strong> <strong>data</strong> process<strong>in</strong>g with a robust statistical procedure<br />

hav<strong>in</strong>g a high breakdown po<strong>in</strong>t. Geophys. J. Int., 152:1–7., 2003.<br />

J. T. Smith and J. R. Booker. Rapid Inversion of Two - and Three-Dimensional <strong>Magnetotelluric</strong><br />

<strong>data</strong>. J. Geophys. Res., 96(B3):3905–3922., 1991.<br />

V. Spichak and I. Popova. Artificial neural network <strong>in</strong>version for magnetotelluric <strong>data</strong> <strong>in</strong><br />

terms of three-dimensional earth macroparameters. Geophys. J. Int., 142:15–26., 2000.<br />

J. A. Stodt. Noise <strong>analysis</strong> for conventional and remote reference magnetotellurics. PhD.<br />

dissertation, Univ of Utah, Logan, 1983.


146<br />

D. Sutarno and K. Vozoff. Phase-smoothed robust m-estimation of magnetotelluric<br />

impedance functions. Geophysics, 36:938–942., 1991.<br />

C. M. J. Swift. A magnetotelluric <strong>in</strong>vestigation of an electrical conductivity anomaly <strong>in</strong><br />

Southwestern US. <strong>in</strong> <strong>Magnetotelluric</strong> methods, Geophys repr<strong>in</strong>t series vol 5 ed. Vozoff,<br />

K., Soc. Explor. Geophys. Tulsa, Okla, pages 156–166, 1986.<br />

L. Szarka. Geophysical aspects of man-made magnetic noise <strong>in</strong> the earth - a review.<br />

Surveys <strong>in</strong> Geophysics, 9:287 – 318., 1988.<br />

N. K. Thakur, M. V. Mahashabde, B. R. Arora, B. P. S<strong>in</strong>gh, B. J. Srivastava, and S. N.<br />

Prasad. Geomagnetic variation anomalies <strong>in</strong> Pen<strong>in</strong>sular India. Geophys. Jour. Res.<br />

astr. Soc., 86:838–854., 1986.<br />

A. Tikhonov. On determ<strong>in</strong><strong>in</strong>g electrical characteristics of the deep layers of the Earth’s<br />

crust. Dokl. Akad. Nauk S.S.S.R., 73:295–297., 1950.<br />

J. M. Travassos and D. Beamish. <strong>Magnetotelluric</strong> <strong>data</strong> process<strong>in</strong>g: A case study. Geophys.<br />

J. Int., 93:377–391., 1988.<br />

K. Vozoff. The magnetotelluric method <strong>in</strong> electromagnetic methods <strong>in</strong> applied geophysics.<br />

ed by M.N. Nabighian, Society of Exploration Geophysics, Tulsa, USA., 1991.<br />

K. A. Vozoff. <strong>Magnetotelluric</strong> method <strong>in</strong> the exploration of sedimentary bas<strong>in</strong>s. Geophysics,<br />

37:98–141, 1972.<br />

J. R. Wait. On the relation between telluric currents and the earth’s magnetic field.<br />

Geophysics, 19:281289, 1954.<br />

S. H. Ward. The electromagnetic method. <strong>in</strong> SEG M<strong>in</strong><strong>in</strong>g geophysics volume II, M<strong>in</strong><strong>in</strong>g<br />

geophysics, Tulsa, pages 236 – 372., 1967.<br />

J. T. Weaver. Mathematical Methods for Geo-Electromagnetic Induction. Wiley, New<br />

York, 1994.<br />

P. Weidelt. The <strong>in</strong>verse problem of geomagnetic <strong>in</strong>duction. Zeitschrift für Geophysik, 38:<br />

257–289., 1972.<br />

P. J. Werbos. Backpropagation through time: What it does and how to do it. Proc.<br />

IEEE, 78:1550 – 1560., 1990.<br />

D. E. Wight and F. X. Bostick. Cascade decimation: A technique for real time estimation<br />

of power spectra. <strong>in</strong> Vozoff, K. (ed) <strong>Magnetotelluric</strong> methods: SEG Geophysics Repr<strong>in</strong>t<br />

Series 5, pages 215–218., 1980.<br />

J. W<strong>in</strong>er, R. J. A., R. J. R., and R. Moll. Predict<strong>in</strong>g carbonate permeabilities from<br />

wirel<strong>in</strong>e logs <strong>us<strong>in</strong>g</strong> back propagation networks. 61st SEG meet<strong>in</strong>g, Houston, USA.(<br />

Expanded abstract), pages 285–288., 1991.


147<br />

E. Yee, P. R. Kosteniuk, and K. V. Paulson. The reconstruction of the magnetotelluric<br />

impedance tensor: An adaptive parametric time-doma<strong>in</strong> approach. Geophysics, 53:<br />

1080–1087., 1988.<br />

Y. Zhang and K. V. Paulson. <strong>Magnetotelluric</strong> <strong>in</strong>version <strong>us<strong>in</strong>g</strong> regularized Hopfield neural<br />

networks. Geophys. Prosp., 45:725–743., 1997.<br />

Z. Zhang. Robust M estimators. Appeared at Internet URL http://wwwsop.<strong>in</strong>ria.fr/robotvis/personnel/zzhzng/Publis/Tutorial-Estim.,<br />

1996.<br />

Y. Zhao and K. Takano. An artificial neural network approach for broadband seismic<br />

pick<strong>in</strong>g. Bull. Seis. Soc. Am., 89:670–680, 1999.<br />

K. L. Zonge and L. J. Hughes. Controlled source audio frequency magnetotellurics. <strong>in</strong> EM<br />

methods <strong>in</strong> Applied Geophysics, ed M.N. Nabighian, Society of Exploration Geophysics.<br />

Tulsa, USA, 1991.


Errata<br />

The Follow<strong>in</strong>g changes/ additions have been <strong>in</strong>corporated <strong>in</strong>to the thesis based on the<br />

exam<strong>in</strong>ers comments.<br />

1. Case <strong>in</strong> which noise is difficult to be removed is given <strong>in</strong> §2.1, page 19. A description<br />

of noises that cause the scatter <strong>in</strong> the processed MT impedances is given <strong>in</strong> the<br />

second paragraph of §6.1, page 89. Error bars are def<strong>in</strong>ed <strong>in</strong> §3.3.4, Page 46.<br />

2. Two sentences on quantify<strong>in</strong>g the noise <strong>in</strong> MT <strong>data</strong> after process<strong>in</strong>g are <strong>in</strong>serted <strong>in</strong><br />

§5.8, page 86 (for ANN) and §6.2.4.6, page 101 (for robust process<strong>in</strong>g).<br />

3. Coherent and <strong>in</strong>coherent noises are def<strong>in</strong>ed <strong>in</strong> the first paragraph of §3.3.3, page<br />

45.<br />

4. Use of coherency to screen MT <strong>data</strong> <strong>in</strong> ”dead band” is given <strong>in</strong> the second paragraph<br />

of §3.3.3, page 45.<br />

5. Choices of <strong>data</strong> rejection gates other than coherency are discussed <strong>in</strong> the last paragraph<br />

of §3.3.3, page 45.<br />

6. Shortcom<strong>in</strong>gs of Kao and Rank<strong>in</strong> [1977]’s iterative MT process<strong>in</strong>g scheme is given<br />

<strong>in</strong> the last paragraph of §3.3.5, page 46.<br />

7. Specific improvements obta<strong>in</strong>ed by band averag<strong>in</strong>g are <strong>in</strong>serted <strong>in</strong> the second paragraph<br />

of §6.4.10, page 119.<br />

8. The thesis does not claim or deals with the averag<strong>in</strong>g of ”up” and ”down” biased<br />

MT impedances.<br />

9. MT process<strong>in</strong>g results along the Vellar-Palani profile <strong>in</strong> the SGT are given <strong>in</strong> §6.4.9,<br />

page 118.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!