Biomagnetics - Principles and Applications of Biomagnetic Stimulation and Imaging (PDFDrive)
Biomagnetics - Principles and Applications of Biomagnetic Stimulation and Imaging (PDFDrive)
Biomagnetics - Principles and Applications of Biomagnetic Stimulation and Imaging (PDFDrive)
Edited by
Shoogo Ueno
Professor Emeritus, The University of Tokyo
Masaki Sekino
Associate Professor, The University of Tokyo
This book contains information obtained from authentic and highly regarded sources. Reasonable
efforts have been made to publish reliable data and information, but the author and publisher cannot
assume responsibility for the validity of all materials or the consequences of their use. The authors and
publishers have attempted to trace the copyright holders of all material reproduced in this publication
and apologize to copyright holders if permission to publish in this form has not been obtained. If any
copyright material has not been acknowledged please write and let us know so we may rectify in any
future reprint.
Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced,
transmitted, or utilized in any form by any electronic, mechanical, or other means, now known or
hereafter invented, including photocopying, microfilming, and recording, or in any information stor-
age or retrieval system, without written permission from the publishers.
For permission to photocopy or use material electronically from this work, please access www.copy-
right.com (http://www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222
Rosewood Drive, Danvers, MA 01923, 978-750-8400. CCC is a not-for-profit organization that pro-
vides licenses and registration for a variety of users. For organizations that have been granted a photo-
copy license by the CCC, a separate system of payment has been arranged.
Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are
used only for identification and explanation without intent to infringe.
Visit the Taylor & Francis Web site at
http://www.taylorandfrancis.com
and the CRC Press Web site at
http://www.crcpress.com
Contents
Foreword, vii
Preface, ix
Acknowledgments, xv
Editors, xvii
Contributors, xix
Chapter 1 ◾ Introduction 1
Shoogo Ueno
v
vi ◾ Contents
INDEX, 321
Foreword
vii
viii ◾ Foreword
ix
x ◾ Preface
The core parts of the book are based on lecture notes that the editors
of the book have used for students in graduate courses at the Department
of Electronic Engineering, Graduate School of Engineering, and the
Department of Biomedical Engineering, Graduate School of Medicine,
University of Tokyo for more than 20 years. Dr. Ueno has been con-
tinuously revising and improving his notes since 1994 when he started
his Laboratory of Biomagnetics and Bioimaging at the Department of
Biomedical Engineering, Faculty of Medicine, at the University of Tokyo,
Tokyo, Japan. Dr. Ueno retired from the University of Tokyo as professor
emeritus. Masaki Sekino is now responsible for a course of biomedical
engineering at the Graduate School of Engineering, University of Tokyo.
In this book, the editors asked five colleagues to join chapter contribu-
tors to further explore and study the field of biomagnetics and bioimaging.
The book is composed of 10 chapters, and a brief outline of each follows.
Shoogo Ueno
Masaki Sekino
xv
Editors
Shoogo Ueno, PhD, earned his BS, MS, and PhD (Dr Eng) degrees in
electronic engineering from Kyushu University in 1966, 1968, and 1972,
respectively. Dr. Ueno was an associate professor with the Department
of Electronics, Kyushu University, from 1976 to 1986. From 1979 to 1981,
he spent his sabbatical with the Department of Biomedical Engineering,
Linkoping University, Linkoping, Sweden, as a guest scientist. He served
as a professor in the Department of Electronics, Graduate School of
Engineering, Kyushu University from 1986 to 1994. He subsequently
served as a professor in the Department of Biomedical Engineering,
Graduate School of Medicine, University of Tokyo from 1994 to 2006.
During this time, he also served as a professor in the Department of
Electronic Engineering, Graduate School of Engineering, University of
Tokyo. In 2006 he retired from the University of Tokyo as professor emeri-
tus. Since then, he has been a professor with the Department of Applied
Quantum Physics, Graduate School of Engineering, Kyushu University,
Fukuoka, Japan, and he served as the dean of the faculty of medical tech-
nology, Teikyo University, Fukuoka, Japan from 2009 through 2012.
Dr. Ueno is a fellow of the Institute of Electrical and Electronics Engineers,
IEEE (2001), and a life fellow (2011) and a fellow of the American Institute
for Medical and Biological Engineering, AIMBE (2001). He is a fellow
(2006) and secretary (2012) of the Governing Council of the International
Academy for Medical and Biological Engineering, IAMBE. He was a pres-
ident of the Bioelectromagnetics Society, BEMS (2003–2004), chairman of
the Commission K on Electromagnetics in Biology and Medicine of the
International Union of Radio Science (URSI) (2000–2003). He received
the Doctor Honoris Causa from Linkoping University, Linkoping, Sweden
(1998). He was a 150th Anniversary Jubilee Visiting Professor at Chalmers
University of Technology, Gothenburg, Sweden (2006) and a visiting pro-
fessor at Simon Frasier University, Burnaby, Canada (1994) and Swinburne
xvii
xviii ◾ Editors
Masaki Sekino, PhD, earned his BS, MS, and PhD (Dr Eng) degrees in
mechanical engineering and electronic engineering from the University
of Tokyo, Tokyo, Japan, in 2000, 2002, and 2005, respectively. Dr. Sekino
was a research associate with the Department of Biomedical Engineering,
Graduate School of Medicine, University of Tokyo, from 2005 to 2006. He
was subsequently a research associate (2006–2010), a lecturer (2010–2011),
and an associate professor (2011–2015) with the Department of Electrical
Engineering and Information Systems, Graduate School of Engineering,
University of Tokyo, Tokyo, Japan.
Dr. Sekino was awarded the International Union of Radio Science
Young Scientist Award (2002), and the International Conference on
Complex Medical Engineering Best Conference Paper Award (2012).
Contributors
xix
Chapter 1
Introduction
Shoogo Ueno
CONTENTS
1.1 History of Biomagnetics 1
1.1.1 From Gilbert (1600) to d’Arsonval (1896) 1
1.1.2 Modern Biomagnetics in the 1970s and 1980s 4
1.2 Biomagnetic Phenomena and Medical Applications 7
1.2.1 Biomagnetic Phenomena at Different Intensities
and Their Frequencies 7
1.2.2 Biomagnetic Measurements of Magnetic Fields
Produced by Living Systems 8
1.2.3 Biomagnetic Stimulation and Magnetic Treatments 9
1.2.4 Magnetic Resonance Imaging of the Living System 10
1.2.5 Magnetic Control of Cell Manipulation, Cell
Orientation, and Cell Growth 11
1.2.6 Biological Interactions of ELF, Pulsed, and RF
Magnetic Fields 12
1.3 Fundamental Bases for Understanding Biomagnetic Phenomena 12
1.3.1 Electromagnetic Induction in Living Tissues 12
1.3.2 Magnetism of Living Systems and Materials 14
1.4 Summary 16
References 16
1.1 HISTORY OF BIOMAGNETICS
1.1.1 From Gilbert (1600) to d’Arsonval (1896)
Biomagnetics is an interdisciplinary field where magnetics, biology, and
medicine overlap. Biomagnetics is also called biomagnetism, which is
often used in medicine and biology. In this book, however, biomagnetics
1
2 ◾ Biomagnetics
The original book was written in Latin, and the above quote is taken
from Busby.2
Gilbert’s hypothesis is very thought-provoking and suggests the study
of interactions between magnetism and human life. It was difficult, how-
ever, to verify Gilbert’s profound hypothesis in 1600 because science and
technology were not well developed at that time. Gilbert was a physicist as
well as a physician for Queen Elizabeth I (1533–1603) in England.
In France in 1896, Jacques-Arsene d’Arsonval (1851–1940) reported a
phenomenon called magnetophosphene, a visual sensation induced in
human subjects when exposed to alternating magnetic fields.3 The magneto
phosphene reported by d’Arsonval was among the first reliable phenomena
related to the interactions of human life with magnetism (Figure 1.2).
The magnetophosphenes were repeated by other groups; for example,
S. P. Thompson in the United Kingdom observed that peripheral areas
in the eye see light more clearly than in the center of the eye.4 C. E.
Magnusson and H. C. Stevens in the United States observed frequency
characteristics in light sensation, changing the frequency of alternating
magnetic fields, and determined that the light is sensitive around 20 Hz.5
P. Lovsund, P. A. Oberg, and S. E. G. Nilsson in Sweden quantitatively
that a chemical yield of the cage or escape product shows a magnetic field
dependence when a singlet–triplet intersystem crossing is subject to mag-
netic perturbations.41,42 This area has been well developed as dynamic spin
chemistry in the 1980s and 1990s.41,42 Possible biomagnetic and chemi-
cal effects can be expected when biological systems are exposed to both
magnetic fields and other energy such as light and radiation.43 However,
biological effects of magnetic fields based on dynamic spin chemistry with
a radical-pair model have not yet been clarified.44
Basic studies on biomagnetics and the related fields in the 1970s and
1980s have been explored for further developments in the 1990s and 2000s.
1.2 BIOMAGNETIC PHENOMENA
AND MEDICAL APPLICATIONS
1.2.1 Biomagnetic Phenomena at Different
Intensities and Their Frequencies
Various biomagnetic phenomena for different intensities and their fre-
quencies are shown in Figure 1.3.45,46
Magnetic stimulation of the brain, (i.e., transcranial magnetic stimula-
tion [TMS] of the human brain), requires strong pulsed magnetic fields
106
of 1–2 T for a short period of time (0.1–0.2 msec) to get sufficient induced
electric fields for the stimulation of neurons in the brain. Magnetic stimu-
lation of the heart requires more strong pulsed magnetic fields (1–5 T) with
a 1–2 msec pulse width to excite the cardiac muscles. In Figure 1.3, the
frequencies of the pulsed magnetic fields are converted from pulse width.
Rapid changes in blood flow in the hands of human subjects are
observed when the hand is exposed to alternating magnetic fields of 32 mT
at 3.8 kHz.47 The explanation for the blood flow changes following mag-
netic field exposure is that nerve system information from the skin sensory
receptors is sent via the spinal cord to the central nervous system, which
modulates the efferent nerve activity and affects the blood flow of the skin.
The magnetophosphene shows a V-shaped curve for light sensation
with a minimum threshold of 10 mT at 20 Hz.6
Effects of magnetic fields on biological systems are observed mostly in the
range of magnetic fields higher than the Earth’s magnetic field of 30–50 μT.
Health effects of extremely low frequency (ELF) electromagnetic fields
are discussed, introducing possible mechanisms including a model of
calcium release from living cells exposed to ELF fields. Amplitudes and
frequencies of ELF electromagnetic fields related to consumer electronics
and power lines are in the oval area marked in Figure 1.1.
Magnetic fields used in clinical MRI systems today are 0.3 T for per-
manent magnets and 1.5–3.0 T for superconducting magnet systems. For
research purposes, 4–7 T MRI systems are used. Magnetic orientation
of biological materials is observed in 4–8 T magnetic fields. “Parting of
water” or the “Moses effect” is observed in 8-T magnetic fields.
In contrast, biomagnetic fields or magnetic fields produced from the
living systems are extremely weak, 10–15 to 10–11 T or 1 fT to 10 pT order of
magnetic fields, buried with urban noise. These biomagnetic fields are usu-
ally measured by superconducting quantum interference device (SQUID)
systems in a magnetically shielded room.
brain stem magnetic fields, are extremely weak, 5 × 10–15 T or 5 f T magnetic
fields, and these magnetic fields are detected by SQUID systems by averaging
several thousand times above signal-to-noise ratio. Spontaneous brain mag-
netic activities such as alpha MEG activities of 10–12 T or pT order of magnetic
fields are detected by SQUID systems without averaging the signals.
Magnetocardiography (MCG) is a noninvasive modality for detecting
electrical activities of the heart, and the biomagnetic signals are 10–12 to
10–10 T or 1–100 pT order of magnetic fields.
Magnetopneumography (MPG) is a technique to measure magnetic
fields from the lung.48,49 The MPG measures remnant magnetic fields
produced by magnetic contaminants in the lung or deposition of inhaled
magnetic nanoparticle dusts in the lung. The lung is magnetized by exter-
nal magnetic fields and a relaxation process from a peak of the magnetized
vector of MPG signals to a low level of signals due to randomization of
the magnetized vector. The relaxation process is reflected by the ability of
phagocytosis of alveolar macrophages.
Origins or sources of magnetic fields of MEG and MCG are electrical
currents generated by electrical activities of neurons and muscles, whereas
the magnetic fields of MPG come from magnetic dust in and around alve-
olar macrophages in the lungs.
In biomagnetic measurements, spatiotemporal distributions of magnetic
fields over the surface of the body are obtained by multichannel SQUID
systems, and the sources of biomagnetic fields are estimated by solving
inverse problems. From a mathematical point of view, there are no unique
solutions in inverse problems, and localizations of sources are estimated by
adding some assumptions and boundary or constraint conditions.
static magnetic fields, and (3) effects of multiplications of both static mag-
netic fields and other energy such as light and radiation.
When we discuss the effects of time-varying magnetic fields, the rela-
tionship between magnetism and electricity is important. Hans C. Oersted
(1777–1851) discovered a phenomenon that electric currents produce mag-
netic fields in 1820 in Denmark. Michael Faraday (1791–1867) discovered
a phenomenon called electromagnetic induction in 1831 in England.
Electric currents are induced in the secondary coil when the currents in
the primary coil change rapidly. In other words, electric fields are induced
by time-varying magnetic fields. James C. Maxwell (1831–1879) derived
the fundamental four equations, called Maxwell’s equations, to describe
the relationship between magnetism and electricity mathematically in
1864. That is, electromagnetic phenomena consist of electric and magnetic
fields that change with time and space. Maxwell’s equations are laws that
define the relationship between temporally and spatially averaged electric
and magnetic fields.
Because the living body is composed of conductive tissues and mate-
rials, electric fields are induced in it when it is exposed to time-varying
magnetic fields. In the case of low-frequency magnetic fields lower than
several 100 kHz or pulsed magnetic fields, stimulatory effects of excitable
membranes such as nervous system and muscular system are dominant.
Transcranial magnetic brain stimulation is realized by this principle.
In contrast, in the case of high-frequency magnetic and electromagnetic
fields higher than several 100 kHz and in the range of RF fields, thermal
effects of living systems are dominant. Health effects of RF electromag-
netic fields used in mobile telephony and MRI systems are discussed
mainly on the basis of the thermal effects. The specific absorption rate
(SAR) (W/kg) is used for the evaluation of biological effects of RF fields.
When an electric field E (V/m) is induced in a tissue exposed to RF elec-
tromagnetic fields, SAR (W/kg) in the tissue is determined by
SAR = σE2/ρm
1.4 SUMMARY
In this biomagnetics overview, we study biomagnetics, an interdisciplin-
ary field. We focus on basic principles and medical applications of bio-
magnetic stimulation and imaging in the following chapters. We hope the
readers enjoy the essence of recent advances in biomagnetics in this book.
We recommend that readers refer to other bibliographies related to bio-
magnetics, bioelectromagnetics, brain science, neuroimaging, and regen-
erative medicine, as well as basic sciences to expand their exposure to and
knowledge of the world of biomagnetics. We provide some other review
articles, book chapters, and books.79–89
REFERENCES
1. Gilbert, W. 1600. De Magnete. Translation to English by Mottelay (1958).
Dover, New York.
2. Busby, D. E. 1968. Space biomagnetics. Space Life Science 1: 23–63.
3. D’Arsonval, J. A. 1896. Disrositifs pour la mesure des courants alterna-
tives des toutes frequencies. Comptes Rendus l’Académie des Sciences 48:
450–451.
4. Thompson, S. P. 1910. A physiological effect of an alternating magnetic
field. Proceedings of the Royal Society, Series B 82: 396–398.
5. Magnusson, C. E., and Stevens, H. C. 1911. Visual sensations caused by
changes in the strength of a magnetic field. American Journal of Physiology
29: 124–136.
Introduction ◾ 17
22. Ueno, S., Lovsund, P., and Oberg, P. A. 1986. Effect of time-varying mag-
netic fields on the action potential in lobster giant axon. Medical and
Biological Engineering and Computing 24: 521–526.
23. Ueno, S., Harada, K., Ji, C., and Oomura, Y. 1984. Magnetic nerve stim-
ulation without interlinkage between nerve and magnetic flux. IEEE
Transactions on Magnetics MAG-20: 1660–1662.
24. Barker, A. T., Jalinous, R., and Freeston, I. L. 1985. Non-invasive magnetic
stimulation of human motor cortex. The Lancet i: 1106–1107.
25. Ueno, S, Tashiro, T., and Harada, K. 1988. Localized stimulation of neu-
ral tissues in the brain by means of a paired configuration of time-varying
magnetic fields. Journal of Applied Physics 64: 5862–5864.
26. Ueno, S., Matsuda, T., Fujiki, M., and Hori, S. 1989. Localized stimulation
of the human motor cortex by means of pair of opposing magnetic fields.
Digest of International Magnetics Conference, Washington, D.C.: GD-10.
27. Ueno, S., Matsuda, T., and Fujiki, M. 1989. Localized stimulation of the
human cortex by opposing magnetic fields. In Advances in Biomagnetism,
Williamson, S. J., Hoke, M., Stroink, G., and Kotani, M., eds. Springer, New
York: 529–532.
28. Ueno, S., Matsuda, T., and Fujiki, M. 1989. Localized stimulation of the
human brain by a pair of opposing pulsed magnetic fields. Memoirs of the
Faculty of Engineering, Kyushu University 49: 161–173.
29. Ueno, S., Matsuda, T., and Fujiki, M. 1990. Functional mapping of the
human motor cortex obtained by focal and vectorial magnetic stimulation
of the brain. IEEE Transactions on Magnetics 26: 1539–1544.
30. Ueno, S. (Ed.) 1994. Biomagnetic Stimulation. Plenum Press, New York:
1–136.
31. Blakemore, R. P. 1975. Magnetotactic bacteria. Science 190: 377–379.
32. Kalmijn, A. J., and Blakemore, R. P. 1977. Geomagnetic orientation in
marine mud bacteria. Proceedings of International Union of Physiological
Society 13: 364.
33. Frankel, R. B., Blakemore, R. P., and Wolf, R. S. 1979. Magnetite in freshwa-
ter magnetotactic bacteria. Science 203: 1355–1356.
34. Kirshvink, J. L., Jones, D. S., and MacFadden, B. J. (Eds.) 1985. Magnetite
Biomineralization and Magnetoreception in Organisms: A New Biomag
netism. Plenum Press, New York: 1–682.
35. Kirshvink, J. L., Kobayashi-Kirshvink, A., and Woodford, B. J. 1992.
Magnetite biomineralization in the human brain. Proc. Natl. Acad. Sci.,
USA (Biophysica) 89: 7683–7687.
36. Matsunaga, T., Nakamura, C., Burgess, J. G., and Sode, K. 1992. Gene trans-
fer in magnetic bacteria; transposon mutagenesis and cloning of genomic
DNA fragments for magnetosome synthesis. Journal of Bacteriology 174:
2748–2753.
37. Torbet, J, Freyssinet, M., and Hudy-Clergeon, G. 1981. Oriented fibrin gels
formed by polymerization in strong magnetic fields. Nature 289: 91–93.
38. Hata, N. 1976. The effect of external magnetic field on the photochemical
reaction of isoquinoline N-oxide. Chemical Letters 5: 547–550.
Introduction ◾ 19
39. Shulten, K., Staek, H., Welter, A., Wernar, H. J., and Nickel, B. 1976.
Magnetic field dependence of the geminate recombination of radical ion
pairs in polar solvents. Zeitschrift fűr Physikalische Chemie 101: 371–390.
40. Tanimoto, Y., Hayashi, H., Nagakura, S., Sakaguchi, H., and Tokumaru, K.
1976. The external magnetic field effect on the singlet sensitized photolysis
of dibenzoyl peroxide. Chemical and Physical Letters 41: 267–269.
41. Nagakura, S., and Molin, Y. (Eds.) 1992. Magnetic field effects upon photo
physical and photochemical phenomena. Chemical Physics, Special Issue
162: 1–234.
42. Nagakura, S., Hayashi, H., and Azumi, T. (Eds.) 1998. Dynamic Spin
Chemistry: Magnetic Controls and Spin Dynamics of Chemical Reactions.
John Wiley, New York: 1–297.
43. Ueno, S., and Harada, K. 1986. Experimental difficulties in observing
the effects of magnetic fields on biological and chemical processes. IEEE
Transactions on Magnetics MAG-22: 868–873.
44. Ueno, S. (Ed.) 1996. Biological Effects of Magnetic and Electromagnetic
Fields. Plenum Press, New York: 1–243.
45. Ueno, S., and Iwasaka, M. 1996. Magnetic nerve stimulation and effects of
magnetic fields on biological, physical and chemical processes. In Biological
Effects of Magnetic and Electromagnetic Fields, Ueno, S., ed. Plenum Press,
New York: 1–21.
46. Ueno, S., and Shigemitsu, T. 2006. Biological effects of static magnetic
fields. In Handbook of Biological Effects of Electromagnetic Fields, 3rd
ed. Bioengineering and Biophysical Aspects of Electromagnetic Fields,
Barnes, F. S., and Greenebaum, B., ed. CRC Press, Boca Raton, Fla.:
203–259.
47. Ueno, S., Lovsund, P, and Oberg, P. A. 1986. Effects of alternating magnetic
fields and low-frequency electrical currents on human skin blood flow.
Medical and Biological Engineering and Computing 24: 57–61.
48. Cohen, D. 1973. Ferromagnetic contamination in the lungs and other
organs of the human body. Science 180: 745–748.
49. Kalliomaki, P-L., Kalliomaki, K, Aittoniemi, K., Korhonen, O, Pasanen, J.,
and Moilanen, M. 1983. In Biomagnetism: An Interdisciplinary Approach,
Williamson, S. J., Romani, G-L., Kaufman, L., and Modena, I., eds. Plenum
Press, New York: 533–568.
50. Bourland, J. D., Mouchawar, G. A., Nyenhuis, J. A., Geddes, L., A., Foster,
K. S., Jones, J. T., and Graber, G. P. 1990. Transchest magnetic (eddy-
current) stimulation of the dog heart. Medical and Biological Engineering
and Computing 28: 196–198.
51. Bourland, J. D., Mounchawar, G. A., Nyenhuis, J. A., Geddes, L. A., Foster,
K. S., Jones, J. T., and Graber, G. P. 1992. Closed-chest cardiac stimula-
tion with a pulsed magnetic field. Medical and Biological Engineering and
Computing 30: 162–168.
52. Georg, M. S., and Wassermann, E. M. 1994. Rapid-rate transcranial mag-
netic stimulation and ECT. Convulsive Therapy 10: 251–254.
20 ◾ Biomagnetics
53. Chandos, B., Khan, A., Lai, H., and Lin, J. C. 1996. The application of elec-
tromagnetic energy to the treatment of neurological and psychiatric dis-
eases. In Biological Effects of Magnetic and Electromagnetic Fields, Ueno, S.,
ed. Plenum Press, New York: 161–169.
54. Fujiki, M., and Stewart, O. 1997. High frequency transcranial magnetic
stimulation for protection against delayed neuronal death induced by tran-
sient ischemia. Journal of Neurosurgery 99: 1063–1069.
55. Muller, M. B., Toschi, N., Kresse, A. E., Post, A., and Keck, M. E. 2000.
Long-term repetitive transcranial magnetic stimulation increases the
expression of brain-derived neurotrophic factor and cholecystokinin
mRNA, but not neuropeptide tyrosine mRNA in specific areas of rat brain.
Neuropsychopharmacology 23: 205–215.
56. Ogiue-Ikeda, M., Kawato, S., and Ueno, S. 2003. The effect of repetitive
transcranial magnetic stimulation on long-term potentiation in rat hippo-
campus depends on stimulus intensity. Brain Research 993: 222–226.
57. Ogiue-Ikeda, M., Kawato, S, and Ueno, S. 2005. Acquisition of ischemic
tolerance by repetitive transcranial magnetic stimulation in the rat hippo-
campus. Brain Research 1037: 7–11.
58. Ueno, S., and Fujiki, M. 2007. Magnetic stimulation. In Magnetism in
Medicine A Handbook Second Completely Revised and Enlarged Edition,
Andra, W., and Nowak H., ed. Wiley-VHC, Weinheim: 511–528.
59. Wassermann, E. M. 1998. Risk and safety of repetitive transcranial mag-
netic stimulation: Report and suggested guidelines from the International
Workshop on the Safety of Repetitive Transcranial Magnetic Stimulation,
June 5-7, 1996. Electroenchephalography and Clinical Neurophysiology 108:
1–16.
60. Pascual-Leone, A., Davey, N., Rothwell, J., Wassermann, E. M., and Puri,
B. K. 2002. Handbook of Transcranial Magnetic Stimulation. Hodder
Arnold, London.
61. Lauterbur, P. C. 1973. Image formation by induced local interactions:
Examples employing nuclear magnetic resonance. Nature 242: 190–191.
62. Mansfield, P. 1977. Multi-planar image formation using NMR spin echoes.
Journal of Physics C: Solid State Physics 10: L55–L58.
63. Ogawa, S., Lee, T. M., Kay, A. R., and Tank, D. W. 1990. Brain magnetic reso-
nance imaging with contrast dependent on blood oxygenation. Proceedings
of the National Academy of Sciences USA 87: 9868–9872.
64. Le Bihan, D., Turner, R., Douek, P., and Patronas, N. 1992. Diffusion MR
imaging: Clinical applications. American Journal of Roentgenology 159:
591–599.
65. Basser, P. J., Mattiello, J., and Le Bihan, D. 1994. Estimation of the effec-
tive self-diffusion tensor from the NMR spin echo. Journal of Magnetic
Resonance B 103: 247–254.
66. Ueno, S., and Iriguchi, N. 1998. Impedance magnetic resonance imaging.
Journal of Applied Physics 83: 6450–6452.
Introduction ◾ 21
67. Kamei, H., Iramina, K., Yoshikawa, K., and Ueno, S. 1999. Neuronal cur-
rent distribution imaging using magnetic resonance. IEEE Transactions on
Magnetics 35: 4109–4111.
68. Ueno, S. 1999. Biomagnetic approaches to studying the brain. IEEE
Engineering in Medicine and Biology 18: 108–120.
69. Sekino, M., Ohsaki, H., Yamaguchi-Sekino, S., and Ueno, S. 2009. Toward
detection of transient changes in magnetic-resonance signal intensity aris-
ing from neuronal electrical activities. IEEE Transactions on Magnetics 45:
4841–4844.
70. Ueno, S., and Iwasaka, M. 1994. Properties of diamagnetic fluid in high
gradient magnetic fields. Journal of Applied Physics 75: 7177–7179.
71. Ueno, S., and Iwasaka, M. 1994. Parting of water by magnetic fields. IEEE
Transactions on Magnetics 30: 4698–4700.
72. Kotani, H., Kawaguchi, H., Shimoaka, T., Iwasaka, M., Ueno, S., Ozawa, H.,
Nakamura, K., and Hoshi, K. 2002. Strong static magnetic field stimulates
bone formation to a definite orientation in vivo and in vitro. Journal of Bone
and Mineral Research 17: 1814–1821.
73. Ueno, S., Ando, J., Fujita, H., Sugawara, T., Jimbo, Y., Itaka, K., Kataoka, K.,
and Ushida, T. 2006. The state of the art of nanobioscience in Japan. IEEE
Transactions on Nanobioscience 5: 54–65.
74. Ueno, S., and Sekino, M. 2006. Biomagnetics and bioimaging for medical
applications. Journal of Magnetism and Magnetic Materials 304: 122–127.
75. Ogiue-Ikeda, M., Sato, Y., and Ueno, S. 2003. A new method to destruct
targeted cells using magnetizable beads and pulsed magnetic force. IEEE
Transactions on Nanobioscience 2: 262–265.
76. Yamaguchi, S., Ogiue-Ikeda, M., Sekino, M., and Ueno, S. 2006. The effect
of repetitive magnetic stimulation on tumor and immune functions in
mice. Bioelectromagnetics 27: 64–72.
77. Ueno, S. 1989. Quenching of flames by magnetic fields. Journal of Applied
Physics 65: 1243–1245.
78. Cespedes, O., and Ueno, S. 2009. Effects of radio frequency magnetic fields
on iron release from cage proteins. Bioelectromagnetics 30: 336–342.
79. Ueno, S. 2012. Studies on magnetism and bioelectromagnetics for 45 years:
From magnetic analog memory to human brain stimulation and imaging.
Bioelectromagnetics 33: 3–22.
80. Ueno, S. 2008. New horizon in biomagnetics and bioimaging. Electro
magnetic Field, Health and Environment. Proceedings of EHE’07, Krawczyk, A.,
Kubacki, R., Wiak, S., and Antunes, C. L., eds. IOS Press, Amsterdam: 8–17.
81. Andra, W., and Nowak, H. (Eds.). 1998. Magnetism in Medicine A Handbook,
Wiley-VCH, Berlin.
82. Andra, W., and Nowak, H. (Eds.). 2007. Magnetism in Medicine A Handbook,
2nd ed. Wiley-VCH, Weinheim.
83. Barnes, F. S., and Greenebaum, B. (Eds.). 2007. Handbook of Biological
Effects of Electromagnetic Fields, 3rd ed., Bioengineering and Biophysical
Aspects of Electromagnetic Fields (one part) and Biological and Medical
Aspects of Electromagnetic Fields (one part). CRC Press, Boca Raton, Fla.
22 ◾ Biomagnetics
Principles of
Biomagnetic Stimulation
Shoogo Ueno and Masaki Sekino
CONTENTS
2.1 Introduction 24
2.1.1 Basic Principles 24
2.1.2 Advantages of Magnetic Stimulation 26
2.2 Magnetic Stimulator 28
2.2.1 Stimulator Coils 28
2.2.1.1 Circular Coil 28
2.2.1.2 Figure-Eight Coil 30
2.2.1.3 Coil Array 31
2.2.1.4 Emerging Coil Designs 32
2.2.1.5 Cooling Systems 34
2.2.1.6 Measurement of Magnetic Field and Coil Current 35
2.2.2 Driving Circuit 36
2.2.2.1 Hardware 36
2.2.2.2 Analysis of Driving Circuit 38
2.2.3 Coil Navigation System 40
2.2.4 Multimodal Imaging Using Magnetic Stimulation 41
2.3 Mechanisms of Magnetic Stimulation 43
2.3.1 Fundamentals of Electrophysiology 43
2.3.1.1 Resting Membrane Potential 43
2.3.1.2 Action Potential 44
2.3.2 Transcranial Magnetic Stimulation 46
2.3.2.1 Resting Motor Threshold 46
2.3.2.2 Stimulation Site 46
2.3.3 Peripheral Nerve Stimulation 47
23
24 ◾ Biomagnetics
2.1 INTRODUCTION
2.1.1 Basic Principles
Electromagnetic induction, which provides the fundamental principle
of biomagnetic stimulation, was discovered by Michael Faraday in 1831.
When a pair of coils is situated close to one another and the current
flowing in one coil varies, an electromotive force arises in the other coil.
The current flowing in the primary coil generates a magnetic field, and a
rapid switching of the current causes a steep variation in the magnetic flux
linkage. The magnitude of electromotive force in the circuit is proportional
to the time rate of change in this flux linkage. The electromotive force is
directed so that the magnetic field arising from the secondary coil current
inhibits the variation of flux linkage. Because biological tissues and fluids
are electrically conductive, the electromagnetic induction occurs also in
the body to induce electric fields and resulting currents. When a rapidly
varying current is applied to a coil located on the surface of the body,
electric fields are induced in nearby tissues owing to the varying magnetic
fields. The magnetic stimulation is understood as an electric stimulation
mediated by magnetic fields because the resulting biological effects
originate from the induced electric fields. Considering that the magnetic
permeabilities of biological tissues are nearly equal to that of free space,
electric properties of biological tissues play important roles in biological
effects of alternating and pulsed magnetic fields. At low frequencies below
Principles of Biomagnetic Stimulation ◾ 25
Coil current
Coil current
FIGURE 2.1 (a) Transcranial magnetic stimulation. (b) Electric fields induced by
a circular coil. (c) Electric fields induced by a figure-eight coil.
In addition to the brain, the spinal root, diaphragm, and heart can also
be magnetically stimulated. Applications of magnetic stimulations to arti-
ficial respiration and defibrillation have been explored.
3000
2000
1000
0
−500 0 500 1000 1500
Time (μsec)
(a) −1000
Coil current (A)
4000
2000
0
−200 0 200 400 600
Time (μsec)
−2000
(b) −4000
FIGURE 2.2 (a) Monophasic and (b) biphasic waveforms of current flowing in
the coil. Magnetic field intensity is proportional to the current.
the electric field by pulsed magnetic field, one can find a difference in the
resulting electric field distributions. In the case of using electrodes, the
streamline of electric field flows out from the positive electrode, diffuses
in the body, and flows into the negative electrode. In the case of TMS, the
induced fields flow along closed-loop pathways in the body, as shown in
Figure 2.1b and c.
Recent clinical studies have shown that TMS improves various symp-
toms of neurological and psychiatric diseases. It should be noted that the
TMS exhibits therapeutic effects even for a percentage of drug-resistant
patients. TMS is advantageous also in terms of rarely causing side effects,
which are sometimes problematic in drug treatment. As discussed pri-
marily in Chapter 3, TMS is a promising technique for a variety of fields
such as brain research, diagnosis of the nervous system, treatment of dis-
eases, and recovery of brain function.
28 ◾ Biomagnetics
2.2 MAGNETIC STIMULATOR
2.2.1 Stimulator Coils
2.2.1.1 Circular Coil
Circular coils have been used since TMS was invented.6 Owing to the simplest
structure among various types of coils, the circular coils are still widely used
today. As shown in Figure 2.1b, the coil conductors are wound up circularly.
Figure 2.3a shows a circular coil for basic studies. The generated magnetic
fields are perpendicular to the surface of the head below the coil center and are
tangential to the surface below the coil conductor. The electric fields generated
in the brain form circular pathways along the surface of the head and in the
direction opposite the coil current. The magnitude of electric field is zero
below the center of the coil and is highest below the coil conductors. As a
result, the electric fields distribute in an area whose dimension is comparable
to the coil diameter. Because the magnitude of magnetic field decreases with
the distance from the coil, the electric fields are highest at the surface of the
head and attenuate at deep regions in the brain. Coils with smaller diameters
(a)
(b)
(c)
FIGURE 2.3 (a) Circular coil, (b) figure-eight coil, and (c) double-cone coil.
These coils are widely used for magnetic stimulation.
Principles of Biomagnetic Stimulation ◾ 29
1 ∂
Br = (Aϕ sinθ) (2.1)
r sinθ ∂θ
1 ∂
Bθ = − (rAϕ ) (2.2)
r ∂r
Bφ = 0 (2.3)
2π
µ Ia cos ϕ ′dϕ ′
Aϕ = 0
4π ∫
0
(a + r − 2ar sinθcos ϕ ′)1/2
2 2
µ0 4Ia ⎡ (2 − k 2 )K(k)− 2E(k) ⎤ (2.4)
=
4π a 2 + r 2 + 2ar sinθ ⎢⎣ k2 ⎥⎦
4ar sinθ
k2 =
a + r 2 + 2ar sinθ (2.5)
2
where K(k) and E(k) are elliptic integrals of the first and second kinds,
respectively.
π/2
∫
1
K(k) = dθ (2.6)
0
1− k 2 sin 2 θ
π/2
E(k) =
∫
0
1− k 2 sin 2 θ dθ
(2.7)
30 ◾ Biomagnetics
2.2.1.2 Figure-Eight Coil
The figure-eight coil was originally developed by Ueno et al.7,8 This coil
is now in widespread clinical use along with the single circular coil. The
figure-eight coil consists of a pair of circular coils adjacent to one another,
and the current is applied in the opposite directions to these circular coils.
Magnetic fields are produced from the two circular coils in the opposite
directions, and the magnetic fields induce electric fields in the brain. As
shown in Figure 2.1c, the electric field flows in the brain in the reversed
direction, forming two loop pathways. The electric fields converge in the
target region in the brain below the intersection of the figure-eight coil,
resulting in an electric field intensity approximately 3× higher than those
around the target.11 In the single circular coil, however, the electric fields
form a circular pathway. The resulting fields get the maximum intensity
below the coil conductors, and the intensity becomes zero below the center
of the circle.
The figure-eight coil enables one to stimulate a target area in the human
cerebral cortex with a spatial resolution as high as 5 mm.7 Thus, this coil
is extremely effective when stimuli should be delivered to a specific target
area. Another advantage of this coil is the ability to provide a vectorial
stimulation because the electric fields in the target area are induced in a
specific direction along the surface of the head. The attitude of the coil
defines the direction of electric field.
Figure 2.3b and c show two major figure-eight coils: planar coil and
double-cone coil. In the double-cone coil, the outer conductors come close
to the surface of the head, which intensify the electric field at the target
area. The electric fields in the brain generally attenuate with the distance
from the coil conductors. An increase in the diameter of the coil leads to
stronger electric fields in deep parts of the brain.
Principles of Biomagnetic Stimulation ◾ 31
2.2.1.3 Coil Array
Researchers have investigated potential benefits of composing an array of
small coils distributed over the surface of the head for TMS. A variety of
stimulations can be delivered according to spatial and temporal patterns
of currents applied to the coil elements. The magnetic field arising from
the coil array is the superposition of magnetic fields from each coil
element. In principle, the circular and figure-eight coils can be imitated
with correspondingly designed current patterns. Various coils including
these two could be replaced with one coil array. Mathematical methods
such as lead field and minimum-norm estimation can be used to obtain
the pattern of coil currents for a given distribution of stimulating electric
fields in the brain. A unique advantage of the coil array is the ability to
deliver a train of stimulating pulses for different target areas with a short
pulse interval.
Advantages of a coil array have been evaluated based on numerical sim-
ulations.13,14 Figure 2.4 shows coil elements distributed over the surface of
a numerical human head model. This simulation showed that the focality
and depth of stimulation can be improved by properly selecting the coil
current direction and current phases in the neighboring coils.
240
210
180
150
z (mm)
120
90
60
30 240
180
0 120
50 100 60 y (mm)
150 200 0
x (mm)
FIGURE 2.4 Coil array designed on a numerical model of human head. The array
consisted of 40 elements. (From Lu, M. et al., IEEE Transactions on Magnetics 45:
1662–1665, 2009.)
32 ◾ Biomagnetics
Magnetic field
Coil current
Eddy current
(a)
Magnetic field
Coil current
Eddy current
(c)
(b)
FIGURE 2.5 (a) Eddy currents induced by an ordinary figure-eight coil. (b) Eddy
currents induced by an eccentric figure-eight coil. (c) Schematic of a bowl-shaped
coil.
that of the concentric coil. A prototype eccentric coil was designed and
fabricated. Instead of winding a wire around a bobbin, eccentric spiral
slits were formed on the insulator cases, and a wire was woven through the
slits. The coils were used to deliver TMS to healthy subjects. The current
slew rate corresponding to motor threshold values for the concentric and
eccentric coils were 86 and 78 A/μsec, respectively. Results indicated that
the eccentric coil consistently requires a lower driving current to reach the
motor threshold than the concentric coil does.
As discussed in Section 3.4.4, the installation of a TMS system at a patient’s
home has been proposed for providing the continuous therapeutic effects
of daily sessions. Because the figure-eight coil induces a highly localized
electric field, it is challenging to achieve accurate coil positioning above the
targeted brain area without help by a medical expert. A bowl-shaped coil
for stimulating a localized but wider area of the brain was proposed.22,23
The bowl-shaped coil consists of the conductors aligned tangentially to the
surface of the head and the counter conductors running over the tangen-
tial conductors, as shown in Figure 2.5c. Electric fields are induced along
the aligned tangential conductors. The length and width of the arrayed
tangential conductors determine the localization of stimulation. Thus, the
bowl-shaped coil exhibits a moderate and tunable localization of electric
34 ◾ Biomagnetics
2.2.1.5 Cooling Systems
Electric currents in the coil cause heat generation due to the resistance
of the coil.27,28 The amount of heat is proportional to the square of coil
current and to the repetition rate of pulses. Heat generation is prominent
when stimulating with high currents and high repetition rates. This issue
is particularly important in therapeutic applications of TMS because
stimulations are delivered at a repetition rate higher than five pulses per
second. Guidelines are provided in each country that define the maximum
temperature on the coil surface attached to the human body. Coils should
be designed and tested following those guidelines.
One approach for addressing heat generation is to introduce a heat
insulator between the coil conductor and the human head. The heat insu-
lator suppresses the rising temperature on the side attached to the head,
while radiating heat from the opposite side.
Coils should be actively cooled when heat generation is relatively high.
Two major methods are air cooling and liquid cooling. Air-cooling is
provided from a fan installed above the coil or from an external blower
through a tube. Air cooling can be implemented with simple machinery. For
liquid cooling, the coolant is circulated between the coil and an external
Principles of Biomagnetic Stimulation ◾ 35
(a)
4
Search coil voltage (V)
400
2
0 0
–2
–400
–4 Search coil
voltage
–6 –800
–100 0 100 200 300 400
(b) Time (μsec)
FIGURE 2.6 (a) Homemade search coil for measuring magnetic fields generated
from the stimulating coil. (b) Measured search coil voltage and estimation of
magnetic field intensity.
36 ◾ Biomagnetics
to the stimulator coil with its axis perpendicular to the winding plane
of the stimulator coil, and the search coil measures the magnetic field
component of this axis. Induced voltage is recorded using an oscilloscope.
The following formula gives the induced voltage V:
dB
V = πa 2 N (2.8)
dt
where B is the generated magnetic field and a and N are the respective
radius and number of turns of the search coil. The integral of voltage with
respect to time gives the estimation of magnetic field intensity.
t
∫
1
B= 2 V(τ)dτ (2.9)
πa N
0
2.2.2 Driving Circuit
2.2.2.1 Hardware
The driving circuit for magnetic stimulation supplies pulsed electric
currents to the stimulator coil. Figure 2.7 shows a simplified circuit
diagram for generating biphasic pulses.29 The circuit is composed of a
power supply, a high voltage generator, and a semiconductor switch. The
power supply circuit generates DC voltages from the AC input. The high
voltage generator outputs voltages of kilovolts for charging the capacitor.
Resistance limits the current for charging the capacitor. The mechanism of
storing electric charges in the capacitor and then discharging for providing
currents to the coil enables the generation of strong current from relatively
small equipment. When the thyristor is turned on after the positive charging
of the capacitor, electric current begins to flow in the coil. Here, the circuit
is equivalent to the coil L and capacitor C connected serially. Sinusoidal
current flows in these elements with the pulse width T given by
Principles of Biomagnetic Stimulation ◾ 37
Semiconductor switch
Resistance
Power supply
High voltage
AC
generator
Coil
Capacitor
FIGURE 2.7 Diagram of driving circuit for magnetic stimulator. The circuit
provides pulsed currents to the coil. (From Sekino, M. et al., Proceedings of the
ICME International Conference on Complex Medical Engineering 728–733, 2012.)
T = 2π LC (2.10)
case, preparation for the subsequent pulse generation takes a longer time
in comparison with the case of biphasic pulse. Generation of the mono-
phase pulse train usually requires multiple capacitors.
diL
v=L (2.11)
dt
dv
iC = C (2.12)
dt
S1 S2
D v
L
V0 C
R
iC iR iL
FIGURE 2.8 Simplified circuit for analyzing a monophasic pulse current. (From
Mano, Y., and Tsuji, S., Magnetic Stimulation—Basic Principles and Applications.
Ishiyaku Publishers, Tokyo, 2005.)
Principles of Biomagnetic Stimulation ◾ 39
d 2i L i L
L + =0 (2.13)
dt 2 C
C t (2.14)
iL = V0 sin
L LC
diL t
v=L = V0 cos (2.15)
dt LC
After the closing of switch S2, voltage v decreases and becomes zero at
the time t1 = (π/2)(LC)1/2. The voltage v then turns to be negative, and the
diode allows the current to flow also in the resistance R. The Ohm’s low
holds between the voltage v and current iR in the resistance.
v = RiR (2.16)
d 2v 1 dv v
C + + = 0 (2.17)
dt 2 R dt L
When the relation L > 4R2C holds, among the solution of the above
equation the following one is connected to Equation 2.15 at t1 = (π/2)(LC)1/2:
t − t1 ⎞ ⎛ ⎞
exp ⎛ −
C 1 1
v = −2V0 R ⎝ ⎠
sinh ⎜ 2 − ⋅(t − t1 )⎟ (2.18)
L − 4R C
2
2RC ⎝ (2RC) LC ⎠
40 ◾ Biomagnetics
t − t1 ⎞
exp ⎛ −
dv v C
iL = −C − = V0
dt R L − 4R C
2 ⎝ 2RC ⎠
⎡ 4R 2C ⎛ 1 1 ⎞
× ⎢ 1− cosh ⎜ 2 − ⋅(t − t1 )⎟
⎢⎣ L ⎝ (2RC) LC ⎠
⎛ 1 1 ⎞⎤
+ sinh ⎜ 2 − ⋅(t − t1 )⎟ ⎥ (2.19)
⎝ (2RC) LC ⎠⎦
Both voltage v and current iL attenuate toward zero over time. The com-
bination of Equations 2.14 and 2.15 gives the monophasic waveform shown
in Figure 2.2. The magnetic field generated from the coil is proportional to
the current iL , and the electric field induced in the body is proportional to
the time derivative of current diL/dt.
Reflective marker
Binocular camera
Stimulator coil
FIGURE 2.9 Coil navigation system using a near-infrared binocular camera and
reflective markers.
Principles of Biomagnetic Stimulation ◾ 41
on the head, such as the nasion and preauricular points, are measured.
Placing the patient supine leads to stabilized positioning between the head
and marker. The system enables us to see the position of the stimulator coil
relative to the head in real time. The system reads the patient’s magnetic
resonance images for graphically showing the location of the coil and the
target position on the brain. The binocular camera achieves submillimeter
accuracy in estimating the position of markers. Introducing such a
navigation system enables us to precisely and quickly place the stimulator
coil above the target position. This is particularly valuable for highly
localized stimulations using figure-eight coils.
20
Amplitude (uV)
10
SW F3
0 F4
Vin C Fz
G1 G2 –10
Vref
–20
–50 0 50 100 150 200
Active Time (ms)
GND
Control
20
Inion
10
Amplitude (uV)
F3
0 F4
Anti-alias filter Multiplexer Fz
–10
S/H Isolator Computer
–20
A/D –50 0 50 100 150 200
Time (ms)
20
Amplitude (uV)
10
TMS F3
0 F4
SW Fz
–10
Attenuator
–20
–50 0 50 100 150 200
(a) (b) Time (ms)
FIGURE 2.10 (a) Block diagram of TMS-compatible EEG amplifier and timing
chart of the circuit. (b) Measurement points of EEG and stimulated point on
the head. The stimulus point, which is located on the cerebellum, was 20 mm
superior to the inion. (From Iramina, K. et al., Journal of Applied Physics 93:
6718–6720, 2003.)
cerebellum was stimulated. The response at the left brain was large when
the right cerebellum was stimulated. The response map gave information
of the neuronal connectivity. Magnetic stimulation will become a useful
tool for the study of cortical reactivity and neuronal connectivity of the
brain.
An MRI is useful for observing spatial distribution of brain activities.
For demonstrating a magnetic stimulation in the MRI system, the head
and coil are placed in a static magnetic field. Applying a pulsed electric
current to the stimulator coil under the magnetic field causes an impul-
sive Lorentz force. The coil and supporting structure should be durable
against the force. In addition, the magnetic field generated in the MRI
system affects acquisition of images. The timings of pulse generation and
MRI acquisition should be staggered with each other. The cable between
the coil and driving circuit should be long enough so that the driving
circuit is placed distant from the MRI magnet. A lot of studies have been
carried out using simultaneous operation of magnetic stimulation and
MRI.34,35
Principles of Biomagnetic Stimulation ◾ 43
2.3.1.2 Action Potential
When a local depolarization occurred at the cellular membrane by the
excitation stimuli, a voltage-dependent Na+ channel on the membrane
opens. Subsequently, Na+ flows into intracellular space due to the gradi-
ent of concentration and electrical potential between intra-extra cellular
spaces. Corresponding to the decrease in net negative potential of the
membrane owing to the Na+ inflow, large numbers of the Na+ channels
open and the greater Na+ influx occurs. According to the Na+ gate open-
ing, an electric current caused by Na+ inflow exceeds those of K+ from K+
leak channels, and finally the membrane potential shows a positive charge
at the intracellular side, which means the resting potential is reversed.
This phenomenon is known as action potential. The depolarization and
subsequent Na+ channels opening spread to surrounding area, and the
propagation of action potential occurs.
The basic properties of the action potential propagation are analyzable
using the electric circuit model of the cellular membrane. As shown in
Figure 2.11a, a cellular membrane has electric properties rm and cm. The rm
represents penetration resistance per unit length at an ion transition, and
cm is an electric capacity of the membrane. Additionally, a nerve cell also
cm
rm
ri
im (x)
V(x) V(x + Δx)
Ii (x) Ii (x + Δx)
(a) Δx
EL EK ENa c
m
gL gK gNa
ri
im (x)
FIGURE 2.11 (a) A passive cable model for analyzing propagation of action
potential. (b) An equivalent circuit of nerve fiber with the Hodgkin-Huxley
model of voltage-gated channels.
Principles of Biomagnetic Stimulation ◾ 45
has a resistance per unit length ri along to the direction of nerve fibers.
Extracellular cellular resistance is negligible compared with rm; thus, it is
considered as zero in this case. On the basis of these conditions, a mem-
brane electric current im per unit length is represented as
∂V V
im = cm + (2.21)
∂t rm
∂V
ri I i = − (2.22)
∂x
∂I i
im = − (2.23)
∂x
∂2V ∂V
λ2 −V = τ
∂x 2 ∂t
rm
λ=
ri (2.24)
τ = cmrm (2.25)
2.3.2.2 Stimulation Site
Neurons are sensitive to externally applied electric fields generated parallel
to dendrites and axons. In transcranial electric stimulation, as shown in
Principles of Biomagnetic Stimulation ◾ 47
Stimulator coil
Anode
Electric field
Axon
Pyramidal
hillock Electric field
neuron
Interneuron
(a) (b)
FIGURE 2.12 (a) Electric stimulation of neurons in the brain. (b) Magnetic
stimulation of neurons in the brain. Interneurons are susceptible to the induced
currents.
Figure 2.12, excitation occurs at the axon hillock where the axon runs out
from the cell body of pyramidal neurons oriented perpendicular to the sur-
face of the brain. In TMS, on the other hand, the electric field parallel to
the brain surface mainly excites interneurons running in this direction.
This leads to an excitation of connecting pyramidal neurons, and muscle
contraction finally occurs. A stimulation of the primary motor cortex evokes
a twitch. A shift of coil position from the maximally sensitive position causes
a delay of 0.7–0.8 msec in the latency of motor evoked potentials. This delay
corresponds to the time passing one synapse.39 The distribution of induced
electric field strongly depends on the structure of gyrus and sulcus.40
∂2V ∂V ∂ε
λ2 2 −V = τ + λ 2 x (2.26)
∂x ∂t ∂x
48 ◾ Biomagnetics
where the second term in the right side expresses the influence of magnetic
stimulation. This equation shows an interesting point that the influence
of magnetic stimulation is not proportional to the electric field strength
itself but proportional to the spatial gradient of electric field ∂εx/∂x. This
term is referred to as the activating function. When an inhomogeneous
magnetic field is induced depending on the geometry of stimulator coil,
peripheral nerve stimulation is likely to occur at the location exhibiting
high activating function.
2.3.3.2 Stimulation Site
The conductivity of biological tissue differs between tissue types. At the
boundary between bone and muscle, for example, there is a discontinuous
change in conductivity. Liu and Ueno42 calculated the activating function
to study the influence of the interface between tissues with different con-
ductivities on nerve excitation. A Neumann-type boundary condition is
derived for applying the finite element method to calculate the induced elec-
tric field. The spatial distributions of electric fields and activating functions
in homogeneous and inhomogeneous volume conductors are calculated
for comparison. The results show that the interface between conductors of
different conductivities exhibits a large magnitude of activating function.
Such locations are susceptible to magnetic nerve excitations. In addition,
the point where a nerve fiber bends is susceptible to magnetic stimulation
because of a high activating function.
can be increased locally at locations where the field steeply varies or the
target object has complex geometry. Various algorithms for automatically
generating meshes have been investigated. After the generation of meshes,
the nodes are assigned on the sides or on the vertexes of each element, and
an expansion of field equation is derived in terms of unknown parameters
on the nodes. Substitution of this equation into the functional gives an
expansion of functional with the unknown parameters. Because of the
variation principle, this expression takes an extremal value with respect
to a minimal variation of the unknown parameters. The differential of
the expression by the unknown parameters is equal to zero. Doing this
calculation for the all unknown parameters gives a simultaneous linear
equation for the unknown parameters. Numerical analyses lead to the
solution for the unknown parameters.
Because the FEM allows the use of the tetrahedral element, which is
applicable to complicated geometries, this method is appropriate for the
analyses of objects with complicated geometries. However, many numeri-
cal models of the human body consist of cubic voxels originating from
voxels of cross-sectional images. In this case, the boundaries between tis-
sues should be reconstructed to fit the tetrahedral meshes. In the case of
the brain, methods for segmenting tissues from image pixels are devel-
oped. Software for generating tetrahedral mesh is available.
6
⎛ 6
⎞ 6
∑n=1
snφn − ⎜
⎝
∑
n=1
sn ⎟ φ0 = jω
⎠
∑(−1) s l A
n=1
n
nn 0n (2.27)
50 ◾ Biomagnetics
where ϕn is the electric scalar potential at the node n, A0n is the external
magnetic vector potential parallel to the side n of the voxel connecting the
nodes 0 and n, ln is the length of side n, and sn is the conductance of side
n given by
sn = σ nan/ln (2.28)
Here, an is the area of face perpendicular to the side n and σn is the mean
conductivity of four voxels sharing the node n.
For calculating the electric field in the voxel, first, a system of equations
is constructed from Equation 2.27 with the electric scalar potentials ϕn
included as unknown parameters and for all nodes having finite conduc-
tivities. The electric scalar potential is obtained by solving the system of
equations using sparse matrix methods such as successive overrelaxation.
Then, the electric fields on the sides of voxel are calculated from magnetic
vector potentials and electric scalar potentials. The electric field of voxel is
obtained for x, y, and z components as the average of electric fields on the
four sides.
(a) (b)
FIGURE 2.13 (a) A numerical model of the brain consisting of gray matter, white
matter, and cerebrospinal fluid. (b) Numerical simulation of induced current in
the brain using the scalar-potential finite difference method.
(a) (b)
(c) (d)
18
(e) (f ) 0
A/m2
the coil position approaching the forehead. The difference decreased with
an increase in the coil diameter.
∫ B dV (2.29)
1
L= 2
µ0I 2
54 ◾ Biomagnetics
0.56
0.32 4.0
FIGURE 2.15 (a) Distribution of magnetic field on the surface of the head model
in magnetic stimulation to cerebellum. (b) Three-dimensional view of the brain.
(c) Distribution of magnetic field on the surface of the brain. (d) Distribution
of eddy current on the surface of the brain. (From Sekino, M. et al., IEEE
Transactions on Magnetics 42: 3575–3577, 2006.)
12
10
8
6
4
2
0
80 90 100 110
(a) Outer diameter (mm)
14
12
10
8
6
4
2
0
0 10 20 30 40
(b) Inner diameter (mm)
14
12
10
8
6
4
2
0
6 7 8 9 10 11 12
(c) Number of turns
Inductance (µH) Voltage (× 100 V)
Pulse width (× 100 µsec) Resistance (mΩ)
Maximum eddy current density (A/m2)
FIGURE 2.16 Influences of (a) outer diameter, (b) inner diameter, and (c) num-
ber of turns on the characteristics of eccentric figure-eight coil. (From Kato, T.
et al., Journal of Applied Physics 111: 07B322, 2012.)
REFERENCES
1. d’Arsonval, A. 1896. Dispositifs pour la mesure des courants alternatifs de
toutes frequencies. Comptes Rendus de Societe Biologique 2: 450–451.
2. Magnusson, C. E., and Stevens, H. C. 1911. Visual sensations caused by
changes in the strength of a magnetic field. American Journal of Physiology
29: 124–136.
3. Irwin, D. D. Rush, S. Evering, R. Lepeschkin, E. Montgomery, D. B., and
Weggel, R. J. 1970. Stimulation of cardiac muscle by a time-varying magnetic
field. IEEE Transactions on Magnetics 6: 321–322.
4. Maass, J. A., and Asa, M. M. 1970. Contactless nerve stimulation and signal
detection by inductive transducer. IEEE Transactions on Magnetics 6:
322–326.
5. Ueno, S., Harada, K., Ji, C., and Oomura, Y. 1984. Magnetic nerve stimulation
without interlinkage between nerve and magnetic flux. IEEE Transactions
on Magnetics 20: 1660–1662.
6. Barker, A. T. Jalinous, R., and Freeston, I. L. 1985. Non-invasive magnetic
stimulation of human motor cortex. Lancet 1: 1106–1107.
7. Ueno, S., Tashiro, T., and Harada, K. 1988. Localized stimulation of neural
tissues in the brain by means of a paired configuration of time-varying
magnetic fields. Journal of Applied Physics 64: 5862–5864.
Principles of Biomagnetic Stimulation ◾ 57
8. Ueno, S., Matsuda, T., and Fujiki, M. 1990. Functional mapping of the
human motor cortex obtained by focal and vectorial magnetic stimulation
of the brain. IEEE Transactions on Magnetics 26: 1539–1544.
9. Ueno, S. 1994. Biomagnetic Stimulation. New York: Plenum Press.
10. Hallett, M. 2000. Transcranial magnetic stimulation and the human brain.
Nature 406: 147–150.
11. Mano, Y., and Tsuji, S. 2005. Magnetic Stimulation—Basic Principles and
Applications. Ishiyaku Publishers, Tokyo.
12. Ueno, S., and Fujiki, M. 2007. Magnetic stimulation. In Magnetism in Medi
cine, Andra, W., and Nowak, H. (Eds.). Wiley-VCH, Weinheim: 511–528.
13. Ruohonen, J., and Ilmoniemi, R. J. 1998. Focusing and targeting of mag
netic brain stimulation using multiple coils. Medical & Biological Engineer
ing & Computing 36: 297–301.
14. Lu, M., Ueno, S., Thorlin, T., and Persson, M. 2009. Calculating the current
density and electric field in human head by multichannel transcranial mag-
netic stimulation. IEEE Transactions on Magnetics 45: 1662–1665.
15. Grandori, F., and Ravazzani, P. 1991. Magnetic stimulation of the motor
cortex: Theoretical considerations. IEEE Transactions Biomedical Engineering
38: 180–191.
16. Lin, V. W., Hsiao, I. N., and Dhaka, V. 2000. Magnetic coil design considerations
for functional magnetic stimulation. IEEE Transactions on Biomedical
Engineering 47: 600–610.
17. Hsu, K. H., and Durand, D. M. 2001. A 3-D differential coil design
for localized magnetic stimulation. IEEE Transactions on Biomedical
Engineering 48: 1162–1168.
18. Zangen, A., Roth, Y., Voller, B., and Hallett, M. 2005. Transcranial magnetic
stimulation of deep brain regions: Evidence for efficacy of the H-coil.
Clinical Neurophysiology 116: 775–779.
19. Kato, T., Sekino, M., Matsuzaki, T., Nishikawa, A., Saitoh, Y., and Ohsaki,
H. 2011. Fabrication of a prototype magnetic stimulator equipped with
eccentric spiral coils. Proceedings of the Annual International Conference of
the IEEE Engineering in Medicine and Biology Society, 1985–1988.
20. Kato, T., Sekino, M., Matsuzaki, T., Nishikawa, A., Saitoh, Y., and Ohsaki,
H. 2012. Electromagnetic characteristics of eccentric figure-eight coils for
transcranial magnetic stimulation: A numerical study. Journal of Applied
Physics 111: 07B322.
21. Sekino, M., Ohsaki, H., Takiyama, Y., Yamamoto, K., Matsuzaki, T., Yasumuro,
Y., Nishikawa, A., Maruo, T., Hosomi, K., and Saitoh, Y. 2015. Eccentric
figure-eight coils for transcranial magnetic stimulation. Bioelectromagnetics
36: 55–65.
22. Sekino, M., and Suyama, M. 2014. A magnetic stimulator coil with high
robustness to positioning error. URSI General Assembly and Scientific
Symposium, Beijing, China.
23. Yamamoto, K. Suyama, M. Takiyama, Y. Kim, D. Saitoh, Y. Sekino, M.
Characteristics of bowl-shaped coils for transcranial magnetic stimulation.
Journal of Applied Physics 117: 17A318.
58 ◾ Biomagnetics
24. Epstein, C. M., and Davey, K. R. 2002. Iron-core coils for transcranial mag-
netic stimulation. Journal of Clinical Neurophysiology 19: 376–381.
25. Han, B. H., Lee, S. Y., Kim, J. H., and Yi, J. H. 2003. Some technical aspects
of magnetic stimulation coil design with the ferromagnetic effect. Medical
& Biological Engineering & Computing 41: 516–518.
26. Lu, M., and Ueno, S. 2009. Calculating the electric field in real human head
by transcranial magnetic stimulation with shield plate. Journal of Applied
Physics 105: 07B322.
27. Bischoff, C., Machetanz, J., Meyer, B. U., and Conrad, B. 1994. Repetitive
magnetic nerve stimulation: Technical considerations and clinical use in
the assessment of neuromuscular transmission. Electroencephalography
and Clinical Neurophysiology 93: 15–20.
28. Weyh, T., Wendicke, K., Mentschel, C., Zantow, H., and Siebner, H. R. 2005.
Marked differences in the thermal characteristics of figure-of-eight
shaped coils used for repetitive transcranial magnetic stimulation. Clinical
Neurophysiology 116: 1477–1486.
29. Sekino, M., Kato, T., Ohsaki, H., Saitoh, Y., Matsuzaki, T., and Nishikawa,
A. 2012. Eccentric figure-eight magnetic stimulator coils. Proceedings of the
ICME International Conference on Complex Medical Engineering, 728–733.
30. Epstein, C. M. 2008. A six-pound battery-powered portable transcranial
magnetic stimulator. Brain Stimulation 1: 128–130.
31. De Sauvage, R. C., Beuter, A., Lagroye, I., and Veyret, B. 2010. Design and
construction of a portable transcranial magnetic stimulation (TMS) appa-
ratus for migraine treatment. Journal of Medical Devices 4: 015002-1-6.
32. Iramina, K., Maeno, T., Nonaka, Y., and Ueno, S. 2003. Measurement of
evoked electroencephalography induced by transcranial magnetic stimula-
tion. Journal of Applied Physics 93: 6718–6720.
33. Iramina, K., Maeno, T., and Ueno, S. 2004. Topography of EEG responses
evoked by transcranial magnetic stimulation to the cerebellum. IEEE
Transactions on Magnetics 40: 2982–2984.
34. Shitara, H., Shinozaki, T., Takagishi, K., Honda, M., and Hanakawa, T.
2011. Time course and spatial distribution of fMRI signal changes during
single-pulse transcranial magnetic stimulation to the primary motor cor-
tex. Neuroimage 56: 1469–1479.
35. Leitão, J., Thielscher, A., Werner, S., Pohmann, R., and Noppeney, U. 2013.
Effects of parietal TMS on visual and auditory processing at the primary
cortical level—A concurrent TMS-fMRI study. Cerebral Cortex 23: 873–884.
36. Groppa, S., Oliviero, A., Eisen, A., Quartarone, A., Cohen, L. G., Mall, V.,
Kaelin-Lang, A., Mimah, T., Rossi, S., Thickbroom, G. W., Rossini, P. M.,
Ziemann, U., Valls-Solém, J., and Siebner, H. R. 2012. A practical guide to
diagnostic transcranial magnetic stimulation: Report of an IFCN commit-
tee. Clinical Neurophysiology 123: 858–882.
37. Balslev, D., Braet, W., McAllister, C., and Miall, R. C. 2007. Inter-individual
variability in optimal current direction for transcranial magnetic stimula-
tion of the motor cortex. Neuroscience Methods 162: 309–313.
Principles of Biomagnetic Stimulation ◾ 59
38. Kowalski, T., Silny, J., and Buchner, H. 2002. Current density threshold for
the stimulation of neurons in the motor cortex area. Bioelectromagnetics 23:
421–428.
39. Mano, Y., Morita, Y., Tamura, R., Morimoto, S., Takayanagi, T., and Mayer,
R. F. 1993. The site of action of magnetic stimulation of human motor cor-
tex in a patient with motor neuron disease. Journal of Electromyography &
Kinesiology 3: 245–250.
40. Lu, M., Ueno, S., Thorlin, T., and Persson, M. 2008. Calculating the acti-
vating function in the human brain by transcranial magnetic stimulation.
IEEE Transactions on Magnetics 44: 1438–1441.
41. Roth, B. J., and Basser, P. J. 1990. A model of the stimulation of a nerve fiber
by electromagnetic induction. IEEE Transactions on Biomedical Engineering
37: 588–597.
42. Liu, R., and Ueno, S. 2000. Calculating the activating function of nerve
excitation in inhomogeneous volume conductor during magnetic stimula-
tion using the finite element method. IEEE Transactions on Magnetics 36:
1796–1799.
43. Dawson, T. W., and Stuchly, M. A. 1996. Analytic validation of a three-
dimensional scalar-potential finite-difference code for low-frequency magnetic
induction. Applied Computational Electromagnetics Society Journal 11: 63–71.
44. Nagaoka, T., Watanabe, S., Sakurai, K., Kunieda, E., Watanabe, S., Taki, M.,
and Yamanaka, Y. 2004. Development of realistic high-resolution whole-
body voxel models of Japanese adult males and females of average height
and weight, and application of models to radio-frequency electromagnetic-
field dosimetry. Physics in Medicine and Biology 49: 1–15.
45. Laakso, I., Hirata, A., and Ugawa, Y. 2014. Effects of coil orientation on the
electric field induced by TMS over the hand motor area. Physics in Medicine
and Biology 59: 203–218.
46. Opitz, A., Legon, W., Rowlands, A., Bickel, W. K., Paulus, W., and Tyler,
W. J. 2013. Physiological observations validate finite element models for
estimating subject-specific electric field distributions induced by tran-
scranial magnetic stimulation of the human motor cortex. Neuroimage 81:
253–264.
47. Sekino, M., and Ueno, S. 2002. Comparison of current distributions in
electroconvulsive therapy and transcranial magnetic stimulation. Journal
of Applied Physics 91: 8730–8732.
48. Sekino, M., and Ueno, S. 2004. Numerical calculation of eddy currents in
transcranial magnetic stimulation for psychiatric treatment. Neurology and
Clinical Neurophysiology 88: 1–5.
49. Sekino, M., and Ueno, S. 2004. FEM based determination of optimum cur-
rent distribution in transcranial magnetic stimulation as an alternative to
electroconvulsive therapy. IEEE Transactions on Magnetics 40: 2167–2169.
50. Sekino, M., Hirata, M., Sakihara, K., Yorifuji, S., and Ueno, S. 2006. Intensity
and localization of eddy currents in transcranial magnetic stimulation to
the cerebellum. IEEE Transactions on Magnetics 42: 3575–3577.
Chapter 3
Applications of
Biomagnetic Stimulation
for Medical Treatments
and for Brain Research
Shoogo Ueno and Masaki Sekino
CONTENTS
3.1 Introduction 62
3.1.1 Applications in Brain Research 62
3.1.2 Applications in Medicine 63
3.2 Functional Mapping of the Brain 63
3.2.1 Principles 63
3.2.1.1 Characteristics of Figure-Eight Coils 63
3.2.1.2 Spatial Resolution of Brain Mapping 64
3.2.1.3 Electroencephalography Responses Evoked
by Magnetic Stimulation 66
3.2.2 Medical Applications 68
3.3 Applications in Brain Research 69
3.3.1 Virtual Lesion 69
3.3.2 Applications in Cognitive Neuroscience 69
3.3.2.1 Episodic Memory 69
3.3.2.2 Influence on P300 Evoked Potentials 70
3.3.3 Paired-Pulse Stimulation 71
3.3.4 Repetitive Magnetic Stimulation 72
3.3.4.1 Changes in Long-Term Potentiation 72
61
62 ◾ Biomagnetics
3.1 INTRODUCTION
3.1.1 Applications in Brain Research
It is well known that each part of the brain has a different function. In the
early days of neuroscience, knowledge about localization of brain function
was accumulated through observations of patients suffering brain
disorders. Cerebrovascular disorders or injuries occurring in a specific
part of the brain often cause a particular brain function abnormality.
This reveals that a specific part of the brain is necessary for fulfilling the
function. Broca and Welnicke identified the speech center based on such an
approach. Observations of patients with brain disorders played important
roles in understanding the localization of brain function. However, a
methodological limitation existed owing to the uncontrollability of the
location of the disorders. Penfield performed direct electric stimulations
to a patient’s brain during a surgical treatment of epilepsy and recorded
the evoked responses. Various mental activities arose according to the
positions of stimulation, which provided a brief map of the localized brain
function. Because transcranial electric stimulations of the brain cause
pain, mapping of brain function using electric stimulation could not be
performed practically except for patients under craniotomy.
As discussed in Chapter 2, transcranial magnetic stimulation (TMS)
enables noninvasive and localized stimulation of the brain. Such local-
ized stimulation, resembling Penfield’s work, makes it possible to depict a
functional map of the brain even for normal subjects. Moreover, magnetic
stimulation exerted on a region where spontaneous brain activities occur
Biomagnetic Stimulation for Medical Treatments and Brain Research ◾ 63
transiently interferes with these activities. This TMS ability represents the
feasibility of producing a virtual lesion that imitates the above disorders.
Such interference of brain function, once controlled, owns superiority in
terms of reproducibility and safety, and this approach affirmatively facili-
tates systematic investigation of functional localization of the brain.
3.1.2 Applications in Medicine
TMS is used for clinical examinations of neuronal pathways that connect
the primary motor cortex and peripheral muscles. To be specific, TMS
applied to the primary motor cortex evokes contraction of muscles, for
example, in a finger, and the motor evoked potential (MEP) measured
through electrodes provides a criterion. It has been well studied that
amplitudes and latencies of MEP are affected by neurological diseases
such as cerebral infarction and spinal cord injury. These examinations
using TMS are now widely used in clinical practice.
Therapeutic applications of TMS for neurological diseases and mental
illnesses have also been significantly developed in recent years. Repetitive
TMS (rTMS) that leads to functional changes in the central nervous system
is reported to be effective for relieving symptoms of a variety of diseases,
including depression, Parkinson’s disease, and neuropathic pain. TMS
treatment is effective even for some patients who failed in drug treatment,
and TMS hardly causes side effects. The Food and Drug Administration
(FDA) gave clearance for TMS treatment on depression. In addition, ani-
mal experiments showed that preconditioning of the hippocampus using
rTMS induced ischemic tolerance of the hippocampus. With the succes-
sive clinical results of TMS treatment on other diseases, therapeutic appli-
cations of rTMS will continue to progress in the future.
In addition to the brain, magnetic stimulation is applied to nerves in
the pelvic floor to treat urinary incontinence and to relieve muscle fatigue.
The rTMS is a novel treatment complementary to drugs, and it will is
expected to improve through further studies and clinical trials.
1.9 3.4
0.0 0.0
Coil diameter: 40 mm A/m2 Coil diameter: 80 mm A/m2
2.8 3.8
0.0 0.0
Coil diameter: 60 mm A/m2 Coil diameter: 100 mm A/m2
targets compared with other types of coils. When an electric current flows
through the figure-eight coil, the current generates time-varying magnetic
fields from the two circular coils in the opposite directions. Eddy currents
are induced in the target in the direction opposite the coils’ currents so that
the magnetic fields arising from the eddy currents attenuate the primary
magnetic fields generated from the figure-eight coil. Eddy currents in the
two loops flow in the same direction below the middle of the coil. This
convergence of eddy currents leads to the locally increased current density.
Furthermore, the figure-eight coil induces eddy current in a specific direc-
tion at this location. Figure 3.1 shows that the intensity of the eddy current
density in the target increases as the diameter of figure-eight coils becomes
larger. The maximum eddy current density induced by the 100-mm coil was
twice as high as that of the 40-mm coil. The maximum eddy current den-
sity was observed on the surface of the target for all the coils. The magnetic
fields attenuate with the distance from the coil. The above mentioned two
characteristics of figure-eight coils, i.e., the intensity of eddy current density
and the penetration depth, can be adjusted by varying the diameters.
Y Cz
B´
X
C A B
C´
T4
B´
C´
0.5 mV
Onset of magnetic stimulation
5 ms
stimulated, the left thumb twitches. Figure 3.2 shows the recorded MEPs
with varied locations of stimulation. A clear MEP was observed when the
motor cortex corresponding to the left thumb, point A, was stimulated. No
specific waveform was acquired at points B, B′, C, and C′, whose distances
from A were all 5 mm. This result indicates that the localized stimula-
tion of the human motor cortex was achieved with a spatial resolution of
5 mm. Figure 3.3 shows the functional mapping of the regions relevant
to the hand and foot. The arrows show the directions of eddy currents
66 ◾ Biomagnetics
Cz
T4
FIGURE 3.3 Functional distribution of the human motor cortex related to the
hand and foot areas. The arrows show current directions for neural excitation.
The distance between grid points is 5 mm.2
3.2.1.3 Electroencephalography Responses Evoked
by Magnetic Stimulation
The combination of TMS with functional brain imaging such as functional
MRI (fMRI), positron emission tomography (PET), and electroencephalog-
raphy (EEG) is an effective approach for investigating the functional con-
nectivity in the brain. EEG is a useful tool for functional mapping because
of its high temporal resolution. The combination of TMS with simultaneous
EEG provides us the possibility to noninvasively probe the brain’s excitabil-
ity, time-resolved activity, and functional connectivity.3 The experimental
setup for recording EEG evoked by TMS is shown in Figure 3.4. The electri-
cal brain activities are recorded through the electrodes on the surface of the
head. The example of recorded EEG is shown in Figure 3.5.
However, combining TMS and EEG is not simple work owing to the
heat and artifact induced by the interference between TMS and elec-
trodes. Dhuna et al.4 reported the possibility of burn injury by heating the
Biomagnetic Stimulation for Medical Treatments and Brain Research ◾ 67
2 3
1
6
4 5
7 12 13
Stimulated 8 11
9 10
24
14 15 18 19 23
17 20 21 22
28 29 30
27 31 32
25 26 33
39 40
38 41
37
42
36 49 50
35 48 43
51
55 44
34 56 52
46 54
58 59
57 53
2 uV
45
60 0.1 s
3.2.2 Medical Applications
Functional mapping of the brain is applied also for neurosurgery.6 Iden
tifying motor and language areas before surgery is important for prevent-
ing damage to these areas. Identification using TMS is shown in Figure 3.6.
R L
(c)
Both legs
Left leg
(a)
Shoulder
Arm/elbow
Fingers
Central sulcus Thumb
(b) (d)
Word
Pattern
TM
S
0.5
sec
FIGURE 3.7 Timeline (left to right) for presentation of Kanji words and abstract
patterns in one trial of 9-sec duration.7 Pulses of TMS are indicated by paired
vertical arrows. Shaded areas beneath the baseline indicate when cognitive stim-
uli are visible.
%
Correct recall
70
60
50
40
30
20
10
0
Left Vertex Right Off
DLPFC DLPFC head
Magnetic stimulation site
FIGURE 3.8 Overall percent correct for subsequent recall of paired associates
according to site of TMS.7 The comparison of interest is between the left and right
DLPFC, with an active TMS control at the vertex. In a second control, the mag-
netic coil is off the head. The asterisk indicates P < 0.05 for right versus left DLPFC.
auditory stimulation, the delay was not observed. Because the TMS inhib-
its the function of the stimulated area, these results suggest that the cortex
around the left supramarginal gyrus contributes to the generation of P300
at around 200 msec after the auditory stimulation.
3.3.3 Paired-Pulse Stimulation
Paired-pulse stimulation is a protocol that has a short time interval
between the first and second stimulation. The first stimulation is applied
with a low intensity that does not evoke MEP, and the second stimula-
tion is applied with a high intensity that evokes MEP. The influence of the
first stimulation on MEP is evaluated, and using this methodology, one
can reveal the synaptic function in the motor cortex or discuss the related
physiological mechanisms.
In the study by Kobayashi et al.9 the stimulation intensity was fixed at
70–80 percent for the first one, and 120–140 percent for the second one,
and the interval was changed from 1 to 15 msec to investigate the effects
on MEP. Observed MEP had two peaks, N1 and N2, and the increase in
the interval resulted in the higher N2 amplitude, while the N1 ampli-
tude was not affected. The authors concluded that the N1 was the direct
wave (D wave) resulting from the directly stimulated pyramidal neurons,
and the N2 was the indirect wave (I wave) resulting from the pyramidal
72 ◾ Biomagnetics
Pre-tetanus fEPSP
Post-tetanus fEPSP
FIGURE 3.9 Basic mechanisms of LTP.13,20 The plots show the changes in field
EPSP (fEPSP) after LTP induction stimuli (tetanus stimulation).
Biomagnetic Stimulation for Medical Treatments and Brain Research ◾ 73
500
= 0.75 T TMS (rat n = 10)
450 = Sham (rat n = 10)
400 Error bar = ± 1SE
500
= 1.25 T TMS (rat n = 8)
450 = Sham (rat n = 8)
Error bar = ± 1SE
400
% of basal EPSP slope
350
300
250
200
150
100
50
–20 –10 0 10 20 30 40 50 60
(b) Time (min)
FIGURE 3.10 Changes in LTP at rat hippocampus slices after rTMS13,20: (a) LTPs
of 0.75-T stimulated and sham control groups and (b) LTPs of 1.25-T stimulated
and sham control groups.
100 µV
10 msec
Non-ischemia, 5 min ischemia, 10 min ischemia, 30 min ischemia,
stimulated stimulated stimulated stimulated
(a)
400
Non-ischemia, sham
Non-ischemia, stimulated
5 min ischemia, sham
350 5 min ischemia, stimulated
300
**
10 min ischemia, sham
10 min ischemia, stimulated *
3.3.4.3 Neuronal Plasticity
Studies on neuronal plasticity attract attention in association with learning,
memory, and recovery of brain function through rehabilitation. Recent
neuroimaging studies using functional MRI and magnetoencephalography
76 ◾ Biomagnetics
have revealed that cerebral infarction and limb amputation cause partial
changes in the localization of brain function.
The primary somatosensory cortex and primary motor cortex are
located posteriorly and anteriorly, respectively, in the central sulcus.
Assume, for example, that one has had the right-hand amputated. The
brain is intact, and the neurons in the motor and somatosensory areas
corresponding to the right hand suddenly lose their roles. In the following
days and weeks, the localization of brain function changes so that adjacent
brain areas, corresponding to the lip, for example, expand their areas to
irrupt into the right-hand area. Similar changes occur also in the motor
cortex. TMS may be used to interpose these processes for the patient to
obtain better quality of life afterward. TMS may enable such rehabilitation
after a cerebral infarction by stimulating the primary motor area, supple-
mentary motor area, and prefrontal area with prospectively programmed
patterns for appropriate stimulations.
3.4 APPLICATIONS IN MEDICINE
3.4.1 Diagnosis of Nervous System
The most popular method of the magnetic stimulation to examine the
nervous system is the measurement of central motor conduction time
(CMCT).24 The CMCT is the time taken for neural impulses to travel
through the central nervous system on their way to the target muscles.
In this method, the primary motor cortex (M1) is stimulated by TMS
and MEP is recorded. Subsequently, the nerve root on the neck or lumbar
is stimulated with recording MEP. Finally, the CMCTs with these MEP
latencies (cortex latency–spinal cord latency) are calculated. The delay of
CMCT reflects the disturbance of pyramidal tract.
A great number of TMS (or magnetic stimulation) studies for the
inspection of nerve disorders have been reported.25–32 Guillain–Barré syn-
drome (GBS) and chronic inflammatory demyelinating polyneuropathy
(CIDP) are the representative inflammatory demyelinating peripheral
nerve diseases. In several reports, TMS has been applied to these diseases
to inspect a conduction disturbance.25–27 Oshima et al.28 reported the use
of magnetic simulation as a tool to determine the involvement of the cor-
ticospinal tract in GBS patients with hyperreflexia. The authors examined
the central motor conduction, and the results revealed axonal motor neu-
ropathy and normal F-wave conduction in these patients. In this case, the
inspection of nerve conduction by the magnetic stimulation indicated the
Biomagnetic Stimulation for Medical Treatments and Brain Research ◾ 77
2.5
1.5
0.5
0
(a) Control MPTP-sham MPTP-rTMS
### ***
90
80
TH positive neurons (%)
70
60
50
40
30
20
10
0
(b) Saline-sham MPTP-sham MPTP-rTMS
3.4.2.2 Neuropathic Pain
Neuropathic pain is caused by a lesion or dysfunction in the nerve systems.
The nerves affected by some abnormalities (e.g., stroke, cancer, diabetes,
and amputation) send wrong pain signals to the brain even without ade-
quate sensory input. Induced pain like burning, aching, stabbing, shoot-
ing, or shocking electrically is often described.36
As a treatment for the patients suffering neuropathic pain, the efficacy
of brain stimulation has been established. Tsubokawa et al.37 first applied
brain stimulation to patients with thalamic pain with implanting elec-
trodes into the cerebral cortex. In this clinical trial, the direct stimu-
lation of the motor cortex was effective for pain relief. Saitoh et al.38
reported that six out of eight patients experienced pain reduction as a
result of motor cortex stimulation, and this kind of stimulation might be
effective for treating peripheral as well as central deafferentation pain.
The direct stimulation of the motor cortex was effective for relieving
neuropathic pain, and the therapeutic effect was considered strongly
related to the stimulated area. The acquired pain relief supposed that
transmission of the pain signal was modulated by the stimulation of the
ascending nerve tracts.
In order to understand the mechanism of pain perception and the
therapeutic effect of brain stimulation, studies using PET and fMRI have
been performed in both animals and humans.39,40 The results of these
studies showed that brain stimulation of the motor cortex induced brain
activations at the areas of thalamus, cognitive cingulate, and insula. Pain-
relieving effect by brain stimulation is supposed to have a relationship
with brain activations of these areas.
As a noninvasive treatment method of brain stimulation for reliev-
ing neuropathic pain, rTMS has been introduced. Recent clinical studies
showed that stimulation of the motor cortex by rTMS is also effective in
relieving neuropathic pain. Saitoh et al.41,42 reported that the rTMS of the
precentral gyrus at frequencies of 10 Hz reduced pain significantly, com-
pared with lower stimulation of 1 or 5 Hz. The effect of rTMS at higher fre-
quency was greater in patients with a noncerebral lesion than those with
a cerebral lesion. These results indicate that patients with a noncerebral
lesion are more suitable candidates for high-frequency rTMS of the pre-
central gyrus.
80 ◾ Biomagnetics
The therapeutic effect of rTMS for pain relief also depends on the stim-
ulated area. Hirayama et al.41 examined several cortical areas of motor
cortex (M1), the postcentral gyrus (S1), premotor area (preM), and sup-
plementary motor area (SMA) as stimulation targets using a navigation-
guided rTMS. The authors reported 10 of the 20 patients (50 percent)
indicated that stimulation of M1, but not other areas, provided significant
and beneficial pain relief. That group’s result indicated that a significant
effect was lasting for 3 hours after the stimulation of M1 and stimulation
of other targets was not effective. Interestingly, this indicates the M1 is the
sole target for treating intractable pain with rTMS, whereas M1, S1, preM,
and SMA are located adjacently.
In the future, the motor cortex stimulation using rTMS will be recog-
nized as a treatment complementary to drug therapy.
(a)
(b) (c)
REFERENCES
1. Ueno, S., Tashiro, T., and Harada, K. 1988. Localized stimulation of neural
tissues in the brain by means of a paired configuration of time-varying
magnetic fields. Journal of Applied Physics 64: 5862–5864.
2. Ueno, S., Matsuda, T., and Fujiki, M. 1990. Functional mapping of the human
motor cortex obtained by focal and vectorial magnetic stimulation of the
brain. IEEE Transactions on Magnetics 26: 1539–1544.
3. Iramina, K., Maeno, T., and Ueno, S. 2004. Topography of EEG responses
evoked by transcranial magnetic stimulation to the cerebellum. IEEE
Transactions on Magnetics 40: 2982–2984.
84 ◾ Biomagnetics
18. Keck, M. E., Welt, T., Post, A., Muller, M. B., Toschi, N., Wigger, A.,
Landgraf, R., Holsboer, F., and Engelmann, M. 2001. Neuroendocrine and
behavioral effects of repetitive transcranial magnetic stimulation in a psy-
chopathological animal model are suggestive of antidepressant-like effects,
Neuropsychopharmacology 24: 337–349.
19. Post, A., Muller, M. B., Engelmann, M., and Keck, M. E. 1999. Repetitive
trans-cranial magnetic stimulation in rats: Evidence for a neuroprotective
effect in vitro and in vivo. European Journal of Neuroscience 11: 3247–3254.
20. Ogiue-Ikeda, M., Kawato, S., and Ueno, S. 2003. The effect of transcranial
magnetic stimulation on long-term potentiation in rat hippocampus. IEEE
Transactions on Magnetics 39: 3390–3392.
21. Ogiue-Ikeda, M., Kawato, S., and Ueno, S. 2005. Acquisition of ischemic
tolerance by repetitive transcranial magnetic stimulation in the rat hippo-
campus. Brain Research 1037: 7–11.
22. Kirino, T. 1982. Delayed neuronal death in the gerbil hippocampus follow-
ing ischemia. Brain Research 239: 57–69.
23. Glickstein, S. B., Ilch, C. P., Reis, D. J., and Golanov, E. V. 2001. Stimulation
of the subthalamic vasodilator area and fastigial nucleus independently pro-
tects the brain against focal ischemia. Brain Research 912: 47–59.
24. Matsumoto, H., Hanajima, R., Shirota, Y., Hamada, M., Terao, Y.,
Ohminami, S., Furubayashi, T., Nakatani-Enomoto, S., and Ugawa, Y. 2010.
Cortico-conus motor conduction time (CCCT) for leg muscles. Clinical
Neurophysiology 121: 1930–1933.
25. Britton, T. C., Meyer, B. U., Herdmann, J., and Benecke, R. 1990. Clinical
use of the magnetic stimulator in the investigation of peripheral conduction
time. Muscle & Nerve 13: 396–406.
26. Takada, H., and Ravnborg, M. 1997. A comparative study of magnetically
evoked motor potentials (MEP) in polyneuropathy: Demyelinating vs. axo-
nal. Electroencephalography and Clinical Neurophysiology 103: 113.
27. Benecke, R. 1996. Magnetic stimulation in the assessment of peripheral
nerve disorders. Bailliere’s Clinical Neurology 5: 115–128.
28. Oshima, Y., Mitsui, T., Endo, I., Umaki, Y., and Matsumoto, T. 2000.
Corticospinal tract involvement in a variant of Guillain-Barre syndrome.
European Neurology 46: 39–42.
29. Clays, D., Waddy, H. M., Harding, A. E., Murray, N. M. F., and Thomas, P. K.
1990. Hereditary motor and sensory neuropathies and hereditary spastic
paraplegia: A magnetic stimulation study. Annals of Neurology 28: 43–49.
30. Tchen, P. H., Fu, C. C., and Chiu, H. C. 1992. Motor-evoked potentials in
diabetes mellitus. Journal of the Formosan Medical Association 91: 20–23.
31. Maetzu, C., Villoslada, C., Cruz Martínez, A. 1995. Somatosensory evoked
potentials and central motor pathways conduction after magnetic stimula-
tion of the brain in diabetes. Electromyography and Clinical Neurophysiology
35: 443–448.
32. Öge, A. M., Yazici, J., Boyaciyan, A., Eryildiz, D., Ornek, I., Konyalioğlu, R.,
Cengiz, S., Okşak, O. Z., Asar, S., and Baslo, A. 1994. Peripheral and central
conduction in n-hexane polyneuropathy. Muscle & Nerve 17: 1416–1430.
86 ◾ Biomagnetics
33. Chokroverty, S., Shah, S., Chokroverty, M., Deutsch, A., and Belsh, J. 1995.
Percutaneous magnetic coil stimulation of the phrenic nerve roots and
trunk. Electroencephalography and Clinical Neurophysiology 97: 369–374.
34. Pascual-Leone, A., Valls-Sole, J., Brasil-Neto, J. P., Cammarota, A., Grafman, J.,
and Hallett, M. 1994. Akinesia in Parkinson’s disease: II. Effects of subthresh-
old repetitive transcranial motor cortex stimulation. Neurology 44: 892–898.
35. Funamizu, H., Ogiue-Ikeda, M., Mukai, H., Kawato, S., and Ueno, S. 2005.
Acute repetitive transcranial magnetic stimulation reactivates dopaminergic
system in lesion rats. Neuroscience Letters 383: 77–81.
36. Siddall, P. J., Taylor, D. A., McClelland, J. M., Rutkowski, S. B., and Cousins,
M. J. 1999. Pain report and the relationship of pain to physical factors in the
first 6 months following spinal cord injury. Pain 81: 187–197.
37. Tsubokawa, T., Katayama, Y., Yamamoto, T., Hirayama, T., and Koyama,
S. 1991. Treatment of thalamic pain by chronic motor cortex stimulation.
Pacing and Clinical Electrophysiology 14: 131–144.
38. Saitoh, Y., Shibata, M., Hirano, S., Hirata, M., Mashimo, T., and Yoshimine,
T. 2000. Motor cortex stimulation for central and peripheral deafferentation
pain: Report of eight cases. Journal of Neurosurgery 92: 150–155.
39. García-Larrea, L., Peyron, R., Mertens, P., Gregoire, M. C., Lavenne, F., Bars,
D., Le Convers, P., Mauguière, F., Sindou, M., and Laurent, B. 1999. Electrical
stimulation of motor cortex for pain control: A combined PET-scan and
electrophysiological study, Pain 83: 259–273.
40. Kim, D., Chin, Y., Reuveny, A., Sekitani, T., Someya, T., and Sekino, M. 2014.
An MRI-compatible, ultra-thin, flexible stimulator array for functional neu-
roimaging by direct stimulation of the rat brain. Proceedings of the Annual
International Conference of the IEEE Engineering in Medicine and Biology
Society 6702–6705.
41. Hirayama, A., Saitoh, Y., Kishima, H., Shimokawa, T., Oshino, S., Hirata,
M., Kato, A., and Yoshimine, T. 2006. Reduction of intractable deafferenta-
tion pain by navigation-guided repetitive transcranial magnetic stimulation
of the primary motor cortex. Pain 122: 22–27.
42. Saitoh, Y., Hirayama, A., Kishima, H., Shimokawa, T., Oshino, S., Hirata,
M., Tani, N., Kato, A., and Yoshimine, T. 2007. Reduction of intractable deaf-
ferentation pain due to spinal cord or peripheral lesion by high-frequency
repetitive transcranial magnetic stimulation of the primary motor cortex.
Journal of Neurosurgery 107: 555–559.
43. Ueno, S. 2012. Studies on magnetism and bioelectromagnetics for 45 years:
From magnetic analog memory to human brain stimulation and imaging.
Bioelectromagnetics 33: 3–22.
44. Kato, T., Sekino, M., Matsuzaki, T., Nishikawa, A., Saitoh, Y., and Ohsaki, H.
2012. Electromagnetic characteristics of eccentric figure-eight coils for trans
cranial magnetic stimulation: A numerical study. Journal of Applied Physics
111: 07B322.
45. Sekino, M., Kato, T., Ohsaki, H., Saitoh, Y., Matsuzaki, T., and Nishikawa,
A. 2012. Eccentric figure-eight magnetic stimulator coils. Proceedings of the
ICME International Conference on Complex Medical Engineering 728–733.
Biomagnetic Stimulation for Medical Treatments and Brain Research ◾ 87
46. Sekino, M., Ohsaki, H., Takiyama, Y., Yamamoto, K., Matsuzaki, T., Yasumuro,
Y., Nishikawa, A., Maruo, T., Hosomi, K., and Saitoh, Y. 2015. Eccentric
figure-eight coils for transcranial magnetic stimulation. Bioelectromagnetics
36: 55–65.
47. Yamamoto, K., Suyama, M., Takiyama, Y., Kim, D., Saitoh, Y., and Sekino,
M. 2015. Characteristics of bowl-shaped coils for transcranial magnetic
stimulation. Journal of Applied Physics 117, 17A318 (2015); http://dx.doi.org
/10.1063/1.4914876.
48. Okada, A., Nishikawa, A., Fukushima, T., Taniguchi, K., Miyazaki, F., Sekino,
M., Yasumuro, Y., Matsuzaki, T., Hosomi, K., and Saitoh, Y. 2012. Magnetic
navigation system for home use of repetitive transcranial magnetic stimula-
tion (rTMS). Proceedings of the ICME International Conference on Complex
Medical Engineering, 112–118.
49. Rossi, S., Hallett, M., Rossini, P. M., and Pascual-Leone, A. 2009. Safety of
TMS consensus group, safety, ethical considerations, and application guide-
lines for the use of transcranial magnetic stimulation in clinical practice and
research. Clinical Neurophysiology 120: 2008–2039.
Chapter 4
Biomagnetic
Measurements
Sunao Iwaki
CONTENTS
4.1 Introduction 90
4.1.1 History of Magnetoencephalography 90
4.1.2 Clinical Applications 91
4.2 Hardware 92
4.2.1 SQUID 92
4.2.2 Multichannel MEG System 93
4.3 Physiological Background of Biomagnetic Signals 94
4.3.1 Cortical Neural Activity 94
4.3.2 Current Dipole Modeling of Synaptic Activities 95
4.4 Formulation of MEG Signal Generation 96
4.4.1 Maxwell’s Equation and Primary Current 96
4.4.2 Dipolar Current in Volume Conductor 97
4.4.3 Boundary Element Method 98
4.4.4 Lead Field 99
4.5 Neural Source Reconstruction from MEG Signals 100
4.5.1 MEG Inverse Problem 100
4.5.2 Equivalent Current Dipole 101
4.5.2.1 Single ECD Model 101
4.5.2.2 Multidipole Model 101
4.5.2.3 Multiple Signal Classification 104
4.5.3 Distributed Source Imaging 105
4.5.3.1 Minimum-Norm Estimates 106
4.5.3.2 Beamformer 107
89
90 ◾ Biomagnetics
4.1 INTRODUCTION
4.1.1 History of Magnetoencephalography
By the end of the twentieth century, there were several methods developed
to measure human brain activities noninvasively. Magnetoencephalogra
phy (MEG) is one of the completely noninvasive techniques that records
distribution of magnetic fields generated by the neuroelectrical activities
of the human brain by using an array of ultrasensitive magnetic sensors
located around the scalp (Hämäläinen et al., 1993). Since the first recording
in 1929 (Berger, 1929), electroencephalography or EEG, which measures
distributions and temporal changes in the electric potential generated by
cerebral neural activities, was widely used both in clinical applications and
in research on the working human brain. MEG is the magnetic counter-
part of EEG that carries information complementary to that of EEG. More
specifically, MEG mainly records neuroelectrical activity whose dipolar
generator orients tangentially to the scalp, whereas EEG is sensitive to both
tangentially and radially oriented sources. Also, MEG is thought to be use-
ful to localize generators of the neural activities that lie in the cortical area.
In EEG, accurate estimation of source location is often difficult because
the distribution of electrical potentials on the scalp is severely affected by
complex inhomogeneity of conductivity in the head and the poor electri-
cal conductivity of the skull. On the other hand, spatial distribution of
the magnetic field mainly reflects neural currents that flow in the macro-
scopically relatively homogeneous intracranial space (Hämäläinen et al.,
1993). Under such favorable conditions where brain activity is modeled as
a single current dipole, the location of the neural source in the brain can
be determined within a few millimeters spatial accuracy.
Although the first MEG recordings were carried out by using an induction-
coil magnetometer (Cohen, 1968), measurements of MEG signals typi-
cally require more sensitive magnetic sensors such as the superconducting
quantum interference device (SQUID) (Jaklevic et al., 1964) because the
magnetic fields generated by the human brain activity are very weak, on
Biomagnetic Measurements ◾ 91
T (Tesla)
10–4 Geomagnetic field
10–6 Urban magnetic noise
4.1.2 Clinical Applications
Identifying the epileptic foci is the most prevalent use of MEG measure-
ments in the clinical applications that had been studied since the early
1980s (Barth et al., 1982; Hari et al., 1993; Nakasato et al., 1994). Patients
with recurrent and pharmacotherapy-resistant epilepsy could benefit from
epilepsy surgery in which resection of epileptic foci is involved (Choi et al.,
2008). In epileptic surgery, it is crucial to remove only the abnormal tissue
and preserve normal functional tissue especially in the cortical regions.
On the other hand, it is difficult to locate the epileptic foci even when the
abnormal structure such as vascular malformation is detected by using
other imaging modalities. Also, the abnormal tissues including epileptic
foci are often surrounded by eloquent brain areas responsible for essen-
tial functions such as language. For these reasons, it is important to pre-
cisely locate the epileptic foci and dissociate the abnormal area from the
92 ◾ Biomagnetics
4.2 HARDWARE
4.2.1 SQUID
Although the first MEG recordings were carried out by using an induction-
coil magnetometer (Cohen, 1968), measurements of MEG signals typically
require more sensitive magnetic sensors such as a superconducting quan-
tum interference device (SQUID) (Jaklevic et al., 1964). The SQUID is a
superconducting ring with one (dc SQUID) or two (RF SQUID) insulating
layers to form Josephson junction (Josephson, 1972). In the current instru-
mentation of MEG system, dc SQUID, which has more sensitivity to tiny
changes in the magnetic field compared to RF SQUID (Clarke et al., 1976),
is usually used. The dc SQUID has a couple of Josephson junctions in par-
allel in the superconducting ring (Figure 4.2). Without external magnetic
flux passing through the SQUID ring, the input bias current Ib is split
equally between both junctions. When a small external magnetic field
is applied to the SQUID ring, Josephson junctions, which are character-
ized by the maximum critical current, limit the flow of induced currents
without losing the superconductivity and magnetic flux quantization that
occurs in the superconducting ring where magnetic flux through the ring
Biomagnetic Measurements ◾ 93
Ib
V
I0 Φ I0
FIGURE 4.3 Gradiometric pickup coils to reduce disturbance from the external
magnetic noise.
insensitive to uniform magnetic field generated far away from the sensor
(Vrba et al., 1982; Duret and Karp, 1984) (Figure 4.3).
The first helmet-shaped MEG system with whole-scalp coverage using
122-channel planar-gradiometer SQUID sensor array was built in 1992
(Ahonen et al., 1993); it enabled users to monitor the entire cortical neu-
ral activities simultaneously. Since the early 1990s, several manufactur-
ers made the helmet-shaped MEG systems commercially available (Vrba
et al., 1993; Kado et al., 1999). Currently, systems with more than 300
SQUID sensors are available and are being used in many academic and
clinical institutes.
4.3 PHYSIOLOGICAL BACKGROUND
OF BIOMAGNETIC SIGNALS
4.3.1 Cortical Neural Activity
The human brain is composed of neurons and glial cells. The glial cells
work primarily for the structural support for neurons and for regulating
the internal environment by maintaining the concentration of cerebro-
spinal fluid and transporting substances between blood vessels and neu-
rons to keep them working. Neurons are the electrically excitable cells that
process and transmit information in the brain. The neuron is composed
of a cell body (soma), dendrites, and an axon. The soma and dendrites are
mostly located in the gray matter, which is distributed on the surface of
the cerebral cortex. Inside the gray matter is a tissue called white mat-
ter that mostly consists of the glial cells and the myelinated axons, which
connects between different brain structures. The distal terminal of the
axon is connected to the dendrites of another neuron through synapse
Biomagnetic Measurements ◾ 95
Dendrites Soma
Axon
FIGURE 4.4 Equivalent current dipole model of the cortical neural activity.
96 ◾ Biomagnetics
where δ(r) is the Dirac delta function and Q is the amplitude of the equiva-
lent current dipole.
Biomagnetic Measurements ◾ 97
µ0 J(r ′)× R
B(r) =
4π ∫ R3
d v ′ (4.3)
B(r) = B 0 (r)+
µ0
4π ∑(σ − σ )∫ V(r ′) RR × dS′ (4.5)
ij
i j 3 ij
Sij
FIGURE 4.5 Configuration of the neural source and the magnetic field
measurement.
98 ◾ Biomagnetics
where V0 and B0 are the distributions of the electric potential and the mag-
netic field produced by Jp and are calculated as (Geselowitz, 1967)
∇′ ⋅ J p
∫ ∫
1 1 1
V0 (r) = − J p (r ′)⋅ ∇′ d v ′ = d v ′ (4.6)
4πσ 0 R 4πσ 0 R
G G
µ0 p
∫
R
B 0 (r) = J (r ′)× 3 d v ′ (4.7)
4π R
G
µ 0 FQ × rQ − (Q × rQ ⋅r)∇F(r,rQ )
B(r) = (4.8)
4π F(r,rQ )2
where
ni
V =
i
∑ H V + g (4.12)
j=1
ij j i
∫
1 2
g ki = V0 (r ′)d S ′i (4.13)
µ k σ i + σ i+
i −
∆ ik
1 σ −j − σ +j 1
H klij =
2π σ i− + σ i+ µ ik ∫ ∆ ik
Ω lj (r ′)d S ′l (4.14)
the triangle ∆ k from the point r (van Oosterom and Strackee, 1983).
i
In solving the linear system Equation 4.12, the inverting matrix becomes
singular reflecting the arbitrariness in determining the reference point to
define the potential distribution. The mathematical technique called defla-
tion can be used to avoid this singularity issue (Lynn and Timlake, 1968).
After obtaining the distribution of the potential distribution V on each
surface, the magnetic field distribution can be calculated by directly inte-
grating Equation 4.5 on the discretized surfaces.
4.4.4 Lead Field
Section 4.4.3 describes the numerical methods to calculate spatial dis-
tribution of the electrical potential on the scalp (i.e., EEG) and the mag-
netic field outside of the head (i.e., MEG) for given primary neural current
distribution Jp. As shown in the equations, there is a linear relationship
between MEG distribution B and amplitudes of the primary current dis-
tribution Jp. Suppose bi is the amplitude of the magnetic field measured by
100 ◾ Biomagnetics
∫
bi = Λ i (r)⋅ J p (r)dv (4.15)
where vector field Λi is usually called the “lead field.” The lead field is
uniquely defined by the conductivity distribution σ(r) and the configura-
tion of the magnetic field sensor (i.e., location of the sensor and the direc-
tion of the pickup coil).
When employing a set of equivalent current dipoles shown in Equation
4.2 as a model of primary current distribution, the MEG signal generated
by the n current dipoles with dipole moments Q j located at rQ j (j = 1, …, n)
and measured by ith sensor is described as
bi = ∑ Λ (r
j=1
i Qj ) ⋅Q j (4.16)
J Forward problem
B = f(J) + n
Bm = fˆ(x)
Inverse problem
Source current Measured
distribution magnetic field
4.5.2.2 Multidipole Model
In the multiple dipole model of the primary current distribution, we
assume that the distribution of the measured magnetic field is generated
102 ◾ Biomagnetics
2
E = b meas − b model (4.17)
2
b − b model (4.18)
g = 1− meas 2
b meas
which indicates how well the field pattern predicted by the equivalent
dipoles agrees with that measured in the experiment.
Another important point in using the multidipole model is how to
determine the number of dipoles to explain the current field distribution.
Typically, multiple sources with their distance larger than a few centi-
meters show field patterns that have only minor overlap if their orienta-
tions are favorable. In these cases, location, amplitude, and orientation of
each primary source may be estimated successfully by applying the single
a a´
Source Sink
d
a a´
d/√2
FIGURE 4.7 Simple geometric method to estimate location of the single ECD
from magnetic field distribution.
Biomagnetic Measurements ◾ 103
dipole model separately. On the other hand, if the field pattern does not
show obvious separation between sources, spatiotemporal characteristics
of the measured MEG data may be used to estimate the number of active
source dipoles (Scherg, 1990 for review) to be assumed in the multidipole
model. The standard method to estimate the number of active sources in
the measured spatiotemporal MEG data is to use singular value decompo-
sition (SVD) (de Munck, 1990).
Emin ≥ ∑s
i=r+1
2
i (4.20)
∑s 2
i
g≤ i=1
m (4.21)
∑s
i=r+1
2
i
Dipole ∆pi+1 = pi – ∆d ∂E
parameters ∂pi
Nonlinear
parameter bmeas – bmodel
GOF = 1 – 100 [%]
optimization bmeas
Calculated E Measured
Forward data Error data MEG
model calculation system
bmodel bmeas
⎡ Ds ⎤ ⎡ Vs ⎤
R meas = VDV T = [ Vs Vn ] ⎢ ⎥⎢ ⎥ (4.22)
⎢⎣ Dn ⎥⎦ ⎢⎣ Vn ⎥⎦
Biomagnetic Measurements ◾ 105
2
Si = P⊥ ⋅ Λ̂ i (4.23)
F
Scanning grids
Measured data matrix: Bmeas
Source location
measured at M sensors, and Λ = [Λ1, Λ2, …, ΛM] as the lead field matrix, the
MEG forward model for distributed source model is written as
bmeans = Λ · Q + n (4.24)
where n is the noise in the measured data.
4.5.3.1 Minimum-Norm Estimates
The linear equation (Equation 4.24) does not generally have a unique solu-
tion because of the underdetermined nature of the system. The number of
parameters to be estimated (N) is on the order of thousands, whereas the typi-
cal number of equations, which corresponds to the number of sensors (M),
is on the order of hundreds at most. In the traditional inverse formulation, a
“minimum-norm constraint” is commonly employed in which the least norm
solution is selected from among the possible set of solutions (Hämäläinen and
Ilmoniemi, 1984). The solution under minimum-norm constraint is given by
ˆ = Λ − b meas (4.25)
Q
where Λ− is the generalized inverse (Menke, 1984) of the lead field matrix
and is written as
( )
−1 −1 −1
The diagonal matrix Wc contains the weighting factors for “depth nor-
malization” to avoid minimum-norm solution bias toward superficial
locations (Jeffs et al., 1987), and γ is a regularization parameter that con-
trols the degree of regularization and is determined in accordance with
the signal-to-noise ratio of measured signal. Using this method, the solu-
tion Q̂ that minimizes the objective function,
2 2
Ec = b meas − b model + γ 2 Wc Q̂ (4.27)
Wc = diag{||Λi2||} (4.28)
4.5.3.2 Beamformer
Beamformer is an adaptive array signal processing technique originally
developed for radar and seismic research (van Veen and Buckley, 1988).
Beamforming techniques can be successfully applied to form an adaptive
spatial filter to extract time series that arise from the neural activity at a
specific brain location and block the signal from other parts of the brain
with linear weighted sum of the MEG signals measured by the sensor
108 ◾ Biomagnetics
Beamforming
{ } {
min trace ⎡⎣ C ( y(r0 )) ⎤⎦ = min ⎡⎣ FT (r0 )C(B)F(r0 )⎤⎦
F( r0 ) F( r0 )
}
where F(r0) is the filter weights for the cortical point of interest r0, C(y(r0))
is a covariance matrix of the neural activity at r0, which is not observable,
C(B) is a covariance matrix of the spatiotemporal measure data B, and
Λ(r0) is a lead field vector at r0. By using the methods of Lagrange multipli-
ers, the solution for the minimization problem is obtained as
Rotation angle
(a) (b)
FIGURE 4.11 Examples of the visual stimuli used in the mental rotation task.
(a) Rotationally-symmetric pair. (b) Minor-reversed pair.
110 ◾ Biomagnetics
t = 180 msec
t = 210 msec
t = 240 msec
(a)
t = 180 msec
(b)
FIGURE 4.12 Results of the neural source imaging during the performance of
the mental rotation task obtained by the weighted minimum-norm method.
(a) Mental rotation. (b) Control.
240 msec was also observed. On the other hand, activity in the primary
visual area was detected in the control condition in which subjects were
not required to rotate the 3D objects (Iwaki et al., 1999).
This example suggests that the distributed source imaging techniques
such as weighted minimum-norm estimates are capable of visualizing
brain activities without a priori knowledge of the number of active sources
when they are located a few centimeters apart from each other.
quite robustly by the MEG sensor covering the temporal cortex. Within
each hemisphere, the MEG field pattern shows that the source of N1m can
be well explained by a single ECD. The results of the single ECD estima-
tion in each hemisphere show that the neural sources of N1m located in
the primary auditory are in the Heschl’s gyrus on the superior temporal
gyrus (Figure 4.13).
The N1m component is evoked in response not only to the onset of the
auditory stimuli but also to the offset of the stimuli. This offset response
N1moff is also measured for the auditory language stream while the sub-
jects listened to (1) a series of spoken sentences and (2) white noise whose
amplitude was modulated by the envelope of the spoken sentences as con-
trol stimuli, and both included sudden and brief interruptions (Figure
4.14a). Auditory evoked magnetic fields corresponding to the interrup-
tion and resumption of the speech were measured. Statistically significant
enhancement was observed only in the N1m component of the auditory
evoked offset response (N1moff ) to the interruptions of the verbal stream
compared to the onset response (N1mon) to the resumptions of the stream
(Figure 4.14b) (Iwaki et al., 2012). These results suggest that the human
auditory cortex plays an important role in perceiving interruptions of ver-
bal streams.
Stimuli
(a)
(b)
FIGURE 4.13 (a) Waveforms and the spatial distribution of the N1m auditory
evoked magnetic fields. (b) Results of the neural source estimation from N1m
MEG distribution. The dipoles are located in the bilateral primary auditory
cortex.
112 ◾ Biomagnetics
–4 dB
–9 dB
–15 dB 500 msec
100 msec
2.90 3.00 3.10 3.20 3.30 3.40 3.50
(a) Elapsed time from the beginning of the recorded speech (sec)
Stimuli to the left ear Stimuli to the right ear Stimuli to the left ear Stimuli to the right ear
mean (+–SE) peak RMS amplitude (fT/cm)
50 50 50 50
N1moff
N1mon
N1moff
N1mon
N1moff
N1mon
N1moff
N1mon
N1moff
N1mon
N1moff
N1mon
N1moff
N1mon
N1moff
N1mon
(b)
FIGURE 4.14 (a) Typical acoustic waveforms of the sudden interruptions with
500 msec duration embedded in the natural spoken language stream. (b) Mean
and standard error of the N1moff and N1mon peak amplitudes elicited by the inter-
ruption of speech stream (SPEECH) and corresponding white noise (NOISE).
4.6.3 Multimodal Neuroimaging
Although MEG can record temporal changes of post-synaptic neural activity
with good temporal resolution (on the order of milliseconds), the principal
difficulty in interpreting MEG data is in reconstructing the spatial distribu-
tion of the neural activity in the brain from MEG spatiotemporal distribu-
tions measured at sensors located at a distance from the brain area from
which activity is being assessed. On the other hand, fMRI makes it pos-
sible to precisely visualize the spatial distribution of human brain activity
with a resolution that is typically approximately a few millimeters; however,
fMRI measures hemodynamic changes (blood oxygenation level dependent
[BOLD] signals) (Ogawa et al., 1990; Belliveau et al., 1991), which are an
indirect measure of neural activity. These changes peak a few seconds after
the onset of neural firing in an area; therefore, this technique has limited
temporal resolution. On the basis of these inherent limitations in each of the
measurement modalities, it is natural to combine data from both techniques
to enhance both the spatial and temporal resolutions. As stated in Section
4.5.3.1, the weighted minimum-norm technique can be used to impose
Biomagnetic Measurements ◾ 113
Parieto-occipital
junction
280 msec
(a)
Intra-parietal
area
340 msec
Inferior-
380 msec temporal area
(b)
FIGURE 4.15 (a) Example of the visual stimuli used in the 3D-SFM experiment.
(b) Results of the fMRI-weighted MEG inverse analysis in the typical subject
performing the 3D-SFM task. The blue circles denote the sites where we observed
the major neural activity during 3D-SFM processing.
114 ◾ Biomagnetics
The results show that the neural activity in the parieto-occipital, the
intraparietal, and the inferior-temporal regions as well as primary visual
area were increased during the 3D SFM at the latencies between 200
and 450 msec after the onset of visual motion (Iwaki et al., 2013). These
results demonstrate that the noninvasive multimodal neuroimaging, in
which structural and functional MRI data are incorporated into the MEG
inverse problem, is capable of visualizing neural dynamics of the human
brain with spatial accuracy comparable to fMRI without compromising
the excellent temporal accuracy of MEG (Iwaki, 2012).
REFERENCES
Ahonen, A. I., Hämäläinen, M. S., Kajola, M. J., Knuutila, J. E. T., Laine, P. L.,
Lounasmaa, O. V., Parkkonen, L. T., Simola, J. T., and Tesche, C. D. 1993.
122-channel SQUID instrument for investigating the magnetic signals from
the human brain. Physica Scripta T49A: 198–205.
Barnard, A. C. L., Duck, I. M., Lynn, M. S., and Timlake, W. P. 1967. The applica-
tion of electromagnetic theory to electrocardiography. II. Numerical solu-
tion of the integral equations. Biophysics Journal 7: 463–491.
Barth, D., Sutherling, W., Engel, J. J., and Beatty, J. 1982. Neuromagnetic localiza-
tion of epileptiform spike activity in the human brain. Science 218: 891–894.
Beers, A. L., Watanabe, T., Ni, R., Sasaki, Y., and Anderson, G. J. 2009. 3D surface
perception from motion involves a temporal–parietal network. European
Journal of Neuroscience 30: 703–713.
Belliveau, J. W., Kennedy, D. N., McKinstry, R. C., Buchbinder, B. R., Weisskoff,
R. M., Cohen, M. S., Vevea, J. M., Brady, T. J., and Rosen, B. R. 1991.
Functional mapping of the human visual cortex by magnetic resonance
imaging. Science 254: 716–719.
Berger, H. 1929. Über das Elektroenkephalogramm des Menschen. Archiv für Psy
chiatrie und Nervenkrankheiten 87: 527–570.
Brenner, D., Williamson, S. J., and Kaufman, L. 1975. Visually evoked magnetic
fields of the human brain. Science 190: 480–481.
Brenner, D., Lipton, J., Kaufman, L., and Williamson, S. J. 1978. Somatically evoked
magnetic fields of the human brain. Science 199: 81–83.
Chapman, R. M., Ilmoniemi, R. J., Barbanera, S., and Romani, G. L. 1984. Selective
localization of alpha brain activity with neuromagnetic measurements.
Electroencephalography and Clinical Neurophysiology 58: 569–572.
Choi, H., Sell, R. L., Lenert, L., Muennig, P., Goodman, R. R., Gilliam, F. G., and
Wong, J. B. 2008. Epilepsy surgery for pharmacoresistant temporal lobe epi-
lepsy: A decision analysis. JAMA 300: 2497–2505.
Clarke, J. 1994. SQUIDs. Scientific American 271: 46.
Clarke, J., Goubau, W. M., and Ketchen, M. B. 1976. Tunnel junction dc SQUID
fabrication, operation, and performance. Journal of Low Temperature Physics
25: 99–144.
Biomagnetic Measurements ◾ 115
Knowlton, R. C., Razdan, S. N., Limdi, N., Elgavish, R. A., Killen, J., Blount, J.,
Burneo, J. G., Ver Hoef, L., Paige, L., Faught, E., Kankirawatana, P., Bartolucci,
A., Riley, K., and Kuzniecky, R. 2009. Effect of epilepsy magnetic source imag-
ing on intracranial electrode placement. Annals of Neurology 65: 716–723.
Knuutila, J., Ahlfors, S., Ahonen, A., Hällström, J., Kajola, M., Lounasmaa, O. V.,
Vilkman, V., and Tesche, C. 1987. Large-area low-noise seven-channel dc
SQUID magnetometer for brain research. Review of Scientific Instruments
58: 2145–2156.
Levenberg, K. 1944. A method for the solution of certain problems in least squares.
Quarterly of Applied Mathematics 2: 164–168.
Liu, A. K., Belliveau, J. W., and Dale, A. M. 1998. Spatiotemporal imaging of human
brain activity using functional MRI constrained magnetoencephalogra-
phy data: Monte Carlo simulations. Proceedings of the National Academy of
Sciences of the United States of America 95: 8945–8950.
Lynn, M. S., and Timlake, W. P. 1968. The use of multiple deflations in the numeri-
cal solution of singular systems of equations to potential theory. SIAM
Journal on Numerical Analysis 5: 303–322.
Marquardt, D. 1963. An algorithm for least-squares estimation of nonlinear
parameters. SIAM Journal of Applied Mathematics 11: 431–441.
McNay, D., Michielssen, E., Rogers, R. L., Taylor, S. A., Akhtari, M., and Sutherling,
W. W. 1996. Multiple source localization using genetic algorithms. Journal of
Neuroscience Methods 64: 163–182.
Menke, W. 1984. Geophysical Data Analysis: Discrete Inverse Theory. Academic,
New York.
Mosher, J. C., Lewis, P. S., and Leahy R. 1992. Multiple dipole modeling and local-
ization from spatio-temporal MEG data. IEEE Transactions on Biomedical
Engineering 39: 541–557.
Nakasato, N., Levesque, M. F., Barth, D. S., Baumgartner, C., Rogers, R. L., and
Sutherling W. W. 1994. Comparisons of MEG, EEG, and ECoG source local-
ization in neocortical partial epilepsy in humans. Electroencephalography
and Clinical Neurophysiology 91: 171–178.
Ogawa, S., Lee, T. M., Nayak, A. S., and Glynn, P. 1990. Oxygenation-sensitive
contrast in magnetic resonance image or rodent brain at high magnetic field.
Magnetic Resonance in Medicine 14: 68–78.
Ohta, H., Takahata, M., Takahashi, Y., Shinada, K., Yamada, Y., Hanasaka, T.,
Uchikawa, Y., Kotani, M., Matsui, T., and Komiyama, B. 1989. Seven-channel
rf SQUID with 1/f noises only at very low frequencies. IEEE Transactions on
Magnetics 25: 1018–1021.
Oostendorp, T. F., and van Oosterom, A. 1989. Source parameter estimation in
inhomogeneous volume conductors of arbitrary shape. IEEE Transactions on
Biomedical Engineering 36: 382–391.
Orban, G. A., Sunaert, S., Todd, J. T., Hecke, P. V., and Marchal, G. 1999. Human cor-
tical regions involved in extracting depth from motion. Neuron 24: 929–940.
Pantev, C., Makaig, S., Hoke, M., Galambos, R., Hampson, S., and Gallan, C. 1991.
Human auditory evoked gamma-band magnetic fields. Proceedings of the
National Academy of Sciences of the United States of America 88: 8996–9000.
118 ◾ Biomagnetics
Papanicolaou, A. C., Simos, P. G., Castillo, E. M., Breier, J. I., Sarkari, S., Pataraia,
E., Billingsley, R. L., Buchanan, S., Wheless, J., Maggio, V., and Maggio, W. W.
2004. Magnetocephalography: A noninvasive alternative to the Wada proce-
dure. Journal of Neurosurgery 100: 867–876.
Paradis, A. L., Cornilleau-Peres, V., Droulez, J., Van de Moortele, P. F., Lobel, E.,
Berthoz, A., Le Bihan, D., and Poline, J. B. 2000. Visual perception of motion
and 3-D structure from motion: an fMRI study. Cerebral Cortex 10: 772–783.
Parsons, L. M., Fox, P. T., Downs, J. H., Glass, T., Hirsch, T. B., Martin, C. C.,
Jerabek, P. A., and Lancaster, J. 1985. Use of implicit motor imagery for visual
shape discrimination as revealed by PET. Nature 375: 54–58.
Peronnet, F., and Farah, M. J. 1989. Mental rotation: An event-related potential
study with a validated mental rotation task. Brain and Cognition 9: 279–288.
Pirmoradi, M., Beland, R., Nguyen, D. K., Bacon, B. A., and Lassonde, M. 2010.
Language tasks used for the presurgical assessment of epileptic patients with
MEG. Epileptic Disorders 12: 97–108.
Ploner, M., Gross, J., Timmermann, L., and Schnitzler, A. 2002. Cortical repre-
sentation of first and second pain sensation in humans. Proceedings of the
National Academy of Sciences of the United States of America 99: 12444–12448.
Raij, T. T., Forss, N., Stancak, A., and Hari, R. 2004. Modulation of motor-cortex
oscillatory activity by painful Adelta- and C-fiber stimuli. NeuroImage 23:
569–573.
Romani, G. L., Leoni, R., and Salustri C. 1985. Multichannel instrumentation
for biomagnetism. In SQUID ’85: Superconducting Quantum Interference
Devices and their Applications, Hahlbohm, H. D. and Lübbig, H., eds. Walter
de Gruyter, Berlin: 919–932.
Rossini, P. M., Altamura, C., Ferretti, A., Vernieri, F., Zappasodi, F., Caulo, M.,
Pizzella, V., Del Gratta, C., Romani, G. L., and Tecchio, F. 2004. Does cere-
brovascular disease affect the coupling between neuronal activity and local
haemodynamics? Brain 127: 99–110.
Salmelin, R., Service, E., Kiesilä, P., Uutela, K., and Salonen, O. 1996. Impaired
visual word processing in dyslexia revealed with magnetoencephalography.
Annals of Neurology 40: 157–162.
Sarvas, J. 1987. Basic mathematical and electromagnetic concepts of the biomag-
netic inverse problem. Physics in Medicine and Biology 32: 11–22.
Scherg, M. 1990. Fundamentals of dipole source potential analysis. In Auditory
Evoked Magnetic Fields and Electric Potentials, Advances in Audiology, vol. 6.
Grandori, F., Hoke, M., and Romani, G. L., eds. Karger, Basel: 40–69.
Schmidt, R. O. 1986. Multiple emitter location and signal parameter estimation.
IEEE Transactions on Antennas and Propagation AP-34: 276–280.
Sekihara, K., Poeppel, D., Marantz, A., Phillips, C., Koizumi, H., and Miyashita, Y.
1998. MEG covariance difference analysis: A method to extract target source
activities by using task and control measurements. IEEE Transactions on
Biomedical Engineering 45: 87–97.
Shepard, R. N., and Metzler, J. 1971. Mental rotation of three dimensional objects.
Science 171: 701–703.
Biomagnetic Measurements ◾ 119
Principles of Magnetic
Resonance Imaging
Masaki Sekino, Norio Iriguchi, and Shoogo Ueno
CONTENTS
5.1 Introduction 121
5.2 Magnetic Resonance Signal and Relaxation Times 123
5.2.1 Electromotive Force 123
5.3 Overview of the Spin-Warp Imaging Technique 127
5.3.1 Recognition of Nuclei Distributed in the First Direction 128
5.3.2 Recognition of Nuclei Distributed in the Second Direction 129
5.3.3 Recognition of Nuclei Distributed in the Third Direction 131
5.3.4 k Space 133
5.3.5 Image Contrast 135
5.4 Diversification of MRI Application Techniques 137
5.4.1 Magnetic Resonance Angiography 137
5.4.2 Perfusion and Diffusion Imaging 138
5.4.3 Functional Neuroimaging 140
5.4.4 Magnetic Resonance Spectroscopy 140
5.4.5 MRI of Other Nuclei than Protons 141
5.5 Concluding Remarks 142
References 142
5.1 INTRODUCTION
Magnetic resonance imaging (MRI) is a way of producing cross-sectional
images of the human body noninvasively. A proton (or hydrogen nucleus)
is spinning with a positive charge, and it is generating a magnetic flux in
the direction of the Ampere’s law. Therefore, protons can be regarded as
121
122 ◾ Biomagnetics
small, spinning magnets with north poles and south poles in Newtonian
physics. When a part of the body is exposed to a static magnetic field, each
spinning magnet makes a precession around the axis of the static mag-
netic field. If a radiofrequency (rf) wave is then transmitted, some of the
magnets begin to make a precession with the same phase of the transmit-
ted rf wave. This is the magnetic resonance (MR). In the resonating state,
a large number of magnets establish a large magnetic flux rotating around
the axis of the static magnetic field at the frequency of the transmitted rf
wave. The rf wave is then turned off, and subsequently an electrical coil
tuned at the proper rf frequency picks up the electromotive force (EMF)
before the rotating magnetic flux decays in amplitude. Resonance fre-
quency is proportional to the external magnetic field. Hence, modifying
the static magnetic field by using a gradient magnetic field, the position
of each proton can be distinguished, and a computer can reconstruct an
image from the EMF signal.
The first successful nuclear magnetic resonance (NMR) experiments
were reported by two independent groups in the United States (Bloch
et al., 1946; Purcell et al., 1946). In 1952, Purcell and Bloch received the
Nobel Prize for their observations. Since chemical shift phenomena were
discovered in the 1950s, NMR techniques were developed as important
tools for chemical analyses. The early experiments were limited in scope
by the relatively poor instrumentation available at that time. In the late
1960s, superconducting magnets were introduced to NMR experiments
and revolutionized the scope of NMR together with the emergence of
Fourier transform NMR (Ernst and Anderson, 1966). Lauterbur (1973)
published the first NMR image of two small tubes of water. By 1980, the
clinical evaluation of MRI had begun, and since that time, there have been
continuing developments in instrumentation and applications that have
led to where we are today. MRI is now established as an important modal-
ity in medical practice. The application techniques of MRI have diversified
in recent years. Those techniques are, for example, magnetic resonance
angiography (MRA), perfusion and diffusion imaging, and functional
neuroimaging (fMRI). Scanning time has been dramatically shortened.
Because there are clear attractions in using noninvasive methods for the
study of living systems, those MRI techniques that investigate the brain
and tissue characteristics appear to have a very bright future. Therefore,
in this chapter, we give an overview of the principles of MRI techniques.
Principles of Magnetic Resonance Imaging ◾ 123
Magnetization M0
Flip angle θ
= γB1Δt
a counter magnetic flux, and this is experienced by the rf coil as the resis-
tance of the coil itself. When the electrical rf coil and the sample are sup-
plied with the rf power P, the current i* of the coil is i* = (P/R)1/2, where
R is the equivalent, gross resistance of the rf coil and the sample. If B1 is
defined as the magnetic flux density produced by a coil carrying unit cur-
rent, the magnetic flux density B1* = B1 ( P /R )1/2. The B1 value of the rf coil
can be determined also by the geometrical configuration of the rf coil.
Figure 5.2 shows the magnetization M 0* of a volume of interest (VOI) of
a sample at the time when the lip angle θ has been made 90 degree by the
rf field (B1 cos ω 0t) generated by the coil, carrying unit current i = cos ω 0t.
Magnetic resonance signals can be acquired with the same coil, and the
electromotive force (EMF) ξs induced by the magnetization M 0* is given by
( )
ξ s = −d B1* M 0* /dt
ξ s = ω 0 B1 M 0* sinω 0t
Unit current i
= cos ω0t
Magnetization M0*
(
ξ s = ξ s0 exp(iω 0t)exp −t/T2* )
After nuclei are excited by a 90 degree pulse, the vector of magnetiza-
tion M0 on the xy plane begins to decay exponentially in amplitude by T2*
relaxation. However, if a 180 degree pulse follows immediately, the phases
Principles of Magnetic Resonance Imaging ◾ 127
T2/2 T2/2
ξs0
ξSE
T2* T2* t
T2
Echo time Te
are refocused to be coherent and the nuclei generate an echo (Hahn, 1950).
This is the spin echo (SE) signal, and the 180 degree pulse is often called a
time-reversing pulse, because it refocuses the phases of nuclear precessions
as if time were reversed (Figure 5.3). The spin echo signal amplitude ξSE is
affected by T1 relaxation as a matter of course during the repetition time
Tr and the T2 relaxation during the echo time Te. ξSE is expressed as ξSE =
ξs0{1 − exp (−Tr/T1)}exp (−Te/T2). The SE signal ξs can be hence expressed as
(
ξ s = ξ s0 {1− exp(−Tr /T1 )}exp(−Te /T2 )exp(iω 0t)exp − t − Te /T2* )
The SE signal ξs has a feature where the amplitude is symmetrical in
regard to the echo time Te. Therefore, if the excitation pulse and refocus-
ing pulse are θ1 and 2θ2 instead of 90 and 180 degree, respectively, then the
signal ξs can be expressed as
( )
ξ s =ξ s0 {1− exp(−Tr /T1 )}exp(−Te /T2 )exp(iω 0t)exp − t − Te /T2* sinθ1sin 2 θ 2
carrying currents in opposite directions. The field from one coil opposes
the field from the other, creating a net field that serves as a gradient field.
Gradient field can be applied in each of the three orthogonal directions
(x, y, and z). A gradient coil of any direction changes the magnetic flux
density of the z direction. When a gradient coil of z direction is applied,
the magnetic flux density in the z direction is the static magnetic field B0
itself at the center of the magnet, but it is higher at a position apart from
the center in the positive z direction. When a gradient coil of x direction is
applied, the magnetic flux density in the z direction is higher than B0 at a
position apart from the center in the positive x direction.
Slice selection is performed by applying a gradient magnetic field and an
excitation rf pulse simultaneously. Also, the recognition of nuclear distribu-
tion in the second and third directions is performed by employing gradient
coils. The spin-warp imaging technique is the traditional term for such popu-
lar methods of MRI as field echo (FE) imaging, spin echo (SE) imaging, and
echo-planar imaging (EPI). The main feature of spin-warp imaging technique
is the process of detecting the location of a spinning nucleus in the phase of
precession by varying a linear, gradient magnetic field (Edelstein et al., 1980).
0
z
Slice thickness 5 mm
FE
Gx
–Gxd ts
FIGURE 5.5 Reversing the gradient magnetic field and a field echo (FE) signal.
Principles of Magnetic Resonance Imaging ◾ 131
FIGURE 5.6 Echo signal acquired with the reading-out gradient field G x.
(
FIGURE 5.8 Fourier-transformed absolute value Av = Re + I m
2 2
)1/2 of the echo.
are frequency in hertz. The curve of the absolute value Av is similar to the
envelope shape of the real part Re curve.
Gz
Gx
Gy
Frequency f [Hz]
Te/2 Te/2
Gz
Gx
Gy
5.3.4 k Space
By varying the amplitude of the gradient field in the phase-encoding
direction, nuclei are labeled according to their phases of precessions. The
acquired data are then put in successive order on a plane traditionally
called k space (Ljunggren, 1983). The echo signals put on the k space are
those demodulated and subtracted by the primary frequency ω 0 = 2πν0.
Obviously, each signal data appear to have waves containing different fre-
quencies. Those waves are specific to the horizontal x axis distribution of
nuclei and generated by the reading-out gradient field Gx.
The waves in the vertical direction of the k space are also specific to the
vertical y axis distribution of nuclei and generated by varying the ampli-
tude of the phase-encoding gradient field G y. When a quadrature demod-
ulation is employed, each data has a real part and imaginary part. For
example, if the number of data points of a signal is 512, then 256 are real
data points and the remaining 256 are imaginary data points. Hence, when
the number of phase-encoding steps is 256, the vertical phase-encoded
data have 256 real data points and imaginary data points. Therefore, on
the k space, there are a real data set of 256 × 256 points and an imaginary
data set of 256 × 256 points. Figure 5.12 shows an example of a real data
set of 256 × 256.
134 ◾ Biomagnetics
the k space. Each horizontal line is a line graph consisting of 256 points,
while the vertical array is not Fourier transformed yet.
A two-dimensional nuclear distribution of an orange is shown in
Figure 5.15. This was restored by a two-dimensional Fourier transforma-
tion of the horizontal and vertical wave lines on the k space. The nuclear
distribution is displayed in absolute values. The matrix size is 256 × 256.
The unit of the horizontal axis is frequency f in hertz and corresponds
linearly to the distance l in meters from the central point. The unit of the
vertical axis is phase θ in rad/step and also corresponds linearly to the
distance l in meters from the point. If the intensity diagram is gray scaled
and displayed on a CRT, then we obtain an ordinary MR image.
5.3.5 Image Contrast
By applying an excitation rf pulse, the magnetization vector is given
a component perpendicular to the z axis and the component decays in
amplitude by T2* relaxation. When a FE image sequence is used, and if
both the repetition time Tr and echo time Te are long, then the resulting FE
image is a T2* -weighted image affected by T2* relaxation. The T2* relaxation
of the brain tissue often reflects the brain activity, and T2* -weighted images
are important in the field of fMRI.
136 ◾ Biomagnetics
(a) (b)
and offered by Siemens Japan K.K. The MRI scanner was a Siemens
Prisma 3Tesla machine. The images are with a slice thickness of 4 mm. The
total acquisition time was 63 sec. The images were obtained by GRAPPA,
standing for generalized autocalibrating partially parallel acquisition, a
Siemens-specific technique.
5.4.3 Functional Neuroimaging
Functional neuroimaging (fMRI) is one of the most important MRI tech-
niques developed recent years and visualizes the active regions of the brain
noninvasively. Functional imaging uses FE imaging sequences of long echo
time Te as 50 msec and reflects T2* characteristics of the tissues. Electrons of
Fe2+ of deoxy-hemoglobin (Deoxy-Hb) have much greater magnetic momen-
tum than those of proton nuclei, and proton resonance spectra of deoxy-
hemoglobin (Deoxy-Hb) solution exhibit broad peaks caused by the effects
of paramagnetic features of Deoxy-Hb. Also, the water resonance frequency
is broadened by the effects of the gradient magnetic field due to the Fe2+ elec-
trons. Therefore, T2*-weighted images of water around Deoxy-Hb generate
few signals. In the living brain, the gradient magnetic fields caused by the Fe2+
electrons strongly influence not only the water in the blood vessels but also,
to a much greater extent, the outer neural tissues of the water volume of the
blood vessels (Ogawa and Lee, 1990; Ogawa et al., 1990; Turner et al., 1991).
When a region in the brain is active, the blood flow increases in the
region, oxy-hemoglobin (Oxy-Hb) increases, and Deoxy-Hb decreases in
relative quantity. By the blood oxygenation level dependent effects, the sig-
nal intensities from the active region in the obtained T2* -weighted images
increases (Bandettini et al., 1992; Frahm et al., 1992). Also, fMRI reflects
internal verbal generations and recalling of sensory or motor stimulation.
(LeBihan et al., 1993; McCauthy et al., 1993; Rueckert et al., 1994).
Figure 5.19 shows a functional neuroimaging routine. A simultaneous
time-course analysis for multiple regions of interest (ROIs) is carried out
in routine. The image was also offered by Siemens Japan K.K.
5.5 CONCLUDING REMARKS
In the last century, MRI scanners were classified simply by their field
strength, namely, low-field (0.5 T and below), mid-field (1.0 T), and high-
field (1.5 T and higher) scanners. Continued progress in high-performance
gradient coil technologies have made high-quality and fast imaging a
reality. Techniques have been diversified dramatically. MRI scanners are
classified in a more sophisticated way, i.e., productivity in scanning real-
ity and real-time post-processing procedures and friendliness in design
concepts for clinical use. In the field of basic technology, it is apparent that
MRI is not simply a method of obtaining radiological images, but it can
also provide a great deal of physiological information. Some MRI tech-
niques will be developed for noninvasive mapping of physical, chemical,
and physiological properties such as temperature, mechanical elasticity,
pressure, electrical conductivity, permittivity, and metabolite concentra-
tion. Development will be continued in this century.
REFERENCES
Bandettini, P. A., Wong, E. C., Hinks, R. S., Tikofsky, R. S., and Hyde, J. S. 1992.
Time course EPI of human brain function during task activation. Magnetic
Resonance Medicine 25: 390.
Bloch, F., Hansen, W. W., and Packard, M. 1946. The nuclear induction experi-
ment. Physical Review 70: 474.
Edelman, R. R., Siewert, B., Darby, D. G., Thangaraj, V., Nobre, A. C., Mesulam,
M. M., and Warach, S. 1994. Qualitative mapping of cerebral blood flow and
functional localization with echo-planar MR imaging and signal targeting
with alternating radiofrequency. Radiology 192: 513.
Edelstein, W. A., Hutchison, J. M. S., Johnson, G., and Redpath, T. 1980. Spin warp
NMR imaging and applications to whole body imaging. Physics in Medicine
and Biology 25: 751.
Ernst, R. R., and Anderson, W. A. 1966. Application of Fourier transform spec-
troscopy to magnetic resonance. Review of Scientific Instruments 37: 93.
Frahm, J., Bruhn, H., Merboldt, K. D., and Haenicke, W. 1992. Dynamic MR imag-
ing of human brain oxygenation during rest and photic stimulation. Journal
of Magnetic Resonance Imaging 2: 501.
Hahn, E. L. 1950. Spin echoes. Physical Review 80: 580.
Iriguchi, N., and Hasegawa, J. 1993. Carbon-13 magnetic resonance imaging of a
human arm. Magnetic Resonance Imaging 11: 269.
Principles of Magnetic Resonance Imaging ◾ 143
Kohno, K., Hoehn-Berlage, M., Mies, G., Back, T., and Hossmann, K. A. 1995.
Relationship between diffusion-weighted MR images, cerebral blood flow,
and energy state in experimental brain infarction. Magnetic Resonance
Imaging 13: 73.
Lauterbur, P. C. 1973. Image formation by induced local interactions: Examples
employing nuclear magnetic resonance. Nature 242: 190.
LeBihan D., and Turner, R. 1992. Diffusion and perfusion. In Magnetic Resonance
Imaging. Mosby Year Books, St. Louis.
LeBihan, D., Turner, R., Zeffiro, T. A., Cuenod, C. A., Jezzard, P., and Bonnerot,
V. 1993. Activation of human primary visual cortex during visual recall: An
MRI study. Proceedings of the National Academy of Sciences of the United
States of America 90: 1802.
Ljunggren, S. 1983. A simple graphical representation of Fourier-based imaging
method. Journal of Magnetic Resonance 54: 338.
McCauthy, G., Blamire, A. M., Rothman, D. L., Gruetter, R., and Schulman, R. G.
1993. Echo-planar MRI studies of frontal cortex activation during word gen-
eration in humans. Proceedings of the National Academy of Sciences of the
United States of America 90: 4952.
Mintrovich, J., Mosley, M. E., Chileuitt, L., Shimizu, H., Cohen, Y., and Weinstein,
R. R. 1991. Comparison of diffusion- and T2-weighted MRI for early detec-
tion of cerebral ischemia and reperfusion in rats. Magnetic Resonance
Medicine 18: 39.
Mosley, M. E., Cohen, Y., Mintorovich, J., Chileuitt, L., Shimizu, H., Kucharczyk,
J., Wendland, F. M., and Weinstein, P. R. 1990. Early detection of regional
cerebral ischemia in cats: Comparison of diffusion- and T2-weighted MRI
and spectroscopy. Magnetic Resonance in Medicine 14: 330.
Ogawa, S., and Lee, T-M. 1990. Magnetic resonance imaging of blood vessels
at high field: In vivo and in in vitro measurements and image simulation.
Magnetic Resonance Medicine 16: 9.
Ogawa, S., Lee, T-M., Nayak, A. S., and Glynn, P. 1990. Oxygenation-sensitive con-
trast in magnetic resonance image of rodent brain at high magnetic field.
Magnetic Resonance Medicine 14: 68.
Pennock, J. M., Cowan, F. W., Schweiso, J. E., Oatridge, A., Rutherfard, M. A.,
Pubowitz, L. M. S., and Bydder, G. M. 1994. Clinical role of diffusion-
weighted imaging: Neonatal studies. Magnetic Resonance Materials in
Physics, Biology and Medicine 2: 273.
Purcell, E. M., Torroy, H. C., and Pound, R. V. 1946. Resonance absorption by
nuclear magnetic moments in a solid. Physical Review 69: 37.
Rueckert, L., Appollonio, I., Grafmaii, J., Jezzard, P., Johnson, Jr., R., LeBihan, D.,
and Turner, R. 1994. MRI functional activation of the left frontal cortex dur-
ing covert word production. Journal of Neuroimaging 4: 67.
Stejskal, E. O., and Tanner, J. E. 1965. Spin diffusion measurements: Spin echoes
in the presence of time-dependent field gradient. Journal of Chemical Phys.
42: 288.
Stejskal, E. O. 1965. Use of spin echo in a pulsed magnetic-field gradient to study
anisotropic, restricted diffusion and flow. Journal of Chemical Physics 43: 3597.
144 ◾ Biomagnetics
CONTENTS
6.1 Introduction 146
6.1.1 Advantages of Magnetic Resonance Imaging
in Impedance Imaging 146
6.1.2 Advantages of MRI in Electric Current Imaging 147
6.2 Influence of Sample Impedance on MRI 148
6.2.1 Electric Properties of Biological Tissues 148
6.2.1.1 Permittivity and Conductivity 148
6.2.1.2 Frequency Characteristics and Anisotropy 150
6.2.1.3 Human Models for Numerical Simulations
of Electromagnetic Fields 153
6.2.2 RF Distribution in a Cylindrical Sample 154
6.2.3 RF Distribution in the Human Head 155
6.2.3.1 Finite Difference Time Domain Method 155
6.2.3.2 RF Distribution in the Human Head
at Varied Frequencies 157
6.3 Magnetic Resonance Imaging of Impedance 159
6.3.1 Large Flip Angle Method 159
6.3.2 Use of an External AC Field 159
6.3.3 Conductivity Imaging Using Diffusion MRI 161
145
146 ◾ Biomagnetics
6.1 INTRODUCTION
6.1.1 Advantages of Magnetic Resonance Imaging
in Impedance Imaging
Analyses of electromagnetic fields play an important role in understanding
electromagnetic phenomena in living bodies. Worldwide, several groups
develop and distribute standard models of the human body,1,2 and these
models benefit research into electromagnetic fields. However, the present
methodology for developing these standard models imposes certain limi-
tations on the scope and accuracy of these analyses. The models consist
of multiple tissues or organs segmented from cross-sectional images of
the body. Each tissue has specific permittivity and conductivity values.
Each organ, such as the brain, is modeled as a uniform medium. The elec-
tric properties of actual organs and tissues, however, are not necessarily
uniform because of a variety of substructures. Moreover, these standard
models do not replicate the anatomy of an individual subject and are not
applicable to subject-specific analyses. One solution to the aforemen-
tioned issues will be the development of new methodology for impedance
Prospects for MRI of Impedance and Electric Currents ◾ 147
v
λ= (6.2)
f
j = σE (6.3)
iσ
ε* = ε − (6.4)
ω
where ω is the angular frequency. In some cases, the sign of the imaginary
part may be positive instead of negative. When analyzing the distribution
of an RF wave, the use of complex permittivity enables us to write the wave
equation in a simple form.
ε0 − ε ∞
ε *(ω) = ε ∞ + (6.5)
1+ iωτ
where τ is the relaxation time and where ε0 and ε∞ indicate the respec-
tive permittivities at sufficiently low and high frequencies (compared with
1/τ). With ω ranging from 0 to ∞, a plot of Equation 6.5 on a complex
plane produces a semicircle, which is called a Cole-Cole plot.5 There are
multiple sources of dielectric polarization in a tissue, such as charges on
cell membranes and polar molecules (including water), and these polar-
izations have different relaxation times. With four relaxation times, an
Prospects for MRI of Impedance and Electric Currents ◾ 151
extended version of Equation 6.5 is used for modeling the tissue’s complex
permittivity:
∆ε
4
ε *(ω) = ε ∞ + ∑ 1+ (iωτ
m=1
)
m
m
1−αm +
σi
jωε0
(6.6)
where Δεm indicates the drop in permittivity for a frequency range across
ω = 1/τm, which corresponds to each relaxation time τm; σi indicates the
ionic conductivity. The above empirically defined equation is well fitted
to the measured complex permittivities in a broad frequency range, from
10 Hz to 100 GHz.6,7 Figure 6.1 shows the frequency characteristics of gray
108
106
104
εr
102
102
101
σ (S/m)
100
10–1
methods that use cubic cells, such as the finite difference time domain
method, the impedance method, and the scalar-potential finite differ-
ence method. The finite element method is also available.
2
l=
µσω (6.7)
1.2 π
0.0 –π
(a) (d)
1.2 π
0.0 –π
(b) (e)
1.2 π
0.0 –π
(c) (f)
FIGURE 6.2 Magnitude |ξ| ([a] a = 30 mm, water; [b] a = 50 mm, water; [c] a =
50 mm, saline) and phase angle arg(ξ) ([d] a = 30 mm, water; [e] a = 50 mm, water;
[f] a = 50 mm, saline) of magnetic resonance signals obtained from cylindrical
phantoms with different radii and electric properties. The signals are normalized
by the values at the center of the phantom. (From Sekino, M. et al., Journal of
Applied Physics 97: 10R303, 2005.)
z
1
1 1 Ez i, j, k +
Hy i + , j, k + 2 1 1
2 2 Hx i, j + , k+
2 2
1
Ey i, j + ,k
i, j, k 2
y
x 1 1 1
Ex i + , j, k Hz i + , j + , k
2 2 2
FIGURE 6.3 Computational cell for the FDTD method. The electromagnetic
field components are assigned on the edges or faces of each cell. (From Sekino, M.
et al., Journal of Applied Physics 103: 07A318, 2008.)
for each time step can be calculated using the Yee algorithm12,13 when an
RF coil is the wave source. For example, the z component of the electro-
magnetic field is calculated as follows:
∆t
ε(i, j,k +1/2) 1
+
σ(i, j,k +1/2)∆t ∆x
1+
2ε(i, j,k +1/2)
∆t
ε(i, j,k +1/2) 1
−
σ(i, j,k +1/2)∆t ∆y
1+
2ε(i, j,k +1/2)
⎧ 1 ⎫
× ⎨ H xn−1/2 ⎛ i, j + ,k + ⎞ − H xn−1/2 ⎛ i, j − ,k + ⎞ ⎬
1 1 1
⎩ ⎝ 2 2 ⎠ ⎝ 2 2⎠ ⎭ (6.8)
Prospects for MRI of Impedance and Electric Currents ◾ 157
H zn+1/2 ⎛ i + , j + ,k ⎞ = H zn−1/2 ⎛ i + , j + ,k ⎞
1 1 1 1
⎝ 2 2 ⎠ ⎝ 2 2 ⎠
where E zn and H zn are, respectively, the electric and magnetic field compo-
nents at the nth time step, i, j, and k are the index numbers of the com-
putational cell, and Δt, Δx, and Δy are the length of a time step and the
lengths of the computational cell in the x and y directions, respectively.
At the outer boundaries of a computational model, absorbing boundary
conditions are applied to eliminate the influence of the boundaries. There
is a variety of formulations for the boundary conditions.
FIGURE 6.4 Distribution of (a) an RF electric field and (b) an RF magnetic field
in a horizontal section of the human head with different magnetic resonance
frequencies. The increase in frequency causes the occurrence of hot spots that
exhibit high RF absorption. (From Sekino, M. et al., Journal of Applied Physics
103: 07A318, 2008.)
F2 F2 F2
(cm) (cm) (cm)
3.5 3.0 2.5 2.0 1.5 1.0 0.5 3.5 3.0 2.5 2.0 1.5 1.0 0.5 3.5 3.0 2.5 2.0 1.5 1.0 0.5
F1 (cm) F1 (cm) F1 (cm)
FIGURE 6.5 Impedance imaging of the mouse head using the large flip angle
method. The images were obtained with different excitation flip angles and were
evaluated in the cerebrospinal fluid. The flips angles were (a) 160°, (b) 180°, and
(c) 200°. (From Ueno, S., and Iriguchi, N., Journal of Applied Physics 83: 6450–
6452, 1998.)
Resistive
material
Conducting
material
The excited slice
position by
prepulse (90°)
Resistive
material
Conducting
material
Slice position
on RF pulse
(90°)
Resistive
material
Conducting
material
Si (b)
= f fast,i exp(−bDfast,i )+ f slow,i exp(−bDslow,i ) (6.10)
Si (0)
where i is the index number of the MPG, Si(b) is the signal intensity of the
diffusion-weighted images, b is the b factor of the MPG, Dfast,i and Dslow,i
are the fast and slow diffusion components, respectively, and ffast,i and fslow,i
are the respective fractions of the fast and slow components.
The conductivity of the extracellular fluid is obtained from the fast dif-
fusion coefficient based on the proportionality between conductivity and
the diffusion coefficient. Finally, the effective conductivity of the tissue is
estimated from the extracellular fluid’s conductivity and volume fraction.
The effective conductivity σi for direction i is given by17
2 f fast,i × (8.1×108 )
σi = Dfast,i (6.11)
3 − f fast,i
6.3.3.2 Experimental Results
Figure 6.7 shows the relationships between the b factor and the signal
intensities of diffusion-weighted images of the human brain.18 The regions
of interest were located on the putamen (gray matter), the posterior limb
of the internal capsule (white matter), and the genu of the corpus callo-
sum (white matter). The plots indicate the signal intensities measured with
the following MPG directions: anterior-posterior, right-left, and superior-
inferior. The signals gradually decreased with an increase in the b factor.
The putamen did not demonstrate clear anisotropy in signal attenuation.
In the internal capsule, the application of the MPG in the superior-inferior
direction caused the most rapid signal attenuation. In the corpus callo-
sum, the application of the MPG in the right-left direction caused the
most rapid signal attenuation.
Prospects for MRI of Impedance and Electric Currents ◾ 163
–1
ln(signal)
–2
–3
–4 Anterior-posterior
Right-left
Superior-inferior
–1
–2
ln(signal)
–3
–4 Anterior-posterior
Right-left
Superior-inferior
–1
–2
ln(signal)
–3
–4 Anterior-posterior
Right-left
Superior-inferior
FIGURE 6.7 Relationships between the b factor and the signal intensities of
diffusion-weighted images in the putamen. The three plots were obtained with
different MPG directions. (a) The putamen did not exhibit clear anisotropy in the
signal attenuations. (b) Signal attenuations in the internal capsule. The applica-
tion of the MPG in the superior-inferior direction caused the most rapid signal
attenuation. (c) Signal attenuations in the corpus callosum. The application of the
MPG in the right-left direction caused the most rapid signal attenuation. (From
Sekino, M. et al., IEEE Transactions on Magnetics 41: 4203–4205, 2005.)
164 ◾ Biomagnetics
2.6 2.6
0.0 0.0
(a) S/m (d) S/m
2.6 1.0
0.0 0.0
(b) S/m (e)
2.6
0.0
(c) S/m
FIGURE 6.8 Conductivity imaging of the human brain using diffusion MRI.
(a through c) Images of the estimated conductivities in the anterior-posterior,
right-left, and superior-inferior directions, respectively. (d, e) Images of the MC
and AI. (From Sekino, M. et al., IEEE Transactions on Magnetics 41: 4203–4205,
2005.)
Prospects for MRI of Impedance and Electric Currents ◾ 165
the brain’s principal neuronal fiber tracts are well established from previ-
ous anatomical and histological studies. The corpus callosum connects
the right hemisphere and the left hemisphere with several neuronal fibers
that run in the right-left direction. The posterior limb of the internal cap-
sule lies on the pyramidal tract, which mainly consists of neuronal fibers
running in the superior-inferior direction. The diffusion of water mol-
ecules in the extracellular fluid is disturbed by the cell membranes, and
diffusibility is higher in the direction of neuronal fibers than in the other
directions. The high AI values in the internal capsule and corpus callosum
are attributable to the anatomical structures of these regions.
The inhomogeneity and anisotropy of the conductivity in each tissue
is not normally considered in conductivity models for current source
estimations of electroencephalography and magnetoencephalography.
However, the results of current source estimations depend on the spatial
distribution of conductivity in the models. The results of conductivity
imaging show significant inhomogeneity and anisotropy in the white mat-
ter. Therefore, the use of an inhomogeneous and anisotropic conductivity
model is desirable, particularly for estimating current sources in the deep
regions of the brain.
Conventional hardware
RF receiver
Sample
RF transmitter (B)
y Electrode
RF transmitter (A) +x
RF receiver –x
RF transmitter (B)
(b) τ
of the magnetic resonance signals. In the next step, B1 turns toward the
y′ direction, and θ π/2 = B y′ γ τ is measured. On the basis of the wave equa-
tion of the electromagnetic fields in the sample, the following equation is
derived to estimate permittivity εr and conductivity σ:
∇ 2 (θ0 − iθ π/2 )
ω 02ε0εr µ 0 − iω 0µ 0σ = − (6.12)
θ0 − iθ π/2
6.3.4.2 Experimental Results
Preliminary results of phantom experiments have been reported in a
previous paper.21 An acrylic tube was filled with a solution comprised of
saline (30%) and ethanol (70%). The permittivity and conductivity of the
solution were approximately 40 and 0.14 S/m, respectively. A pair of plati-
num electrodes was attached to both ends of the sample. Using a coaxial
cable, the electrodes were connected to an RF transmitter. Signals were
detected using a birdcage-type RF coil.
As shown in Figure 6.10, images of magnetization angles were obtained
from magnetic resonance signals, and the sample’s permittivity and con-
ductivity were estimated. The lower left part of the sample exhibited a rela-
tively large error in the estimated permittivity and conductivity. The lead
wire caused inhomogeneity in the RF electromagnetic field, which resulted
in an error around the wire. At the center of the sample, we obtained an
estimated relative permittivity of 40 and an estimated conductivity of
2.0 S/m, which are reasonable results. In a part of the image, however, the
results showed errors that are mainly attributable to inhomogeneity in the
field generated by the RF coil.
This methodology has potential advantages over conventional imped-
ance imaging techniques because of its high spatial resolution and its
capacity to measure permittivity. In addition, this method is easily appli-
cable to both biological tissues and human subjects.
θ0 θπ/2
–π π –π π
εr σ
Jk V *
ρk+1 = ρk
J * V k (6.13)
ϕ = γBz τ (6.14)
To map the generated magnetic field, two phase angle images are
acquired with and without the application of an electric current, and the
two images are subtracted.25 Spontaneously generated magnetic fields,
such as neuronal fields, generally have time-varying currents. In this case,
Equation 6.14 is modified as
τ
φ=γ
∫ B (t)dt (6.15)
0
z
The above discussion indicates that only the z component of the mag-
netic field can be measured using MRI. The process of rotating the sample
inside the MRI scanner such that the sample is oriented along the other
two orthogonal directions and then measuring the corresponding mag-
netic field maps provides the three components of the magnetic field vec-
tor. Taking the rotation of the magnetic field gives the electric current
distribution:
1
j= ∇ × B (6.16)
µ0
170 ◾ Biomagnetics
i kx2 + k y2
j z′ = − bx′ (6.17)
µ0 ky
where j z′ and bx′ are the Fourier transforms of the z component of the
current density and the x component of the magnetic field, respectively.
According to Equation 6.17, estimating one component of the current
density requires one orthogonal component of the magnetic field to be
mapped.
When the generated magnetic field is strongly inhomogeneous, the field
causes the loss of signal coherence in a voxel, which results in a decrease in
signal magnitude. This effect may occur in close proximity to the current
source. While the measurement based on the phase angle has relatively
high sensitivity to the magnetic field, the image analyses include a some-
what complicated process for phase unwrapping. The measurement using
the signal magnitude allows for simple post processing.
When the intensity of the magnetic field is relatively high, the influ-
ence of the magnetic field can also be observed as a shift in the Larmor
frequency. Numerical simulations based on the Bloch equation provide
a quantitative relationship between the generated field and the signal
intensity.31
µ0 r − r′
B(r) = Q(r ′)×
r − r ′ (6.18)
3
4π
Prospects for MRI of Impedance and Electric Currents ◾ 171
where B(r) is the neuronal magnetic field at location r, Q(r′) is the cur-
rent dipole at location r′, and μ0 is the permittivity of free space. Here, we
assume that the main static magnetic field, B0, is applied in +z direction
and that the current dipole is located at r′ = (0, 0, 0), pointing in the +x
direction. When B(r) is much smaller than B0, the magnitude of the total
magnetic field can be approximated as33
µ0 Qy
B 0 + B(r) ≈ B 0 + Bz (r) = B 0 + (6.19)
4π (x + y 2 + z 2 )3/2
2
where Bz(r) is the z component of B(r) and Q is the strength of the current
dipole. The frequency shift caused by the neuronal magnetic field is
µ0γ Qy
∆ω = γ ( B 0 − B(r) − B 0 ) ≈ (6.20)
4π (x + y 2 + z 2 )3/2
2
For an image voxel centered at (x0, y0, z0), the following equation gives
the signal intensity normalized by the intensity without a neuronal mag-
netic field:
x0 +hx /2 y0 +h y /2 z0 +hz /2
∫ ∫ ∫
1
S= dx dy dz exp(iTE∆ω) (6.21)
hx hy hz x0 −hx /2 y0 −h y /2 z0 −hz /2
where hx, hy, and hz are the dimensions of the voxel in the x, y, and z direc-
tions, respectively, and TE is the echo time for MRI acquisition. Examples
of calculating the above signal intensity are reported elsewhere.33
N
σB = (6.22)
SγTE
where N is the intensity of noise in the MRI signal, S is the signal intensity,
and TE is the echo time.34 The magnetic field can be detected when inten-
sity β is higher than uncertainty σB. Therefore, σB gives the theoretical
sensitivity for magnetic fields.
Magnetization M0, which is induced in a sample by the main static field,
B0, is given by
where TR is the repetition time, T1 and T2 are the sample’s relaxation times,
and θ is the flip angle. B1 is the RF field intensity that the unit current flow-
ing through the RF receiver coil produces at the voxel.
In an MRI consisting of n × n voxels, noise N per voxel is
∆N = n 4kBTs∆fR (6.25)
where Δf is the spectral width of the receiver circuit and R is the effective
resistance in the receiver circuit. The resistance is partly caused by the
conductors in the receiver coil and partly caused by the sample. Under
typical MRI conditions in the human head, the resistance caused by the
sample is dominant. Produced by the sample, equivalent resistance R in
the receiver circuit is related to the absorption, which the unit current
flowing in the receiver coil at the Larmor frequency causes in the sample;
namely,
Prospects for MRI of Impedance and Electric Currents ◾ 173
| j|2
R=2
∫ σ
dVs (6.26)
(a) (b)
Birdcage
coil I11 I0 I1 I2 I3
I0 – I11 I1 – I0 I2 – I1 I3 – I2
I0 I1 I2 I3
FIGURE 6.11 Evaluation of the sensitivity for detecting weak magnetic fields
generated in the human brain. Numerical analyses were conducted on a realistic
human head model and a birdcage-type receiver coil. (a) Coronal and (b) sagit-
tal slices of the model. (c) Three-dimensional view of the model and the receiver
coil. (d) A circuit schematic of the receiver coil. (From Hatada, T. et al., Journal of
Applied Physics 97: 10E109, 2005.)
174 ◾ Biomagnetics
6.4.3 Animal Studies
6.4.3.1 Brain Slices
In addition to the magnetic field arising from neuronal electric activi-
ties, associated changes in blood oxygenation cause considerable fluctua-
tion in the magnetic field and result in baseline signal fluctuations. This
Prospects for MRI of Impedance and Electric Currents ◾ 175
Multielectrode array
Space filler 8 × 8 ch.
150 µm
Sinker
Rat brain slice
150 µm
(a)
Space filler
Sinker
(b)
FIGURE 6.12 Experimental setup for detecting magnetic fields arising from an
extracted brain slice. A multielectrode array measures and stimulates the living
brain slice. The perfusion of artificial cerebrospinal fluid washes the blood from
the brain slice. (a) Construction of perfusion chamber and perfusing system.
(b) Position relation in perfusion chamber. (From Kim, D. et al., Proceedings of
the Annual International Conference of the IEEE Engineering in Medicine and
Biology Society 1370–1373, 2013.)
176 ◾ Biomagnetics
0.025 Left S1
0.02
Signal intensity
0.015
Stimulation
0.01
0.005
Control
0 50 100 150 200 250
(a) Time after stimulation (msec)
0.025
Stimulation
0.02
Signal intensity
0.015 Control
Right S1
0.01
0.005
FIGURE 6.13 Time courses of the signal intensity in the (a) left and (b) right
somatosensory areas in the rat brain after electric stimulation of the left sciatic
nerve. (From Sekino, M. et al., IEEE Transactions on Magnetics 45: 4841–4844,
2009.)
stimulation. In the right ROI, the signal intensity with stimulation was
higher than the intensity without stimulation at most time points. This
increase in the signal intensity’s baseline was caused by the blood oxygen-
ation level-dependent effect associated with increased blood flow in the
right somatosensory cortex. A transient decrease in the signal intensity
was found at 0 to 30 msec after stimulation. This result is consistent with
a local change in the magnetic field resulting from neuronal activities.
The images were compared at adjacent time points to investigate temporal
changes, as shown in Figure 6.14. A significant difference was observed
between 30 to 60 msec and 60 to 90 msec in the right somatosensory area,
which suggests that these effects arose temporarily from neuronal mag-
netic fields. However, both the somatosensory cortex and other areas of
178 ◾ Biomagnetics
0.05
240–270 msec P value
270–300 msec
FIGURE 6.14 Detection of magnetic fields arising from neuronal activities. The
signal intensities from the images obtained at adjacent time points were com-
pared after electric stimulation. The time “0–30 msec, 30–60 msec” means a
comparison between the image excited at 0 msec (and acquired at 30 msec) and
the image excited at 30 msec (and acquired at 60 msec). Color indicates the level
of statistical significance (P value) in a Student’s t-test. (From Sekino, M. et al.,
IEEE Transactions on Magnetics 45: 4841–4844, 2009.)
6.4.4 Human Studies
Several papers have reported human studies seeking to detect neuronal
currents in the brain or the median nerve.42–44 A pioneering work in this
topic used gradient magnetic fields with different polarities to detect neu-
ronal fields and to eliminate signal variations resulting from other effects.42
Measurements were conducted using a 1.5-T MRI system. In addition to
Prospects for MRI of Impedance and Electric Currents ◾ 179
neuronal magnetic fields, the blood oxygenation level and blood volume
affect images. As shown in Table 6.2, functional images were obtained
from gradient fields with different polarities, and the images were edited
to eliminate intensity changes resulting from causes other than neuronal
currents normal to the static magnetic field. Identical effects on Ip and In
are produced by changes in blood’s magnetic susceptibility. Signal inten-
sity changes resulting from causes other than neuronal currents are elimi-
nated in subtracted images of Ip and In. Neuronal currents cause different
effects on Ip and In, and signal intensity changes are observed in subtracted
images (Ip – In and In – Ip). The component of currents normal to the read-
out gradient causes symmetrical signal intensity changes about the com-
ponent’s axis. The parallel component produces asymmetrical intensity
changes. Areas with symmetrical intensity changes were extracted to
avoid misrepresentation by artifacts.
A double-subtracted image was produced for the activated area of the
human brain during repeated tapping of the right hand’s middle fin-
ger and thumb. Figure 6.15 shows a neuronal current map based on the
double-subtracted image. The y component (normal to the static magnetic
field and the readout gradient) of the currents in the activated sensory
area was clearly observed. The currents in the motor area, however, were
not detected because the y component of the currents in the motor area
was too small.
I
Ipr Rest p
Ipa Activated p
Inr Rest n
Ina Activated n
II
Ipa − Ipr = Ip → fMRI
Ina − Inr = In → fMRI
III
In − Ip → Current distribution image
Ip − In → Current distribution image
Source: Kamei, H. et al., IEEE Transactions on Magnetics 35: 4109–4111, 1999.
180 ◾ Biomagnetics
REFERENCES
1. Nagaoka, T., Watanabe, S., Sakurai, K., Kunieda, E., Watanabe, S., Taki, M.,
and Yamanaka, Y. 2004. Development of realistic high-resolution whole-
body voxel models of Japanese adult males and females of average height
and weight, and application of models to radio-frequency electromagnetic-
field dosimetry. Physics in Medicine and Biology 49: 1–15.
2. Christ, A., Kainz, W., Hahn, E. G., Honegger, K., Zefferer, M., Neufeld, E.,
Rascher, W., Janka, R., Bautz, W., Chen, J., Kiefer, B., Schmitt, P., Hollenbach,
H. P., Shen, J., Oberle, M., Szczerba, D., Kam, A., Guag, J. W., and Kuster,
N. 2010. The virtual family—Development of surface-based anatomical
models of two adults and two children for dosimetric simulations. Physics
in Medicine and Biology 55: 23–38.
3. Metherall, P., Barber, D. C., Smallwood, R. H., and Brown, B. H. 1996.
Three-dimensional electrical impedance tomography. Nature 380: 509–512.
4. Hashemzadeh, P., Fhager, A., and Persson, M., 2006. Experimental investiga-
tion of an optimization approach to microwave tomography. Electromagnetic
Biology and Medicine 25: 1–12.
5. Cole, K. S., and Cole, R. H. 1941. Dispersion and absorption in dielectrics: I.
Alternating current characteristics. Journal of Chemical Physics 9: 341–351.
6. Gabriel, C., Gabriel, S., and Corthout, E. 1996. The dielectric properties of
biological tissues: I. Literature survey. Physics in Medicine and Biology 41:
2231–2249.
7. Gabriel, S., Lau, R. W., and Gabriel, C. 1996. The dielectric properties of bio-
logical tissues: II. Measurements in the frequency range 10 Hz to 20 GHz.
Physics in Medicine and Biology 41: 2251–2269.
8. Safety Committee of the Japan Society of Magnetic Resonance in Medicine.
2010. MR Safety—Principles, Standards and Clinical Concerns. Gakken
Medical Shujunsha, Tokyo.
9. Hoeltzell, P. B., and Dykes, R. W. 1979. Conductivity in the somatosensory
cortex of the cat—Evidence for cortical anisotropy. Brain Research 177: 61–82.
10. Nagaoka, T., Togashi, T., Saito, K., Takahashi, M., Ito, K., and Watanabe,
S. 2007. An anatomically realistic whole-body pregnant-woman model and
specific absorption rates for pregnant-woman exposure to electromagnetic
plane waves from 10 MHz to 2 GHz. Physics in Medicine and Biology 52:
6731–6745.
182 ◾ Biomagnetics
11. Sekino, M., Mihara, H., Iriguchi, N., and Ueno, S. 2005. Dielectric reso-
nance in magnetic resonance imaging: Signal inhomogeneities in samples
of high permittivity. Journal of Applied Physics 97: 10R303.
12. Yee, K. 1966. Numerical solution of initial boundary value problems involv-
ing Maxwell’s equations in isotropic media. IEEE Transactions on Antennas
and Propagation 14: 302–307.
13. Wada, H., Sekino, M., Ohsaki, H., Hisatsune, T., Ikehira, H., and Kiyoshi,
T. 2010. Prospect of high-field MRI. IEEE Transactions on Applied
Superconductivity 20: 115–122.
14. Sekino, M., Kim, D., and Ohsaki, H. 2008. FDTD simulations of RF electro-
magnetic fields and signal inhomogeneities in ultrahigh-field MRI systems.
Journal of Applied Physics 103: 07A318.
15. Ueno, S., and Iriguchi, N. 1998. Impedance magnetic resonance imaging:
A method for imaging of impedance distributions based on magnetic reso-
nance imaging. Journal of Applied Physics 83: 6450–6452.
16. Yukawa, Y., Iriguchi, N., and Ueno, S. 1999. Impedance magnetic resonance
imaging with external AC field added to main static field. IEEE Transactions
on Magnetics 35: 4121–4123.
17. Sekino, M., Yamaguchi, K., Iriguchi, N., and Ueno, S. 2003. Conductivity
tensor imaging of the brain using diffusion-weighted magnetic resonance
imaging. Journal of Applied Physics 93: 6730–6732.
18. Sekino, M., Inoue, Y., and Ueno, S. 2005. Magnetic resonance imaging of
electrical conductivity in the human brain. IEEE Transactions on Magnetics
41: 4203–4205.
19. Sekino, M., Ohsaki, H., Yamaguchi-Sekino, S., Iriguchi, N., and Ueno, S.
2009. Low-frequency conductivity tensor of rat brain tissues inferred from
diffusion MRI. Bioelectromagnetics 30: 489–499.
20. Tuch, D. S., Wedeen, V. J., Dale, A. M., George, J. S., and Belliveau, J. W.
2001. Conductivity tensor mapping of the human brain using diffusion
tensor MRI. Proceedings of the National Academy of Sciences of the United
States of America 98: 11,697–11,701.
21. Sekino, M., Tatara, S., and Ohsaki, H. 2008. Imaging of electric permittivity
and conductivity using MRI. IEEE Transactions on Magnetics 44: 4460–4463.
22. Katscher, U., Voigt, T., Findeklee, C., Vernickel, P., Nehrke, K., and Dössel,
O. 2009. Determination of electric conductivity and local SAR via B1 map-
ping. IEEE Transactions on Medical Imaging 28: 1365–1374.
23. Voigt, T., Katscher, U., and Doessel, O. 2011. Quantitative conductivity and
permittivity imaging of the human brain using electric properties tomog-
raphy. Magnetic Resonance in Medicine 66: 456–466.
24. van Lier, A. L. H. M. W., Brunner, D. O., Pruessmann, K. P., Klomp, D. W. J.,
Luijten, P. R,. Lagendijk, J. J. W., and van den Berg, C. A. T. 2012. B1+
phase mapping at 7T and its application for in vivo electrical conductivity
mapping. Magnetic Resonance in Medicine 67: 552–561.
25. Joy, M., Scott, G., and Henkelman, M. 1989. In vivo detection of applied
electric currents by magnetic resonance imaging. Magnetic Resonance
Imaging 7: 89–94.
Prospects for MRI of Impedance and Electric Currents ◾ 183
26. Khang, H. S., Lee, B. I., Oh, S. H., Woo, E. J., Lee, S. Y., Cho, M. H., Kwon, O.,
Yoon, J. R., and Seo, J. K. 2002. J-substitution algorithm in magnetic reso-
nance electrical impedance tomography (MREIT): Phantom experiments for
static resistivity images. IEEE Transactions on Medical Imaging 21: 695–702.
27. Lee, C. O., Jeon, K., Ahn, S., Kim, H. J., and Woo, E. J. 2011. Ramp-preserving
denoising for conductivity image reconstruction in magnetic resonance
electrical impedance tomography. IEEE Transactions on Biomedical Engi
neering 58: 2038–2050.
28. Oh, T. I., Jeong, W. C., McEwan, A., Park, H. M., Kim, H. J., Kwon, O. I.,
and Woo, E. J. 2013. Feasibility of magnetic resonance electrical impedance
tomography (MREIT) conductivity imaging to evaluate brain abscess lesion:
In vivo canine model. Journal of Magnetic Resonance Imaging 38: 189–197.
29. Seo, J. K., Yoon, J. R., Woo, E. J., and Kwon, O. 2003. Reconstruction of
conductivity and current density images using only one component of mag-
netic field measurements. IEEE Transactions on Biomedical Engineering 50:
1121–1124.
30. Ider, Y. Z. Onart, S., and Lionheart, W. R. 2003. Uniqueness and reconstruc-
tion in magnetic resonance-electrical impedance tomography (MR-EIT).
Physiological Measurement 24: 591–604.
31. Sekino, M., Matsumoto, T., Yamaguchi, K., Iriguchi, N., and Ueno, S. 2004.
A method for NMR imaging of a magnetic field generated by electric cur-
rent. IEEE Transactions on Magnetics 40: 2188–2190.
32. Demachi, K., Rybalko, S., and Fujita, M. 2008. Inverse analysis of the
current dipoles distribution in a human brain applied with the shifting-
aperture method. IEEE Transactions on Magnetics 44: 1426–1429.
33. Sekino, M., Ohsaki, H., Yamaguchi-Sekino, S., and Ueno, S. 2009. Toward
detection of transient changes in magnetic resonance signal intensity aris-
ing from neuronal electrical activities. IEEE Transactions on Magnetics 45:
4841–4844.
34. Scott, G. C., Joy, M. L. G., Armstrong, R. L., and Henkelman, R. M. 1992.
Sensitivity of magnetic-resonance current-density imaging. Journal of
Magnetic Resonance 97: 235–254.
35. Hatada, T., Sekino, M., and Ueno, S. 2005. FEM-based calculation of the
theoretical limit of sensitivity for detecting weak magnetic fields in the
human brain using magnetic resonance imaging. Journal of Applied Physics
97: 10E109.
36. Hatada, T., Sekino, M., and Ueno, S. 2004. Detection of weak magnetic
fields induced by electrical currents with MRI: Theoretical and practical
limits of sensitivity. Magnetic Resonance in Medical Sciences 3: 159–163.
37. Halpern-Manners, N. W., Bajaj, V. S., Teisseyre, T. Z., and Pines, A. 2010.
Magnetic resonance imaging of oscillating electrical currents. Proceedings
of the National Academy of Sciences of the United States of America 107:
8519–8524.
38. Sekino, M., Chin, Y., Takewa, T., Kim, D., and Someya, T. 2014. Submillimeter-
scale mapping of evoked magnetic fields in the rat brain for neuronal current
MRI. IEEE International Magnetics Conference.
184 ◾ Biomagnetics
39. Petridou, N., Plenz, D., Silva, A. C., Loew, M., Bodurka, J., and Bandettini,
P. A. 2006. Direct magnetic resonance detection of neuronal electrical
activity. Proceedings of the National Academy of Sciences of the United States
of America 103: 16,015–16,020.
40. Luo, Q., Lu, H., Lu, H., Senseman, D., Worsley, K., Yang, Y., and Gao, J. H.
2009. Physiologically evoked neuronal current MRI in a bloodless turtle
brain: Detectable or not? NeuroImage 47: 1268–1276.
41. Kim, D., Someya, T., and Sekino, M. 2013. Sensitivity of MRI for directly
detecting neuronal electrical activities in rat brain slices. Proceedings of the
Annual International Conference of the IEEE Engineering in Medicine and
Biology Society 1370–1373.
42. Kamei, H., Iramina, K., Yoshikawa, K., and Ueno, S. 1999. Neuronal cur-
rent distribution imaging using magnetic resonance. IEEE Transactions on
Magnetics 35: 4109–4111.
43. Xiong, J., Fox, P. T., and Gao, J. H. 2003. Directly mapping magnetic field
effects of neuronal activity by magnetic resonance imaging. Human Brain
Mapping 20: 41–49.
44. Truong, T. K., and Song, A. W. 2006. Finding neuroelectric activity under
magnetic-field oscillations (NAMO) with magnetic resonance imaging in
vivo. Proceedings of the National Academy of Sciences of the United States of
America 103: 12,598–12,601.
Chapter 7
Magnetic Control
of Biological Cell Growth
Shoogo Ueno and Sachiko Yamaguchi-Sekino
CONTENTS
7.1 Introduction 186
7.2 Basic Principles for Biomagnetic Effects 188
7.2.1 Effects of Static Magnetic Fields on Biological Systems 188
7.2.1.1 Magnetic Force Acting on a Biological System
in a Homogeneous Magnetic Field 189
7.2.1.2 Magnetic Torque on Biological Cells
in a Homogeneous Magnetic Field 191
7.2.1.3 Magnetic Force Acting on a Biological System
in an Inhomogeneous Magnetic Field 192
7.2.2 Effects of Time-Varying Magnetic Fields on Biological
Systems 194
7.2.2.1 Low-Frequency Magnetic Fields and Pulsed
Magnetic Fields 194
7.2.2.2 High-Frequency Magnetic Fields
and Microwaves 195
7.2.3 Effects of Multiplicity of Magnetic Fields and Other
Forms of Energy 196
7.2.3.1 Radical Pair Model and Singlet-Triplet
Intersystem Crossing 196
7.3 Magnetic Orientation of Adherent Cells 197
7.3.1 Polymerization Processes of Fibrin Fibers under Strong
Static Magnetic Fields 197
7.3.2 Fibrin, Collagen, Osteoblast Cells, Endothelial Cells,
Smooth Muscle Cells, and Schwann Cells 198
185
186 ◾ Biomagnetics
7.1 INTRODUCTION
Cell proliferation and cell death are fundamental phenomena that help in
maintaining a human body’s condition. The number of cells in the body
increases by cell proliferation and decreases by cell death. Because the
regulation of cell proliferation and death is strongly connected to vari-
ous disorders (e.g., cancer or nerve degeneration), these phenomena are
strictly regulated by genes or proteins. The basic principle utilized in the
magnetic control of cell growth is to control the cells’ physiological status
using physical stimuli from static or time-varying (and pulsed) magnetic
fields. Application of these stimuli might affect the cells’ morphological
and functional (including genetic and protein expression) features. The
greatest advantage of this approach is its noninvasiveness, thus avoiding
patient surgery. Another advantage is that this method subjects the patient
to nonionizing radiation, thereby minimizing risks to operators and
patients from harmful ionizing radiation. Furthermore, the simplicity-of-
handling feature of this method implies that the installation of large facili-
ties such as radiation therapy is not required. For these reasons, attempts
at controlling the cell state by magnetic fields have been the focus of active
research in this decade (Ueno 2012).
Magnetic Control of Biological Cell Growth ◾ 187
Sources of exposure for the magnetic control of cell growth can be static
or time-varying (pulsed) magnetic fields. Magneto-mechanical effects and
magnetically induced electric currents are the basic mechanisms of physi-
cal stimuli in the magnetic control of cell growth. A static magnetic field
generates magneto-mechanical effects known as “magnetic orientation”
and “magnetic force,” although this chapter only focuses on medical appli-
cations of the former effect. Historically, the first instance of the magnetic
control of cell growth by a static magnetic field was the polymerization of
fibrin gel in strong magnetic fields, performed by Torbet et al. (1981) and
Torbet and Ronziere (1984). Torbet el al. demonstrated the polymeriza-
tion of fibrin fibers in an 11 T static magnetic field and reported that the
fibers aligned in a particular direction with respect to the magnetic field
(known as the “magnetic orientation”) when the polymerization was car-
ried out slowly (Torbet et al. 1981). Magnetic orientation has been observed
in various biological molecules such as collagen (Murthy 1984; Torbet
and Ronziere 1984; Kotani et al. 2000), and cells such as osteoblast cells
(Kotani et al. 2000, 2002), Schwann cells (Eguch et al. 2003; Eguchi and
Ueno 2005), vascular endothelial cells, and smooth muscle cells (Umeno
et al. 2001; Umeno and Ueno 2003; Iwasaka and Ueno 2003a,b; Iwasaka et
al. 2003). Owing to these experiments, this phenomenon is now a scien-
tifically established effect of static magnetic fields (WHO EHC 232 2006;
ICNIRP 2009). Medical applications of the magnetic orientation have
been demonstrated in various targets (Ueno and Sekino 2006).
Apart from static magnetic fields, studies on the medical applications of
time-varying magnetic fields (pulsed magnetic fields) are also underway.
Magnetic stimulation, established by Barkar et al. (1985) and developed
by Ueno et al. (1988), is a noninvasive electrical stimulation technique.
Because the promising biological effects of this method toward excit-
able tissues (e.g., brain or neurons) have been shown, various medical
applications of magnetic stimulation have been developed in the fields of
brain science or neurology (Ueno 2012). However, applying magnetically
induced force or currents to cancers has only recently been studied and
the noninvasiveness of this method has been found to be a great advantage
for cancer therapy.
We first review the basic principles for biology-magnetism interaction
(named biomagnetic effects) in Section 7.2 and introduce the physical
background of biological effects caused by exposure to static and time-
varying magnetic fields and their multiplicative use with other forms of
energy. In the latter half of this chapter, attempts on the magnetic control
188 ◾ Biomagnetics
of tissues for disorders (Section 7.4, bone growth acceleration; Section 7.5,
nerve regeneration; and Sections 7.6 and 7.7, cancers) will be introduced.
F
Biological
effects
B
B0
a force (Lorentz force) and voltage (flow potential) at right angles to each
other. This effect becomes visible as an artifact in ECGs (electrocardio-
grams), which are recorded under strong static magnetic fields. The latter
effect is a phenomenon that the molecules are aligned in the certain direc-
tion in which molecules are magnetically stable with dependent on their
magnetic anisotropy. Various types of molecules and cells are known to
respond to a static magnetic field as well as exhibit morphological changes.
In an inhomogeneous magnetic field, two major properties affect bio-
logical organisms; these are (1) the “magnetic force” and (2) magnetically
induced currents named “motion-induced currents,” which occur because
of motion in an inhomogeneous field. Magnetic forces are frequently
observed when a paramagnetic material is placed in a magnetic field gra-
dient. This projectile effect is a major concern among magnetic resonance
imaging (MRI) operators from the viewpoint of safe handling of the
equipment. A diamagnetic material such as water also experiences a repul-
sive force when exposed to an inhomogeneous magnetic field. A strong
magnetic field with a high field gradient, on the order of 50–100 T/m, can
generate a partition in water (Ueno and Iwasaka 1994a,b). The latter phe-
nomenon is observable through temporary symptoms, such as dizziness
and headache, near a strong inhomogeneous magnetic field such as that
produced in the vicinity of an MRI system. Proper management and edu-
cation are required to protect MRI operators from these symptoms.
The effect of a static magnetic field on a chemical reaction has been
described as the “radical pair effect” (Ritz et al. 2000; Okano 2008; ICNIRP
2009). The static magnetic field alters the yield of chemical substances by
affecting the intersystem crossing of radicals produced as intermediates in
a photochemical reaction.
concern that blood or the body itself may be subject to Lorentz force and
a flow potential because of magnetic field exposure, because these tissues
are electrically conductive. Indeed, it is well established that when major
arteries of the circulatory system are placed in a magnetic field, an elec-
tric voltage is induced in the blood flowing through them (Kangarlu and
Robitaille 2000).
Figure 7.2a shows examples of magnetic forces acting on biological
systems in homogeneous magnetic fields. The electrical potential (E) that
develops when blood (velocity v) flows in a vessel of cylindrical shape with
a diameter D placed at an angle θ with respect to a static magnetic field B
can be represented by
E = vBD sin θ
E
P T
Q
Ao Ac
S
(a) (b)
1
T = − µ 0 ⋅ B 2∆χsin2θ
2
the two water walls, similar to Moses parting the Red Sea as described in the
Bible (Ueno 2012). Figure 7.3 is a photograph of the Moses effect. The Moses
effect has been a concerning issue in MRI users; however, the degree of the
acceleration was suppressed to 1 percent of that of gravity when the force
was recalculated for the situation of a whole-body in a 3-T magnet setup.
Schenck calculated that the magnetic force at 10 T leads to a change in pres-
sure between the inside of the magnet to the outside of less than 40 mm of
H2O, suggesting that the effect on blood flow is minor (Schenck 2005).
of the induced field strength and current density are proportional to the
radius of the loop, the electrical conductivity of the tissue, and the rate of
change in the magnitude of the magnetic flux density (ICNIRP 2010). As
described in the ICNIRP guideline of 2010, the responsiveness of elec-
trically excitable nerve and muscle tissues to electric stimuli, including
those induced by exposure to low-frequency electromagnetic fields, has
been well established (e.g., Reilly 2002; Saunders and Jefferys 2007). On
the basis of a theoretical calculation using a nerve model, the threshold
value of myelinated nerve fibers in human peripheral nervous system has
been estimated as 6 Vpeak /m (Reilly 1998, 2002; ICNIRP 2010). The most
widely established effect of magnetically induced electric currents is mag-
netic phosphenes: the perception of a faint flickering light in the periphery
of the visual field occurring in the retinas of volunteers exposed to low-
frequency magnetic fields (ICNIRP 2010). As mentioned in the ICNIRP
guideline of 2010, the minimum threshold flux density of 5 mT occurs at
20 Hz. The phosphenes are thought to be a result of the interaction of the
induced electric field with electrically excitable cells in the retina.
Transcranial magnetic stimulation (TMS) is an example of a medical
application of pulsed magnetic fields. Various applications of TMS include
neurological diseases, functional mapping of the brain, and cancer thera-
pies (Ueno 2012). A pulsed magnetic field generated by a coil attached to
the scalp is applied to the brain during TMS. The applied magnetic field
typically has an intensity of 1 T and a pulse duration in the submillisecond
range. It induces an electric current in the brain aimed at nerve excita-
tion. Single- or double-pulse stimulation is normally used for diagnosis
and functional mapping of the brain.
The specific absorption rate should be minimized to the extent that blood
flow and other bodily heat-transfer mechanisms can dissipate the heat.
The magnitude of the specific absorption rate is a critical problem, par-
ticularly in the case of high-field MRI systems (IEC 2010).
Singlet Triplet
H
1 µm
1 µm
(a) (b)
FIGURE 7.5 Polymerization of fibrin fibers under strong static magnetic fields.
Scanning electron micrographs of fibrin prepared without (a) or with (b) mag-
netic field. H represents the direction of the field. The diameter of the fibrin fibers
is −1000 Å. (From Torbet, J. et al., Nature 289: 91–93, 1981.)
198 ◾ Biomagnetics
200 µm 200 µm 50 µm
50 µm 100 µm 100 µm
(×1)
y
5 mm 5 mm
z
(a) x
(×100)
y
Magnetic fields (T)
8.20 T 10 µm 10 µm
8.09
8 7.91
4 7.84
0 z
–200 0 200
Distance (mm)
Alizarin red (21 days)
Superconducting magnet
y
700 mm 10 µm 10 µm
100 mm
Water
x
bath
(b) (c) (d)
FIGURE 7.7 Bone growth acceleration by magnetic fields (in vitro experiments).
(a) Photograph of the horizontal cylindrical-type superconducting magnet.
(b) The longitudinal section of the bore and the distribution of the magnetic
field along the z axis. (c) The cross section of the bore and the magnetic field at
the center of the z axis (x-y plane). (d) Effects of the SMF on the differentiation
and matrix synthesis of cultured MC3T3-E1 cells. Fourteen days and 21 days in
culture after 8-T SMF exposure for 60 h; cells were stained with ALP and aliza-
rin red, respectively. Arrows indicate the direction of the magnetic fields. (From
Kotani, H. et al., Journal of Bone and Mineral Research 17: 1814–1821, 2002.)
Magnetic Control of Biological Cell Growth ◾ 201
Exposed
1 mm
Control
1 mm
(a)
10 µm 10 µm
FIGURE 7.8 Bone growth acceleration by magnetic fields (in vivo experiments).
Histological analyses of the SMF effect on bone formation in and around the
BMP-2/collagen pellets implanted subcutaneously in mice. Twenty-one days
after BMP-2/collagen pellet implantation; samples were obtained and embedded
in paraffin and 5-μm-thick paraffin sections were stained with H&E. (a) Light
microscopic and (b) differential interference contrast microscopic findings of a
representative pellet from each group. The squares in panel a represent respective
areas in panel b. The arrow indicates the direction of the magnetic fields. (From
Kotani, H. et al., Journal of Bone and Mineral Research 17: 1814–1821, 2002.)
202 ◾ Biomagnetics
Neuron
Axonal remnants and Axon Actin fiber
Axonotmesis
myelin debris Neurofilament
Lamellipodium
Lesion Mϕ Schwann cell
Regeneration Schwann cell column
Sprouting
(Bungner band)
Control Exposed
Control Exposure
Incubation for 2 h 8-T exposure for 2 h (37°C)
(37°C)
Implanted
(b)
COL M-COL
20 µm
Numbers and diameters of myelinated fibers (po. 12 weeks)
COL M-COL
FIGURE 7.10 Magnetic field effects on Schwann cells and nerve regeneration.
(a) Light micrographs of a mixed culture of Schwann cells and collagen on the sec-
ond day in culture after 2 h with and without 8-T magnetic field exposure. Scale bars:
50 mm. (b) Set up for in vivo experiment. (c) Morphological examination with tolu-
idine blue staining. The number and the diameter of regenerated axons increased
significantly in the M-COL group compared with the COL group by histological
evaluation. (b and c, From Eguchi, Y. et al., Bioelectromagnetics 36: 233–243, 2015.)
Magnetic Control of Biological Cell Growth ◾ 205
of the Schwann cells was performed microscopically after 60 h for the two
groups, namely, the exposed group (magnetically aligned collagen) and the
control group (unaligned collagen). As shown in Figure 7.10a, light micro-
graphs of a mixed culture of Schwann cells and collagen indicated that cells
cultured with magnetically oriented collagen aligned along the collagen
fibers oriented perpendicular to the magnetic field.
The same author also investigated the effectiveness of using magneti-
cally aligned collagen (after exposure to a maximum 8-T magnetic field)
for nerve regeneration in both an in vitro and in vivo model (Eguchi et al.
2015). Eguchi et al. investigated the orientation of Schwann cells embedded
in collagen gel within a silicone tube as an artificial nerve guide to bridge
sciatic nerve defects. A collagen-cell suspension by mixing collagen type
I solution with culture medium containing Schwann cells were prepared
(Figure 7.10b). As shown in Figure 7.10b, the collagen-cell suspensions were
seeded into a silicone tube (length, 15.0 mm; inner diameter, 1.5 mm) and
the tube was exposed to an 8-T magnetic field for 2 h with the axis perpen-
dicular to the magnetic field. The sciatic nerve trunk was cut in the cen-
tral part of the thigh and the silicone tubes were affixed to both stumps
of the nerve (Figure 7.10b). Four groups were examined: collagen gel only
(COL), magnetically aligned collagen gel (M-COL), collagen gel mixed with
Schwann cells (S-COL), and magnetically aligned collagen gel mixed with
Schwann cells (M-S-COL). The ratio of infiltrating regenerated nerves was
higher in the M-COL group compared to the COL group at 8 weeks post-
operation, and there were no significant differences between the two groups
with and without Schwann cells. The number and diameter of regenerated
axons increased significantly in the M-COL compared with the COL group
at 12 weeks post-operation (Figure 7.10c).
The authors demonstrated that magnetically oriented collagen pro-
moted nerve regeneration using both an in vitro and in vivo model and also
suggest that magnetically aligned collagen acts as a scaffold for Schwann
cells and in combination with aligned Schwann cells can also be used for
biohybrid nerve guide implants to bridge nerve lesions.
investigated (Overgaard 1989; Yoo et al. 2006; Japan Society for Thermal
Medicine). Hyperthermia is a type of cancer treatment in which a body
tissue is exposed to high temperatures (42°C–43°C) for 30–60 min (Japan
Society for Thermal Medicine). The idea of using heat to treat cancer is
based on the fact that temperatures higher than 42.5°C lead to the death
of somatic cells. Research reports indicate that hyperthermia may render
some cancer cells more sensitive to radiation or chemotherapy, in addition
to a killing effect by heating itself. Numerous clinical trials of hyperther-
mia for a variety of cancers such as breast cancer, melanoma (malignant
melanoma), cervical cancer, colorectal cancer, bladder cancer, and cervi-
cal lymph node metastasis have been demonstrated (Yoo et al. 2006; Japan
Society for Thermal Medicine).
However, irradiation of only the targeted tumors noninvasively
remains an issue. For example, to accomplish sufficient localization of
heat requires the implantation of an electrode to the body, instead of using
a whole-body electrode. In addition, whole-body electrodes are ineffec-
tive for deep-seated tumors, implying that only invasive methods using
implanted antennas are available for deep-site tumors.
Alternatively, hyperthermia using harmless magnetizable particles as
a method to treat targeted tumors has been reported (Yanase et al. 1997;
Hafeli and Pauer 1999; Wada et al. 2001; Ogiue-Ikeda et al. 2003, 2004).
The most important properties of magnetic particles for clinical diagnos-
tics and medical therapies are nontoxicity, biocompatibility, injectability,
and high-level accumulation in the target tissue or specific organ, i.e.,
being strictly and spatially confined to the planned region of the internal
body. However, the possibility that the drugs might damage normal tis-
sues as they circulate throughout the body before reaching their target
remains. Further, although magnetized particles are accumulated only in
the cancer tissue, application of a whole-body electrode for hyperthermia
may induce heat inside the normal tissue as well as the cancer because of
poor localization.
Instead, pulsed magnetic stimulation has recently been proposed as
a source of electromagnetic field for cancer therapy using magnetizable
particles (Ogiue-Ikeda et al. 2003, 2004). As introduced earlier, pulsed
magnetic stimulation is a method for stimulating biomedical tissues non-
invasively and with good localization. The next section introduces a new
method to destruct targeted cells by using magnetizable beads and pulsed
magnetic force.
Magnetic Control of Biological Cell Growth ◾ 207
a b c d a b c d a b c d a b c d
Nonstimulated Stimulated
60
40
20
0
Control1 2 3 4 Control1 2 3 4
(b) 4.5 µm beads 2.8 µm beads
2.25
2
Tumor weight (g)
1.75 ***
1.5
Coil 1.25
1
Tumor Tumor 0.75
0.5
0.25
0
(c) Stim (+) Sham
Stim 0.25 T Sham
(a) (b)
(d)
120
% Control (TCC-S + RPMS)
100
80 **
*
60
: Control
40 : 0.1 T 1000 pulses
: 0.25 T 1000 pulses
20 : 0.5 T 1000 pulses
: 0.25 6000 pulses
0
TCC-S + RPMS TCC-S +Imatinib 100 nM
+ RPMS
the stimulus condition described above. The authors concluded that these
results showed the first evidence of antitumor and immunomodulatory
effects brought about by the application of repetitive magnetic stimulation
and also suggested the possible relationship between antitumor effects and
the increase of TNF-α levels caused by pulsed magnetic stimulation.
Next, the combined effect of magnetic stimulation and an anticancer
agent was examined on human chronic myelogenous leukemia-derived
cell line TCC-S using molecular target drug (selective tyrosine kinase
inhibitor) imatinib mesylate (imatinib). The stimulus conditions were
determined as follows: 0.1, 0.25, and 0.5 T; 25 pulses/sec; 1000, 3000, and
6000 pulses/day. TCC-S cells were cultured with 100 nM imatinib and
exposed to RPMS at 1, 12, 24, 36, 48, and 56 h after drug treatment. The
combined effect of magnetic stimulation and imatinib depended on
the stimulus intensity and the pulse dose (Figure 7.14). To further clar-
ify the effect of magnetic stimulation on human normal lymphocytes,
they too were exposed to magnetic stimulation with or without imatinib.
Magnetic stimulation was found to have no effect on the viability of nor-
mal lymphocytes. These results indicate that magnetic stimulation pos-
sibly improves the effectiveness of anticancer agents.
REFERENCES
Barkar, A. T, Jalinous, R. I., and Freestone, I. L. 1985. Non-invasive magnetic stim-
ulation of human motor cortex. Lancet 1: 1106–1107.
Bradley, J. K., Nyekiova, M., Price, D. L., Lopez, L. D., and Crawley, T. 2007.
Occupational exposure to static and time-varying gradient magnetic fields
in MR units. Journal of Magnetic Resonance Imaging 26: 1204–1209.
Magnetic Control of Biological Cell Growth ◾ 213
Bruce, G. K., Howlen, C. A., and Huckstep, R. L. 1987. Effect of a static magnetic
field on fracture healing in a rabbit radius. Clinical Orthopaedics 222: 300–306.
Cabrales, L. B., Ciria, H. C., Bruzon, R. P., Quevedo, M. S., Aldana, R. H., De Oca,
L. M., Salas, M. F., and Pena, O. G. 2001. Electrochemical treatment of mouse
Ehrlich tumor with direct electric current. Bioelectromagnetics 22(5): 316–322.
Capstick, M., McRobbie, D., Hand, J., Christ, A., Kuhn, S., Hansson Mild, K.,
Cabot, E., Li, Y., Melzer, A., Papadaki, A., Prüssmann, K., Quest, R., Rea,
M., Ryf, S., Oberle, M., and Kuster, N. 2008. An investigation into occupa-
tional exposure to electro-magnetic fields for personnel working with and
around medical magnetic resonance imaging equipment. Report on Project
VT/2007/017 of the European Commission Employment, Social Affairs and
Equal Opportunities DG. 2008. Available from http://www.myesr.org/html
/img/pool/VT2007017FinalReportv04.pdf (accessed August 15, 2013).
Chakeres, D. W., and de Vocht, F. 2005. Static magnetic field effects on human sub-
jects related to magnetic resonance imaging systems. Progress in Biophysics
and Molecular Biology 87: 255–265.
Chou, C.-K., McDougall, J. A., Ahn, C., and Vora, N. 1997. Electrochemical treat-
ment of mouse and rat fibrosarcomas with direct current. Bioelectromagnetics
18: 14–24.
Cook, S. D. 1999. Preclinical and clinical evaluation of osteogenic protein-1 (BMP-
7) in bony sites. Orthopedics 22: 669–671.
Crozier, S., and Liu, F. 2005. Numerical evaluation of the fields induced by body
motion in or near high-field MRI scanners. Progress in Biophysics and
Molecular Biology 87: 267–278.
David, S. L., Absolom, D. R., Smith, C. R., Gams, J., and Herbert, M. A. 1985. Effect
of low level direct current on in vivo tumor growth in hamsters. Cancer
Research 45(11 Pt 2): 5625–5631.
de Vocht, F., van Drooge, H., Engels, H., and Kromhout, H. 2006. Exposure, health
complaints and cognitive performance among employees of an MRI scan-
ners manufacturing department. Journal of Magnetic Resonance Imaging 23:
197–204.
Eguchi, Y., and Ueno, S. 2005. Stress fiber contributes to rat Schwann cell orienta-
tion under magnetic field. IEEE Transactions on Magnetics 41: 4146–4148.
Eguchi, Y., Ogiue-Ikeda, M., and Ueno, S. 2003. Control of orientation of rat Schwann
cells using an 8-T static magnetic field. Neuroscience Letters 351: 130–132.
Eguchi, Y., Ohtori, S., and Ueno, S. 2015. The effectiveness of magnetically aligned
collagen for neural regeneration in vitro and in vivo. Bioelectromagnetics 36:
233–243.
Glazer, P. A., Heilmann, M. R., Lotz, J. C., and Bradford, D. S. 1997. Use of elec-
tromagnetic fields in a spinal fusion. A rabbit model. Spine 22: 2351–2356.
Glover, P. M., Cavin, I., Qian, W., Bowtell, R., and Gowland, P. A. 2007.
Magnetic-field-induced vertigo: A theoretical and experimental investiga-
tion. Bioelectromagnetics 28: 349–361.
Grissom, C. B. 1995. Magnetic field effects in biology—A survey of possible mech-
anisms with emphasis on radical-pair recombination. Chemical Reviews 95:
3–24.
214 ◾ Biomagnetics
Hafeli, U. O., and Pauer, G. J. 1999. In vitro and in vivo toxicity of magnetic micro-
spheres. Journal of Magnetism and Magnetic Materials 194: 76–82.
Harguindey, S. 1982. Hydrogen ion dynamics and cancer: An appraisal. Medical
and Pediatric Oncology 10(3): 217–236.
Harkins, T. T., and Grissom, C. B. 1994. Magnetic field effects on B12 ethanol-
amine ammonia lyase: Evidence for a radical mechanism. Science 263(5149):
958–960.
Higashi, T., Yamagishi, A., Takeuchi, T., Kawaguchi, N., Sagawa, S., Onishi, S., and
Date, M. 1993a. Orientation of erythrocytes in a strong static magnetic field.
Blood 82: 1328–1334.
Higashi, T., Sagawa, S., Kawaguchi, N., and Yamagishi, A. 1993b. Effects of a strong
static magnetic field on blood platelets. Platelets 4: 341–342.
Higashi, T., Yamagishi, A., Takeuchi, T., and Date, M. 1995. Effects of static mag-
netic fields on erythrocyte rheology. Bioelectrochemistry and Bioenergetics
36: 101–108.
Hudson, M. 2006. The EU Physical Agents (EMF) Directive and its impact on
MRI imaging in animal experiments: A submission by FRAME to the HSE.
Alternatives to Laboratory Animals 34: 343–347.
Humphrey, C. E., and Seal, E. H. 1959. Biophysical approach toward tumor regres-
sion in mice. Science 130: 388–389.
Ide, C. 1996. Peripheral nerve regeneration. Neuroscience Research 25: 101–121.
International Commission on Non-Ionizing Radiation Protection. 1998. Guide
lines for limiting exposure to time-varying electric, magnetic, and electro-
magnetic fields (up to 300 GHz). Health Physics 74: 494–522.
International Commission on Non-Ionizing Radiation Protection. 2004. Medical
magnetic resonance (MR) procedures: Protection of patients. Health Physics
87: 197–216.
International Commission on Non-Ionizing Radiation Protection. 2009. Guidelines
on limiting exposure to static magnetic fields. Health Physics 96: 504–514.
International Commission on Non-Ionizing Radiation Protection. 2010. ICNIRP
guidelines for limiting exposure to time-varying electric and magnetic fields
(1 Hz to 100 kHz). Health Physics 99: 818–836.
International Commission on Non-Ionizing Radiation Protection. 2014. Guide
lines for limiting exposure to electric fields induced by movement of the
human body in a static magnetic field and by time-varying magnetic fields
below 1 Hz. Health Physics 106(3): 418–425.
International Electrotechnical Commission. 2010. Medical electrical equipment—
Part 2Y33: Particular requirements for the basic safety and essential perfor-
mance of magnetic resonance equipment for medical diagnosis. Geneva:
IEC; IEC 60601-2-33 ed 3.0.
Iwasaka, M., and Ueno, S. 2003a. Polarized light transmission of smooth muscle cells
during magnetic field exposures. Journal of Applied Physics 93: 6701–6703.
Iwasaka, M., and Ueno, S. 2003b. Detection of intracellular macromolecule behav
ior under strong magnetic fields by linearly polarized light. Bioelectro
magnetics 24: 564–570.
Magnetic Control of Biological Cell Growth ◾ 215
Iwasaka, M., Ueno, S., and Tsuda, H. 1994. Effects of magnetic fields on fibrinoly-
sis. Journal of Applied Physics 75: 105–107.
Iwasaka, M., Miyakoshi, J., and Ueno, S. 2003. Magnetic field effects on assembly
pattern of smooth muscle cells. In Vitro Cellular & Developmental Biology—
Animal 39: 120–123.
Japan Society for Thermal Medicine. http://www.jsho.jp.
Kangarlu, A., and Robitaille, P. M. L. 2000. Biological effects and health implica-
tions in magnetic resonance imaging. Concepts in Magnetic Resonance, 12:
321–359.
Kangarlu, A., Burgess, R. E., Zhu, H., Nakayama, T., Hamlin, R. I., Abduljalh,
A. M., and Robitaille, P. M. 1999. Cognitive, cardiac, and physiological safety
studies in ultrahigh field magnetic resonance imaging. Magnetic Resonance
Imaging 17: 1407–1416.
Kano, Y., Akutsu, M., Tsunoda, S., Mano, H., Sato, Y., Honma, Y., and Furukawa, Y.
2001. In vitro cytotoxic effects of a tyrosine kinase inhibitor STI571 in com-
bination with commonly used antileukemic agents. Blood 97: 1999–2007.
Karpowicz, J., Gryz, K., Politański, P., and Zmyślony, M. 2011. Exposure to static
magnetic field and health hazards during the operation of magnetic reso-
nance scanners (in Polish). Medycyna Pracy 62: 309–321.
Kato, M. (Ed.) 2006. Electromagnetics in Biology. Springer.
Kotani, H., Iwasaka, M., Ueno, S., and Curtis, A. 2000. Magnetic orientation of col-
lagen and bone mixture. Journal of Applied Physics 87: 6191–6193.
Kotani, H., Kawaguchi, H., Shimoaka, T., Iwasaka, M., Ueno, S., Ozawa, H.,
Nakamura, K., and Hoshi, K. 2002. Strong static magnetic field stimulates
bone formation to a definite orientation in vivo and in vitro. Journal of Bone
and Mineral Research 17: 1814–1821.
Marie, P. J., Debiais, F., and Haÿ, E. 2002. Regulation of human cranial osteo-
blast phenotype by FGF-2, FGFR-2 and BMP-2 signaling. Histology and
Histopathology 17(3): 877–885.
McRobbie, D. W. 2012. Occupational exposure in MRI. British Journal of Radiology
85: 293–312.
Miklavcic, D., An, D., Belehradek, Jr., J., and Mir, L. M. 1997. Host’s immune
response in electrotherapy of murine tumors by direct current. European
Cytokine Network 8(3): 275–279.
Miller, G. J., Burchardt, H., Enneking, W. F., and Tylkowski, C. M. 1984.
Electromagnetic stimulation of canine bone grafts. Journal of Bone and Joint
Surgery, American 66: 693–698.
Murthy, N. S. 1984. Liquid crystallinity in collagen solutions and magnetic orien-
tation of collagen fibrils. Biopolymers 23: 1261–1267.
Nagakura, S., Hayashi, H., and Azumi, T. 1998. Dynamic Spin Chemistry: Magnetic
Controls and Spin Dynamics of Chemical Reactions. Kodansha, Tokyo.
Nakajima, F., Nakajima, A., Ogasawara, A., Nakajima, A., Goto, K., Shimizu, S.,
Moriya, H., and Einhorn, T. A. 2007. Effects of a single percutaneous injec-
tion of basic fibroblast growth factor on the healing of a closed femoral shaft
fracture in the rat. Calcified Tissue International 81: 132–138.
216 ◾ Biomagnetics
Sersa, G., Miklavcic, D., Batista, U., Novakovic, S., Bobanovic, F., and Vodovnik,
L. 1992. Anti-tumor effect of electrotherapy alone or in combination with
interleukin-2 in mice with sarcoma and melanoma tumors. Anti-Cancer
Drugs 3(3): 253–260.
Sersa, G., Jarm, T., Kotnik, T., Coer, A., Podkrajsek, M., Sentjurc, M., Miklavcic,
D., Kadivec, M., Kranjc, S., Secerov, A., and Cemazar, M. 2008. Vascular dis-
rupting action of electroporation and electrochemotherapy with bleomycin
in murine sarcoma. British Journal of Cancer 98(2): 388–398.
Shellock, F. G., and Kanal, E. 1994. Magnetic Resonance: Bioeffects, Safety, and
Patient Management. Raven Press, New York.
Shiga, T., Okazaki, M., Seiyama, A., and Maeda, N. 1993. Paramagnetic attrac-
tion of erythrocyte flow due to an inhomogeneous magnetic field. Bioelec
trochemistry and Bioenergetics 30: 181–188.
So, P. P. M., Stuchly, M. A., and Nyenhuis, J. A. 2004. Peripheral nerve stimu-
lation by gradient switching fields in magnetic resonance imaging. IEEE
Transactions on Biomedical Engineering 51: 1907–1914.
Takano-Yamamoto, T., Kawakami, M., and Sakuda, M. 1992. Effect of a pulsing elec-
tromagnetic field on demineralized bone-matrix-induced bone formation in a
bony defect in the premaxilla of rats. Journal of Dental Research 71: 1920–1925.
Togawa, T., Okai, O., and Oshima, M. 1967. Observation of blood flow E.M.F. in
externally applied strong magnetic field by surface electrodes. Medical and
Biological Engineering 5(2): 169–170.
Torbet, J., and Ronziere, M. 1984. Magnetic alignment of collagen during self-
assembly. Biochemical Journal 219: 1057–1059.
Torbet, J., Freyssinet, J. M., and Hudry-Clergeon, G. 1981. Oriented fibrin gels
formed by polymerization in strong magnetic fields. Nature 289: 91–93.
Ueno, S. 2012. Studies on magnetism and bioelectromagnetics for 45 years:
From magnetic analog memory to human brain stimulation and imaging.
Bioelectromagnetics 33(1): 3–22.
Ueno, S., and Iwasaka, M. 1994a. Properties of diamagnetic fluid in high gradient
magnetic fields. Journal of Applied Physics 75: 7177–7179.
Ueno, S., and Iwasaka, M. 1994b. Parting of water by magnetic fields. IEEE
Transactions on Magnetics 30: 4698–4700.
Ueno, S., and Sekino, M. 2006. Biomagnetics and bioimaging for medical applica-
tions. Journal of Magnetism and Magnetic Materials 304: 122–127.
Ueno, S., Tashiro, T., and Harada, K. 1988. Localized stimulation of neuronal tis-
sues in the brain by means of paired configuration of time-varying magnetic
fields. Journal of Applied Physics 64: 5862–5864.
Ueno, S., Iwasaka, M., and Tsuda, H. 1993. Effects of magnetic fields on fibrin poly
merization and fibrinolysis. IEEE Transactions on Magnetics 29: 3352–3354.
Umeno, A., and Ueno, S. 2003. Quantitative analysis of adherent cell orientation
influenced by strong magnetic fields. IEEE Transactions on Nanobioscience
2: 26–28.
Umeno, A., Kotani, H., Iwasaka, M., and Ueno, S. 2001. Quantification of adher-
ent cell orientation and morphology under strong magnetic fields. IEEE
Transactions on Magnetics 37: 2909–2911.
218 ◾ Biomagnetics
Vassilev, P. M., Dronzine, R. T., Vassileva, M. P., and Georgiev, G. A. 1982 Parallel
assays of microtubules formed in electric and magnetic fields. Bioscience
Reports 2: 1025–1029.
Wada, S., Yue, L., Tazawa, K., Furuta, I., Nagae, H., Takemori, S., and Minamimura,
T. 2001. New local hyperthermia using dextran magnetic complex (DM) for
oral cavity: Experimental study in normal hamster tongue. Oral Diseases
7(3): 192–195.
Wilén, J., and de Vocht, F. 2010. Health complaints among nurses working near
MRI scanners—A descriptive pilot study. European Journal of Radiology 80:
510–513.
World Health Organization. 1987. Environmental Health Criteria 69: Magnetic
Fields. Geneva, Switzerland.
World Health Organization. 2006. Environmental Health Criteria 232: Static
Fields. Geneva, Switzerland.
World Health Organization. 2007. Environmental Health Criteria 238: Extremely
Low Frequency (ELF) Fields. Geneva, Switzerland.
Yamagishi, A., Takeuchi, T., Higashi, T., and Date, M. 1992. Diamagnetic orienta-
tion of blood cells in high magnetic field. Physica B 177: 523–526.
Yamaguchi, S., Ogiue-Ikeda, M., Sekino, M., and Ueno, S. 2004. The effect of repet-
itive magnetic stimulation on the tumor development. IEEE Transactions on
Magnetics 40: 3021–3023.
Yamaguchi, S., Ogiue-Ikeda, M., Sekino, M., and Ueno, S. 2005. Effects of mag-
netic stimulation on tumors and immune function. IEEE Transactions on
Magnetics 41: 4182–4184.
Yamaguchi, S., Ogiue-Ikeda, M., Sekino, M., and Ueno, S. 2006a. The effect of
repetitive magnetic stimulation on tumor and immune functions in mice.
Bioelectromagnetics 27: 64–72.
Yamaguchi, S., Sato, Y., Sekino, M., and Ueno, S. 2006b. Combination effects of
the repetitive pulsed magnetic stimulation and the anticancer agent imatinib
on human leukemia cell line TCC-S. IEEE Transactions on Magnetics 42:
3581–3583.
Yamaguchi-Sekino, S., Sekino, M., and Ueno, S. 2011. Biological effects of elec-
tromagnetic fields and recently updated safety guidelines for strong static
magnetic fields. Magnetic Resonance in Medical Sciences 10: 1–10.
Yamaguchi-Sekino, S., Nakai, T., Imai, S., Izawa, S., and Okuno, T. 2014.
Occupational exposure levels of static magnetic field during routine MRI
examination in 3T MR system. Bioelectromagnetics 35: 70–75.
Yanase, M., Shinkai, M., Honda, H., Wakabayashi, T., Yoshida, J., and Kobayashi,
T. 1997. Intracellular hyperthermia for cancer using magnetite cationic lipo-
somes: Ex vivo study. Japanese Journal of Cancer Research 88: 630–632.
Yiu, G., and Zhigang, H. 2006. Glial inhibition of CNS axon regeneration. Nature
Reviews Neuroscience 7: 617–627.
Yoo, J. L., Kim, H. R., and Lee, Y. J. 2006. Hyperthermia enhances tumour necro-
sis factor-related apoptosis-inducing ligand (TRAIL)-induced apoptosis in
human cancer cells. International Journal of Hyperthermia 22(8): 713–728.
Chapter 8
Effects of Radio
Frequency Magnetic
Fields on Iron Release
and Uptake from and
into Cage Proteins
Oscar Cespedes and Shoogo Ueno
CONTENTS
8.1 Introduction 220
8.1.1 Interactions of Radio Frequency Magnetic Fields
with Iron Release and Uptake 220
8.1.2 Possible Medical Consequences and Applications
of RF Magnetic Fields on Iron Biochemistry 221
8.2 Iron Cage Proteins 222
8.2.1 Introduction to Ferritins and Their Physiological Role 222
8.2.2 Iron Release and Absorption 224
8.2.2.1 Spectroscopic Techniques in the Determination
of Iron Contents 224
8.2.2.2 Function of the Symmetry Points in Iron
Release and Uptake 225
8.2.2.3 Role of Iron in Oxidative Stress and Dementia 227
8.3 Magnetic Properties of Ferrihydrite Bionanoparticles 229
8.3.1 Nanoscale Magnetism 229
219
220 ◾ Biomagnetics
8.1 INTRODUCTION
8.1.1 Interactions of Radio Frequency Magnetic
Fields with Iron Release and Uptake
Most biological components are diamagnetic or paramagnetic and only
interact weakly with magnetic fields. For this reason, the effect of nonion-
izing radio frequency (RF) radiation in biology is commonly described
in terms of heating.1 However, iron is an essential element to most living
organisms, including, of course, humans. A healthy adult needs between
10 (males) and 18 mg (females) of iron per day, the highest concentra-
tion of any metallic element.2 Although it has a crucial role in biochemical
mechanisms such as oxygen transport and redox processes, iron is also a
potential toxic agent to cells via the Fenton reaction. Therefore, it requires
a complex set of regulatory mechanisms to meet the demands of the body
while avoiding accumulation or release in the wrong environment.3
A key constituent in the chemical control of iron is ferritin. This cage
protein is the main form of storage for mineral iron in biology, to the
Effects of RF Magnetic Fields on Iron Release and Cage Proteins ◾ 221
point that the clinical assessment of iron content in blood is done via
ferritin concentration. The role of the protein is to oxidize and store
iron ions in the form of a ferrihydrite nanoparticle. Ferritin proteins
have also been associated with the function of the mitochondria.4,5 This
emphasizes the role of the protein in protecting against oxidative dam-
age, given that most of the iron must pass through the organelle to par-
ticipate in biological processes.4–6 The core nanoparticle can contain up
to 4500 iron ions, which results in a high magnetic moment and a super-
paramagnetic magnetization at room temperature. The only compo-
nents found in biology with a higher magnetic moment than ferritin are
mineral biogenic magnetite nanocrystals. These have also been consid-
ered as a possible transducer for field effects,7,8 but no effect was observed
in the mortality rates of magnetic bacteria when the nanocrystals were
exposed to RF fields.9,10
Being the organic bioelement with the highest magnetic moment,
it is perhaps not unexpected that ferritin shows strong effects even at
weak magnetic fields and relatively low frequencies in the RF range. The
inner superparamagnetic ferrihydrite nanoparticle increases its internal
energy when exposed to alternating magnetic fields via Néel absorption/
relaxation. This energy is irradiated to the surrounding proteic cage,
altering the protein ability to uptake and release iron. Proteins exposed
during several hours have iron intake and chelation rates up to a factor of
3 smaller than control samples although for concentrations not observed
in physiological conditions. These results are examined more closely in
the following sections, as they open new paths of research for biological
effects of alternating magnetic fields.
proportional to the amount of iron in the body, the relaxation times and
frequency dependence of power dissipation will vary greatly with the type
of ferritin. From estimates of the power released, it is easy to see that the
energy loss by the nanoparticle inside ferritin is unlikely to lead to a ther-
mal effect at the concentrations usually found in the body, even at rela-
tively large magnetic fields and frequencies. This is due to the relatively
low susceptibility of ferrihydrite (as compared to, for example, magnetite
nanoparticles) and to the low concentration of ferritin in healthy organ-
isms. It is also difficult to estimate how much power is transmitted to the
organic cage or how well the peptides may dissipate the energy into the
surrounding medium. If the effects were to be significant for in vivo exper-
iments, there are two possible considerations: Will the RF exposure result
in impaired protein function? And, can we use this exposure to regulate
iron chemistry?
As for the first question, any damage to the role of ferritin in biochem-
istry may have serious consequences to the health of an organism. For
example, it could result in iron release in the brain, which would then
lead to oxidative-stress related illnesses, e.g., some forms of dementia.
However, a priori it seems that the frequency and amplitude of the fields
that would be needed for these severe effects would have to be very large,
unlikely to be found in any natural environment. These large RF fields
would probably generate effects via heating of the bio-environment.
On the other hand, it is possible to generate a controlled exposure
to high-frequency/amplitude magnetic fields without generating dipo-
lar electric heat, see the numerous experiments in hyperthermia.15–17
Similarly, it might be possible to expose certain regions in the body to
alternating magnetic fields so that the iron release/uptake is done more
slowly, which could be beneficial in case of chelating agents being present
in the body, genetic misregulations, etc. Further research could also help
to distinguish the mechanisms affecting iron release from those in iron
uptake, in which case the magnetic field exposure could be used to stimu-
late one above the other.
harmful Fe (II) ions, oxidize, and store them in a safe manner. This is
done inside the peptidic cage of ferritin as a relatively harmless ferrihy-
drite nanoparticle (9H2O·5Fe2O3) with small concentrations of P, Cs, Cd,
Zn, Cu, and S. Ferritin will later deliver the iron ions under certain chemi-
cal signals or agents when it is needed for metabolic purposes. The proteic
cage (apoferritin) is formed by 24 subunits about 180 peptides each and
arranged in a 4-3-2 symmetry to form a roughly spherical cage of 440 kDa
in mass and 12 nm in diameter. The inner cavity is roughly 8 nm in size,
so it can store particles with up to 4500 iron ions. A typical ferritin protein
contains of the order of 1000–1500 ions.18 Usually, ferritin is found mostly
in the liver and the spleen, the main places for iron storage, but in some
pathologies, ferritin can also be found in high concentrations in other
organs, such as the heart and the brain. In order to ensure that the iron
is protected in the different biological environments, ferritin has evolved
into a very stable and robust configuration. The peptidic arrangement
only dissociates at extreme conditions with temperatures above 80°C or
at pH extremes of 2.8 and 11.2, unlikely to be found in environments that
host living organisms (see Figure 8.1).
Peptidic cage
(apoferritin)
~12 nm
Ferrihydrite
nanoparticle
(a) (b)
(c)
Charge
FIGURE 8.2 (a) Molecular representation of the polar threefold symmetry point,
exit point for Fe (III) ions, and the protein aggregation center. (b) Zoom on the
six carboxyl groups from the glutamic acids (COO−) that form the hydrophilic
terminals at the threefold symmetry point. (c) Charge distribution at the sym-
metry point.
originates from the six carboxyl groups of aspartic and glutamic acids at
the terminals. If the threefold point is blocked, for example, because of a
rapid release and aggregation of iron at the pore, the proteins may aggre-
gate and precipitate. The fourfold symmetry point is nonpolar and there-
fore does not contribute to molecular aggregation. Whereas iron release
seems to take place predominantly or exclusively at the threefold symme-
try point, iron uptake occurs probably at the two- and fourfold symmetry
points.21,22
Effects of RF Magnetic Fields on Iron Release and Cage Proteins ◾ 227
The process of ferroxidation by which iron ions are oxidized and later
incorporated into the core ferrihydrite nanoparticle is similar to that of
Fenton chemistry (see Section 8.2.2.3). Ferroxidase enzyme molecules in
the proteic cage act as catalysts and binding agents to facilitate the oxi-
dation and incorporation of iron (II). These ferroxidase centers are com-
monly found at the protein symmetry point with four helical bundles,
although they can also be present in the two subunit links. For some bac-
terial ferritins, these centers may also be found at all 24 subunits. The role
of the ferroxidase is to act as catalyst in the reactions:23
Ferritins without a ferroxidase center can still incorporate iron and form a
mineral core, but the process is slower and these proteins can probably func-
tion only in environments low in iron and/or where it is not essential.
Amyloid-β mediated
lipid peroxidation
Cell rupture
synapse disruption
Anemia Fibromyalgia
thrombocytosis stroke Alzheimer’s
etc. etc. disease
FIGURE 8.3 Potential physiological pathways for iron-related illnesses and the
role of ferritin function.
Effects of RF Magnetic Fields on Iron Release and Cage Proteins ◾ 229
8.3.1.1 Superparamagnetism
In order to minimize their internal energy, ferro-, ferri-, and anti-
ferromagnetic materials form domains with aligned magnetization.
These domains can have sizes from ~0.1 to 100 μm, depending on the
exchange energy, temperature, and magnetic anisotropy of the mate-
rial. If the magnetic material is nanostructured below the single domain
size, its magnetic behavior is different from that of conventional fer-
romagnets or paramagnets. The particle behaves then as a single mac-
rospin whose direction fluctuates owing to the thermal energy but can
also be aligned in a magnetic field. Below a certain temperature, the
magnetization direction at the measuring speed is blocked, and the
hysteresis loop has a net remanence and coercivity different from zero.
Above the blocking temperature, the nanomaterial has no coercivity
or remanence but still has a very large susceptibility well above that
of conventional paramagnets, hence the name of superparamagnetic
particles. Different from paramagnets with a constant susceptibility,
superparamagnets above the blocking temperature display two differ-
ent susceptibilities for low and high magnetic fields, the characteristic
sigma shape.
One of the most common methods to study superparamagnetic nano
particles is to measure the difference between zero field cooled and field
cooled low field susceptibility. If the particles are cooled with no field
applied below the blocking temperature, only a small percentage of the par-
ticles will align once a small field is applied. As the temperature increases,
more particles can surmount the energy barrier with the aligned state and
the magnetization increases with increasing temperature. Once the sys-
tem reaches the blocking temperature, the magnetization reaches a maxi-
mum and then decays as the magnetic energy becomes weaker compared
to the temperature. However, if the sample is cooled in a field, particles
that are aligned with the field at high temperatures are frozen in the high
magnetization state as we cross the blocking temperature, and, as we cool
down the sample, more particles go into a high magnetization state. In the
case of the ferrihydrite nanoparticles in ferritin, this blocking temperature
is of the order of 10–20 K (see Figure 8.4).
2.5 τN = 0.13 ns
5
ln f
0
2.0
' (10–3 m3/kg) 50 60 70
1.5 1000/T (K)
0.1 Hz
1 Hz
1.0 10 Hz
100 Hz
1 kHz
0.5
0 10 20 30 40 50 60
Temperature (K)
FIGURE 8.4 Real component of the susceptibility χ′ and the dependence of the
blocking temperature (TB) with the measuring frequency for superparamagnetic
ferrihydrite nanoparticles (ferritin solution). The dependence using Equation 8.11
leads to a Néel relaxation time τN of 0.13 ns (τ0 = 3 psec). Here, the AC field was set
to 0.5 mT and DC field to zero.
∆U = −µ 0 !∫ M dH (8.7)
where μ0 is the magnetic permeability in vacuum, H is the magnetic field
intensity, M is the magnetization, and ΔU is the energy absorbed/dissipated
per cycle.
In the case of nanoparticles, the physical description of the power
absorption can be more complicated. Having no coercivity, the area of
the hysteresis loop of a superparamagnetic nanoparticle is nominally
zero and therefore should dissipate no energy. However, for sufficiently
fast alternating magnetic fields, the nanoparticles’ magnetization lags the
external applied field. In that case, the imaginary component of the sus-
ceptibility is different from zero, and this results in an energy gain by the
nanoparticle of:
2π/ω
∆U = 2µ 0 H χ′′
2
0
∫ sin ωt dt (8.8)
0
2
τ N = τ0 exp ⎛⎜
KV ⎞
(8.9)
⎝ kBT ⎟⎠
K
TB =
f (8.10)
kB ln ⎛⎜ ⎞⎟
⎝ f0 ⎠
with TB the Blocking temperature, f0 the attempt frequency (1/τ0), and f the
frequency of the applied magnetic field. By plotting ln(f) versus 1/TB, K/kB
is the slope, and ln(f0) the crossing with the abscise. In ferritin samples we
obtain τ0 ~3 × 10–11 s and K ~ 370 K, which gives τN ~ 10–10 s (Figure 8.4).
On the other hand, the Brownian relaxation time τB is:
3ηVH
τB = (8.11)
kBT
with η the medium viscosity (~1 mPa · s–1) and VH the hydrodynamic
particle volume (~10−25–10−24 m3 for ferritin). The result is that τB ~ 10−7–
10−8 s ≫ τN ⇨ τN ≅ τ, so the relaxation mechanism for ferrihydrite
nanoparticles in ferritin is of Néel type, i.e., the particle, and the protein
around it, will not move will the magnetic field.
8.3.2.2 AC Susceptibility
If the magnetic field varies faster that the characteristic time for the mag-
netization (τN or τB), the particle finds itself in the “wrong” magnetic state
after each cycle; that is, it is not aligned with the magnetic field. This rep-
resents an energy gain that is then dissipated in the form of spin waves
and/or heat. From Equation 8.8, it can be derived that the power loss of a
superparamagnetic nanoparticle in a magnetic field is:36
2 πf τ
P = πµ 0 χ0 H 02 f × (8.12)
1 + (2 πf τ)2
150
FC
(10–6 m3/kg)
100
50
ZFC
0
4 16 28 40 160
Temperature (K)
FIGURE 8.5 Zero field cooled – field cooled (ZFC – FC) magnetic susceptibility
χ for a saturated ferritin solution. The curve is characteristic of superparamag-
netic nanoparticles with a blocking temperature of ~15 K (note the log x-scale).
the molecule may penetrate the proteic cage through the threefold sym-
metry points. The protein is then well protected against strong, nega-
tively charged, iron chelators such as ferrozine. When using ferrozine as
chelating agent, we found that samples previously exposed to magnetic
fields of 1 MHz and 30 μT released up to 40 to 50 percent less iron than
control samples.11 We define the magnetic field effect on the change of
iron released as ΔFereleased ≡ (Fereleased|control − Fereleased|exposed)/Fereleased|control,
where Fereleased|control and Fereleased|exposed are the total iron released 1 hour
after adding the chelating agent discounting the initial burst in control
and exposed samples, respectively. This effect was dependent on several
parameters, including the magnetic field amplitude and frequency, the
exposure time, the ferritin and ferrozine concentration, and the pH of the
solution.
The variation of the magnetic field effect with chemical concentrations
of ferrozine and ferritin (Figure 8.6) can be explained as due to pH varia-
tions and changes in the intermolecular interactions. The pH of the ferri-
tin solution decreases when the acidic ferrozine is added, which will result
in iron reduction and increased rates of release. On the other hand, the
protein itself and the NaCl solution in which it is dissolved act as a buf-
fer to the pH change. Higher rates of iron release due lower pH will affect
both control and exposed samples, decreasing the magnetic field effect.
The reason for smaller effects at lower ferritin concentrations (1 μM) may
60
Ferritin concentration
14 µM
3.5 µM
1 µM
40
∆Fe released (%)
20
0
50 µM 100 µM 200 µM
Chelator (ferrozine) concentration
FIGURE 8.6 Changes in iron release after exposure to magnetic field as a func-
tion of the chemical concentrations of ferritin and ferrozine (chelating agent)
used.
238 ◾ Biomagnetics
of the field. On the other hand, if the effect depended on the electrical
field, such as due to induced polarization changes, it would be expected to
increase at higher frequencies because the measured electrical field gener-
ated by the coils at 15 μT and 2 MHz is 4 times higher than at 120 μT and
250 kHz (see Figure 8.7). The effect should also in that case vanish shortly
after stopping the applied field, in a time scale similar to the dipolar relax-
ation, i.e., the relaxation of the electrical dipoles in the protein cage. In
fact, the equivalent effect obtained for constant products is in agreement
with a mechanism mediated by the RF power dissipation of superpara-
magnetic nanoparticles as calculated in Equation 8.12.
Assuming that the energy irradiated by the nanoparticle is responsible
for the effects in iron chelation, it remains the question of what are the
changes induced by the irradiated power that lead to the slower release
50
40
∆Fe released (%)
30
20 3 hours exposure
ω × B = 190 Ts–1
10
0
(a) 250 500 1000 2000
f (kHz)
50 3 hours exposure
500 kHz
40
∆Fe released (%)
30
20
10
(b) 0 10 20 30 40 50 60
B (µT)
pH 2 buf.
Exposed
Control
0.25% vol
2 hours
Stirring
FIGURE 8.8 The threefold point act as hydrophilic terminals essential to pro-
tein solubility. Surface protonation and a sudden release of iron via pH reduction
and iron reducing agents lead to blocking of the negative terminals and protein
precipitation. This effect is quenched in solutions exposed to RF magnetic fields.
Effects of RF Magnetic Fields on Iron Release and Cage Proteins ◾ 241
0.08
DMSA
∆OD 562 nm
0.04
0.00
(a)
0.08
∆OD 480 nm
6-OHDA
0.04
0.00
(b) 0 2 4 6 8 10 12 14 16 18 20 22 24
Time (hours)
FIGURE 8.9 Changes in the optical absorption after the addition of (a) meso
2,3-dimercaptosuccinic acid (DMSA) and (b) 6-hydroxydopamine (6-OHDA)
between control and samples exposed for 5 hours to fields of 1 MHz and 30 μT.
ferritin solution are uptaken by the protein, there is an increase in the opti-
cal absorption of the solution. The changes in OD at 310 and 420 nm have
been calibrated as ε = 2745 and 550 per mol of iron uptaken and centime-
ter of solution, respectively.48,49 The temperature of the samples, measured
during and after the field exposure and during the titration measure-
ments, is once again equal for control and exposed samples within 0.1°C.
Measurements are done after (not during) field exposure.
As it was the case for iron release (Section 8.4.2), we define the mag-
netic field effect on the change of iron uptaken by the protein as ΔFeuptake ≡
(Feup|control − Feup|exposed)/Feup|control, where Feup|control and Feup|exposed are the
concentrations of iron uptaken 1 hour after adding the Fe (II) ions in con-
trol and exposed samples, respectively. We found that the ability of ferritin
to oxidize and store iron is reduced after being exposed to RF magnetic
fields, i.e., ΔFeuptake > 0. The change is a function of the molecular con-
centrations, exposure time and ω·B product (Figures 8.10 and 8.11). The
iron absorbed in 3.5 μM ferritin solutions exposed to field of 30 μT (H =
25 Am–1) at 1 MHz for five hours (total energy released ~ 35 JM–1), is
20 ± 10% smaller than for control samples for added iron concentrations
between 0.25 and 1 mM. However, for the same conditions in apoferritin
(protein without inner ferrihydrite nanoparticle), the effect is not signifi-
cant: ΔFeuptake (apoferritin) = 1 ± 4%. The effect depends on the relative fer-
ritin-iron concentrations, and it has not statistical significance for large Fe
(II) concentrations (~1 mM). The fact that there is no effect in apoferritin
40 3 µM ferritin
1 µM ferritin
3 µM apoferritin
Decrease in iron uptake (%)
30
20
10
250 µM 500 µM 1 mM
Fe2+ concentration
FIGURE 8.10 Changes in iron uptake after exposure to magnetic field as a func-
tion of the chemical concentrations of iron cations and ferritin.
Effects of RF Magnetic Fields on Iron Release and Cage Proteins ◾ 245
ω × B = 190 Ts–1
30 2 hours exposure
10
0
100 1000
f (kHz)
70
60
∆Fe release (%)
50
Ferrozine
40 chelation
30
20
10
0
(a) 0 5 10 15
Time (hours)
0
–10
2 hours exposure
–20
ω × H = 190 Ts–1
∆Fe release (%)
–30
–40
–50
–60
–70
FIGURE 8.12 (a) The effect has a maximum just 20 min after the addition of
the iron chelator (note that the magnetic field exposure is stopped previous to
the measurement) and then decays slowly over several hours. (b) Effect of RF
magnetic fields in the iron release from cationized ferritin when using ferrozine
as chelating agent. The change has the opposite sign than for natural ferritin with
negatively charged threefold symmetry points.
248 ◾ Biomagnetics
We find that the exposure to the same fields that resulted in a factor 2
slower release rate in ferritin (500 kHz to 2MHz; 60–15 μT), now result
in faster release rates from exposed ferritin. Furthermore, the effect now
extends over longer periods of time, with a roughly constant increase of 30
to 50 percent one day after the magnetic field exposure has been stopped
and the ferrozine added. Following our initial argument, we would con-
clude that the charge redistribution induced by the magnetic effect has
now the effect of making it easier for the negative ferrozine to penetrate
the proteic cage across the now neutralized carboxyl groups.
700
600
λ Emission 1
500 4
10
40
400 80
Y 120
300 160
200
300 400 500 600 240
(a) λ Excitation 280
500
Fluorescence polarization (mP)
400
300
200
100
0
0 500 1000 1500
(b) Iron content (ions)
• Increase in the iron chelation rate for cationized ferritin when using
the same molecules; opposite effect to standard protein.
done for example on comet analysis for DNA strands, which are not a
direct measurement of radical lifetime, have little reproducibility and
many variables with a complex procedure.
Another consideration when using RF, rather than DC magnetic fields,
is that once the period of the magnetic field is comparable with the life-
time of the free radical, the band split effect decreases as the frequency
increases. The lifetime of free radicals is for example 4 μsec (→ 40 kHz)
for a free radical involved in Fenton chemistry.38,53 If our effect was due
to changes in the Fenton reaction, as we increase the frequency from
250 kHz to 2 MHz, the effect should quickly decrease because of (1) the
reduced averaged field over the lifetime of the radical and (2) the fields
used at 250 kHz are 4 times higher than those at 2 MHz: from 60 to 15 μT.
In fact, at frequencies of the order of 1 MHz, the averaged field over the
radical lifetime would be zero, and no effect would be expected. However,
we do measure an effect, and it remains constant over the frequency range.
We think that if the effect was due to changes in the Fenton chemistry, the
use of pH buffers would surely reduce it because the buffer would control
the oxidation state of iron. We see in our measurements that it is not the
case. Finally, the addition of chelating agents is done after the exposure.
Our measurements are performed at frequencies of 0.25 to 2 MHz,
but ferrihydrite nanoparticles can have small relaxation times down to
10 psec.14 In theory, the described mechanism could vary with ω·B up to
much higher frequencies, with a dependence only on B once the period of
the field is smaller than the relaxation time of the nanoparticle.36 However,
measurements at higher frequencies would be necessary to asses this
hypothesis and the effects at higher frequencies. In the case of physiologi-
cal experiments, the iron/ferritin levels together with the frequency and
amplitude of the magnetic (not only electric) field should be considered.
From this calculation, it is easy to see that the energy loss by the nanopar-
ticle inside ferritin is unlikely to lead to a thermal effect at the concentra-
tions usually found in the body, even at relatively large magnetic fields
and frequencies. This calculation is nevertheless independent of the effects
that the loss power may have on the proteic cage itself or on whether our
suggested mechanism is indeed the correct explanation for the observed
effects (see Figure 8.14).
We do not know what the implications could be. For example, our
estimates calculate that to denaturalize human spleen ferritin we would
need fields of the order of 100 mT at frequencies of 100 kHz operating for
about 15 min, but these fields are one order of magnitude higher than the
252 ◾ Biomagnetics
Experiment
10–6
102 103 104 105 106 107 108
f (kHz)
FIGURE 8.14 Differences in the effect of iron releasing agent using the time scale
for a ferrihydrite nanoparticle with τN ~ 0.13 nsec. The denaturalization limit
assumes that all the power dissipated is absorbed by the protein, which is not
likely and would need to be probed experimentally.
maintained, but below the relaxation time of the nanoparticle only a field
dependence would be expected, limiting the experiments we could do to
confirm the mechanism. Given the characteristic time of the ferrihydrite
nanoparticle, which can be as low as 10 to 100 psec in some ferritins but
depends on many factors, even at higher frequencies the power dissipa-
tion would theoretically remain constant as long as the field amplitude
was maintained. That is, in principle, the effect at 1 μT and 1 GHz would
be approximately equivalent to also 1 μT at 10 GHz, but it could be similar
to 10 mT at 100 kHz. Of course, at those frequencies other factors, such as
molecular vibrations, would also play a role and may change the dissipa-
tion and absorption mechanisms. Furthermore, experiments in vitro with
high ferritin concentrations may differ very significantly from tests in vivo
and biological effects due to proteic relaxation, chemical buffer, and other
environmental considerations.
ACKNOWLEDGMENTS
We thank Bioelectromagnetics for the permission to reproduce ideas and
results as reported in References 11,13.
REFERENCES
1. Bavrnes, F. S. 2005. Mechanisms for electric and magnetic fields effects on
biological cells. IEEE Transactions on Magnetics 41: 4219–4224.
2. Mertz, W. 1981. The essential trace-elements. Science 213: 1332–1338.
3. Beard, J. L. 2001. Iron biology in immune function, muscle metabolism and
neuronal functioning. Journal of Nutrition 131: 568S–579S.
4. Levi, S., Corsi, B., Bosisio, M., Invernizzi, R., Volz, A., Sanford, D.,
Arosio, P., and Drysdale, J. 2001. A human mitochondrial ferritin encoded
by an intronless gene. Journal of Biological Chemistry 276: 24,437–24,440.
5. Arosio, P., and Levi, S. 2002. Ferritin, iron homeostasis, and oxidative dam-
age. Free Radical Biology and Medicine 33: 457–463.
6. Carrondo, M. A. 2003. Ferritins, iron uptake and storage from the bacterio-
ferritin viewpoint. Embo Journal 22: 1959–1968.
7. Kirschvink, J. L. 1996. Microwave absorption by magnetite: A possible
mechanism for coupling nonthermal levels of radiation to biological sys-
tems. Bioelectromagnetics 17: 187–194.
8. Kirschvink, J. L., Kobayashikirschvink, A., and Woodford, B. J. 1992.
Magnetite Biomineralization in the Human Brain. Proceedings of the
National Academy of Sciences of the United States of America 89: 7683–7687.
9. Cranfield, C. G., Weiser, H. G., and Dobson, J. 2003. Exposure of magnetic
bacteria to simulated mobile phone-type RF radiation has no impact on
mortality. IEEE Transactions on Nanobioscience 2: 146–149.
254 ◾ Biomagnetics
10. Cranfield, C., Wieser, H. G., Al Madan, J., and Dobson, J. 2003. Preliminary
evaluation of nanoscale biogenic magnetite-based ferromagnetic trans-
duction mechanisms for mobile phone bioeffects. IEEE Transactions on
Nanobioscience 2: 40–43.
11. Cespedes, O., and Ueno, S. 2009. Effects of radio frequency magnetic fields
on iron release from cage proteins. Bioelectromagnetics 30: 336–342.
12. Cespedes, O., Inomoto, O., Kai, S., and Ueno, S. 2009. Effects of cationiza-
tion and 6-hydroxydopamine on the reduced iron release rates from ferri-
tin by radio-frequency magnetic fields. IEEE Transactions on Magnetics 45:
4865–4868.
13. Cespedes, O., Inomoto, O., Kai, S., Nibu, Y., Yamaguchi, T., Sakamoto, N.,
Akune, T., Inoue, M., Kiss, T., and Ueno, S. 2010. Radio frequency mag-
netic field effects on molecular dynamics and iron uptake in cage proteins.
Bioelectromagnetics 31: 311–317.
14. Allen, P. D., St Pierre, T. G., Chua-anusorn, W., Strom, V., and Rao, K. V.
2000. Low-frequency low-field magnetic susceptibility of ferritin and
hemosiderin. Biochimica Et Biophysica Acta-Molecular Basis of Disease
1500: 186–196.
15. Pankhurst, Q. A., Connolly, J., Jones, S. K., and Dobson, J. 2003. Applications
of magnetic nanoparticles in biomedicine. Journal of Physics D-Applied
Physics 36: R167–R181.
16. Jordan, A., Scholz, R., Wust, P., Fahling, H., and Felix, R. 1999. Magnetic
fluid hyperthermia (MFH): Cancer treatment with AC magnetic field
induced excitation of biocompatible superparamagnetic nanoparticles.
Journal of Magnetism and Magnetic Materials 201: 413–419.
17. Hergt, R., Andra, W., d’Ambly, C.G., Hilger, I., Kaiser, W.A., Richter, U.,
and Schmidt, H. G. 1998. Physical limits of hyperthermia using magnetite
fine particles. IEEE Transactions on Magnetics 34: 3745–3754.
18. Theil, E. C. 2011. Ferritin protein nanocages use ion channels, catalytic
sites, and nucleation channels to manage iron/oxygen chemistry. Current
Opinion in Chemical Biology 15: 304–311.
19. Jameson, G. N. L., Jameson, R. F., and Linert, W. 2004. New insights into
iron release from ferritin: direct observation of the neurotoxin 6-hydroxy-
dopamine entering ferritin and reaching redox equilibrium with the iron
core. Organic & Biomolecular Chemistry 2: 2346–2351.
20. Gallois, B., d’Estaintot, B. L., Michaux, M. A., Dautant, A., Granier, T.,
Precigoux, G., Soruco, J. A., Roland, F., Chavas Alba, O., Herbas, A., and
Crichton, R. R. 1997. X-ray structure of recombinant horse L-chain apo-
ferritin at 2.0 angstrom resolution: Implications for stability and function.
Journal of Biological Inorganic Chemistry 2: 360–367.
21. Levi, S., Luzzago, A., Cesareni, G., Cozzi, A., Franceschinelli, F., Albertini,
A., and Arosio, P. 1988. Mechanism of ferritin iron uptake-activity of the
H-chain and deletion mapping of the ferro-oxidase site—A study of iron
uptake and ferro-oxidase activity of human-liver, recombinant H-chain
ferritins, and of 2 H-chain deletion mutants. Journal of Biological Chemistry
263: 18,086–18,092.
Effects of RF Magnetic Fields on Iron Release and Cage Proteins ◾ 255
22. Harrison, P. M., and Arosio, P. 1996. Ferritins: Molecular properties, iron
storage function and cellular regulation. Biochimica Et Biophysica Acta-
Bioenergetics 1275: 161–203.
23. Arosio, P., Ingrassia, R., and Cavadini, P. 2009. Ferritins: A family of mol-
ecules for iron storage, antioxidation and more. Biochimica Et Biophysica
Acta-General Subjects 1790: 589–599.
24. Winterbourn, C. C. 1995. Toxicity of iron and hydrogen peroxide: The
Fenton reaction. Toxicology Letters 82-3: 969–974.
25. Valko, M., Morris, H., and Cronin, M. T. D. 2005. Metals, toxicity and oxi-
dative stress. Current Medicinal Chemistry 12: 1161–1208.
26. Stohs, S. J., and Bagchi, D. 1995. Oxidative mechanisms in the toxicity of
metal-ions. Free Radical Biology and Medicine 18: 321–336.
27. Pankhurst, Q., Hautot, D., Khan, N., and Dobson, J. 2008. Increased levels
of magnetic iron compounds in Alzheimer’s disease. Journal of Alzheimers
Disease 13: 49–52.
28. Everett, J., Cespedes E., Shelford, L. R., Exley, C., Collingwood, J. F.,
Dobson, J., van der Laan, G., Jenkins, C. A., Arenholz, E., and Telling,
N. D. 2014. Evidence of redox-active iron formation following aggregation
of ferrihydrite and the Alzheimer’s disease peptide beta-amyloid. Inorganic
Chemistry 53: 2803–2809.
29. Bartzokis, G., Tishler, T. A., Shin, I. S., Lu, P. H., and Cummings, J. L. 2004.
Redox-Active Metals in Neurological Disorders, vol. 1012, Annals of the New
York Academy of Sciences. LeVine, S. M., Connor, J. R., and & Schipper,
H. M. Academy of Sciences, New York: 224–236.
30. Rottkamp, C. A., et al. 2001. Redox-active iron mediates amyloid-beta tox-
icity. Free Radical Biology and Medicine 30: 447–450.
31. Zecca, L., Youdim, M. B. H., Riederer, P., Connor, J. R, and Crichton, R. R.
2004. Iron, brain ageing and neurodegenerative disorders. Nature Reviews
Neuroscience 5: 863–873.
32. Lee, J.-H., Huh, Y.-M., Jun, Y.-W., Seo, J.-W., Jang, J.-T., Song, H.-T., Kim,
S., Cho, E.-J., Yoon, H.-G., Suh, J.-S., and Cheon, J. 2007. Artificially engi-
neered magnetic nanoparticles for ultra-sensitive molecular imaging.
Nature Medicine 13: 95–99.
33. Zhang, Y., Pilapong, C., Guo, Y., Zhenlian, L., Cespedes, O., Quirke, P., and
Zhou, D. 2013. Sensitive, simultaneous quantitation of two unlabeled DNA
targets using a magnetic nanoparticle-enzyme sandwich assay. Analytical
Chemistry 85: 9238–9244.
34. Neuberger, T., Schopf, B., Hofmann, H., Hofmann, M., and von Rechenberg,
B. 2005. Superparamagnetic nanoparticles for biomedical applications:
Possibilities and limitations of a new drug delivery system. Journal of
Magnetism and Magnetic Materials 293: 483–496.
35. Brem, F., Stamm, G., and Hirt, A. M. 2006. Modeling the magnetic behavior
of horse spleen ferritin with a two-phase core structure. Journal of Applied
Physics 99: 123906.
36. Rosensweig, R. E. 2002. Heating magnetic fluid with alternating magnetic
field. Journal of Magnetism and Magnetic Materials 252: 370–374.
256 ◾ Biomagnetics
37. Gilles, C., Bonville, P., Wong, K. K. W., and Mann, S. 2000. Non-Langevin
behaviour of the uncompensated magnetization in nanoparticles of artifi-
cial ferritin. European Physical Journal B 17: 417–427.
38. Cheeseman, K. H., and Slater, T. F. 1993. An introduction to free-radical
biochemistry. British Medical Bulletin 49: 481–493.
39. Chou, C. K., Bassen, H., Osepchuk, J., Balzano, Q., Petersen, R., Meltz,
M., Cleveland, R., Lin, J. C., and Heynick, L. 1996. Radio frequency
electromagnetic exposure: Tutorial review on experimental dosimetry.
Bioelectromagnetics 17: 195–208.
40. Stefanini, S., Cavallo, S., Wang, C. Q., Tataseo, P., Vecchini, P., Giartosio, A.,
and Chiancone, E. 1996. Thermal stability of horse spleen apoferritin and
human recombinant H apoferritin. Archives of Biochemistry and Biophysics
325: 58–64.
41. Yau, S. T., Petsev, D. N., Thomas, B. R., and Vekilov, P. G. 2000. Molecular-
level thermodynamic and kinetic parameters for the self-assembly of
apoferritin molecules into crystals. Journal of Molecular Biology 303:
667–678.
42. Funk, F., Lenders, J. P., Crichton, R. R., and Schneider, W. 1985. Reductive
mobilization of ferritin iron. European Journal of Biochemistry 152: 167–172.
43. Santambrogio, P., Levi, S., Cozzi, A., Corsi, B., and Arosio, P. 1996. Evidence
that the specificity of iron incorporation into homopolymers of human fer-
ritin L- and H-chains is conferred by the nucleation and ferroxidase cen-
tres. Biochemical Journal 314: 139–144.
44. Fahn, S., and Cohen, G. 1992. The oxidant stress hypothesis in Parkinsons-
disease—Evidence supporting it. Annals of Neurology 32: 804–812.
45. Blum, D., Torch, S., Lamberg, N., Nissou, M. F., Benabid, A. L., Sadoul, R.,
and Verna, J. M. 2001. Molecular pathways involved in the neurotoxicity of
6-OHDA, dopamine and MPTP: Contribution to the apoptotic theory in
Parkinson’s disease. Progress in Neurobiology 65: 135–172.
46. Bove, J., Prou, D., Perier, C., and Przedborski, S. 2005. Toxin-induced mod-
els of Parkinson’s disease. Journal of the American Society for Experimental
NeuroTherapeutics 2: 484–494.
47. Li, P., Sage, J. T., and Champion, P. M. 1992. Probing picosecond processes
with nanosecond lasers—Electronic and vibrational-relaxation dynamics
of heme-proteins. Journal of Chemical Physics 97: 3214–3227.
48. Macara, I. G., Hoy, T. G., and Harrison, P. M. 1973. Formation of ferritin
from apoferritin—Inhibition and metal ion-binding studies. Biochemical
Journal 135: 785–789.
49. Macara, I. G., Hoy, T. G., and Harrison, P. M. 1973. Formation of ferritin
from apoferritin—Catalytic action of apoferritin. Biochemical Journal 135:
343–348.
50. Theil, E. C., Takagi, H., Small, G.W., He, L., Tipton, A. R., and Danger, D.
2000. The ferritin iron entry and exit problem. Inorganica Chimica Acta
297: 242–251.
Effects of RF Magnetic Fields on Iron Release and Cage Proteins ◾ 257
51. Danon, D., Goldstein, L., Marikovs, Y., and Skutelsk, E. 1972. Use of cat-
ionized ferritin as a label of negative charges on cell surfaces. Journal of
Ultrastructure Research 38: 500–510.
52. vanDijk, B., Gast, P., and Hoff, A. J. 1996. Control of radical pair lifetime by
a switched magnetic field. Physical Review Letters 77: 4478–4481.
53. Nedoloujko, A., and Kiwi, J. 1997. Transient intermediate species active
during the Fenton-mediated degradation of quinoline in oxidative media:
Pulsed laser spectroscopy. Journal of Photochemistry and Photobiology A:
Chemistry 110: 141–148.
Chapter 9
Safety Aspects
of Magnetic and
Electromagnetic Fields
Sachiko Yamaguchi-Sekino,
Tsukasa Shigemitsu, and Shoogo Ueno
CONTENTS
9.1 Introduction 260
9.2 Mechanisms for Biological Interaction 262
9.2.1 Static Magnetic Field 262
9.2.2 Gradient (Time-Varying) Magnetic Field 263
9.2.3 Radio Frequency Electromagnetic Field 264
9.3 Biological Effects Related to MRI 265
9.3.1 Static Magnetic Field 266
9.3.1.1 In Vivo Studies 266
9.3.1.2 In Vitro Studies 269
9.3.1.3 Human Experimental and Epidemiological
Studies 270
9.3.2 Gradient (Time-Varying) Magnetic Field 272
9.3.3 Radio Frequency Electromagnetic Field 273
9.3.4 Possible Health Effect Related to MRI Operation 274
9.4 Exposure Guidelines Related to MRI 276
9.4.1 Static Magnetic Field 277
9.4.2 Gradient (Time-Varying) Magnetic Field 278
9.4.3 Radio Frequency Electromagnetic Field 279
9.4.4 Movement-Related Magnetic Field 280
259
260 ◾ Biomagnetics
9.1 INTRODUCTION
The magnetic resonance imaging (MRI) was introduced into the diagnostic
imaging technology in medicine in the early 1980s. It originally comes from
the technique of nuclear magnetic resonance (NMR). In 1970s, Damadian
observed the differences of the relaxation times between tumor and nor-
mal tissues and proposed the use of NMR as imaging for the detection of
cancer (Damadian 1971). In Japan, during the 1970s, using NMR, Abe et
al. proposed and measured noninvasively detection of the biological image
information by magnetic focusing method (Abe et al. 1974).
The basic principle and the technical development of imaging forma-
tion using NMR were proposed by Lauterbur, who developed the MRI
(Lauterbur 1973). The Nobel Prize in Medicine and Physiology was
awarded in 2003 to Lauterbur and Mansfield for the fundamental appli-
cation of a static magnetic field in combination with a gradient (time-
varying) magnetic field and the development of the image acquisition and
processing (Mansfield and Maudsley 1977). In this way, the root of MRI is
known as nuclear magnetic resonance. To avoid alarming the public and
medical experts, the word “nuclear” was deleted from the term magnetic
resonance, although it has nothing to do with radioactivity.
The operation of MRI basically utilizes three different types of elec-
tromagnetic field, the strong static magnetic field, the rapidly changing
gradient (time-varying) magnetic field, and the radio frequency electro-
magnetic field. The static magnetic field is the main magnetic field in MRI.
It aligns the proton spins and generates a total magnetization in the human
body. The static magnetic field is responsible for a measure of the proton
density. It is usually generated by a strong superconducting magnet. The
increase in the static magnetic field strength goes toward improvements in
image resolution. Today, the commercially available MRI system typically
has a static magnetic field strength of up to 3 T. The gradient magnetic
Safety Aspects of Magnetic and Electromagnetic Fields ◾ 261
ϕ = ν ∙ B ∙ d ∙ sin θ
is the same as the external field. Human and animal bodies do not sig-
nificantly perturb the field. The main interaction of magnetic fields is the
Faraday induction of electric fields and associated currents in the tissues.
Key features of dosimetry for exposure of humans to low-frequency mag-
netic fields include
J = π · r · f · σ · B0
This means that the induced currents are proportional to the loop
radius r and tissue conductivity σ. A well-known biological effect of mag
netic fields in this range is the induction of visual sensations called
magnetophosphene.
2
σE
SAR =
2ρ
In parallel with the brief introduction of recent studies in each section, the
evaluation was conducted with the help of authoritative reviews by well-
recognized organizations such as World Health Organization (WHO)
and the International Commission on Non-Ionizing Radiation Protection
(ICNIRP).
to fields greater than about 1 T (0.1 T in larger animals) will induce flow
potentials around the heart and major blood vessels, but the physiologi-
cal consequences of this remain unclear. Several hours of exposure to
very high flux densities of up to 8 T in the heart region did not result in
any cardiovascular effects in pigs. In rabbits, short and long exposures to
fields ranging from geomagnetic levels to the millitesla range have been
reported to affect the cardiovascular system, although the evidence is not
strong” (WHO 2006a, p. 5).
serum proteins and hormone levels, body, and skin temperature (WHO
2006a).
The document of the WHO stated that “The results do not indicate that
there are effects of static magnetic field exposure on neurophysiological
responses and cognitive functions in stationary volunteers, nor can they
rule out such effects. A dose-dependent induction of vertigo and nausea
was found in workers, patients and volunteers during movement in static
fields greater than about 2 T. One study suggested that eye-hand coordina-
tion and near visual contrast sensitivity are reduced in fields adjacent to
a 1.5 T MRI unit. Occurrence of these effects is likely to be dependent on
the gradient of the field and the movement of the subject. A small change
in blood pressure and heart rate was observed in some studies, but were
in the range of normal physiological variability. There is no evidence of
effects of static magnetic fields on other aspects of cardiovascular physi-
ology, or on serum proteins and hormones. Exposure to static magnetic
fields of up to 8 T does not appear to induce temperature changes in
humans” (WHO 2006a, p. 7).
In the past, several epidemiological studies have been conducted to
examine mortality and cancer incidence among workers exposed to static
magnetic fields at aluminum reduction and chloralkali plants. Historically,
Marsh et al. presented the occupational exposure study (Marsh et al. 1982).
This study with 320 male workers carried out the occupational exposure
to static magnetic fields ranging from about 3 to 15 mT generated from
large electrolytic cells. The control group was 186 male workers. Although
the cohort was small, they found that there are significant increases in
heart disease and cancer. Barregard et al. carried out very limited occupa-
tional study on workers exposed to static magnetic field from 4 to 29 mT
(average 14 mT) (Barregard et al. 1985). They studied mortality and can-
cer incidence among workers in a Swedish chloralkali plant where about
100 kA direct currents are used to produce chlorine. The exposed group
comprised of 157 men employed between 1951 and 1983, and their mortal-
ity has been compared with that of Swedish men using calendar-year and
age-specific mortality rates. Cancer incidence has been compared with the
expected incidences among Swedish men. They reported that there is no
increased mortality among the exposed men. No excess incidence of can-
cer is found.
Regarding the human and epidemiological studies, a document stated
that “Increased risks of various cancers, e.g., lung cancer, pancreatic can-
cer, and haematological malignancies, were reported. But results were
272 ◾ Biomagnetics
MRI for 1 h followed by the movement of their heads for 16 sec. The pos-
tural body sway was expressed in sway path length, sway area, and sway
velocity. The healthy volunteers performed two tasks, standing with eyes
closed and feet in parallel and then in tandem position, after standardized
head movements in a sham, low-exposure (stray magnetic fields: 0.24 T
with time-varying magnetic field of 0.49 T/sec) and high-exposure con-
ditions (0.37 T with 0.70 T/sec). The results show that sway path length,
sway area, and the velocity of body movements were significantly higher
with higher static magnetic fields. They commented that the investigation
of the practical safety implications of this finding for surgeons and others
working near MRI scanner is needed.
which covers the prevention of nervous system functions. The ICNIRP set
the basic restrictions in terms of the induced electric field strength in the
body to prevent perception of nerve stimulation.
Guidelines for exposure above 100 kHz are covered in another docu-
ment (ICNIRP 1998). Between 100 kHz and 10 GHz, basic restrictions
consider the SAR to prevent whole-body and localized tissue heating.
For frequencies from 10 to 300 GHz, basic restrictions are set in terms
of power density to prevent heating effects in tissue at or near the body
surface (ICNIRP 1998).
As mentioned in the ICNIRP document, induced electric field strength
and SAR cannot be measured directly. So the reference level was set by
ICNIRP. For practical exposure assessment purposes, the ICNIRP has
provided the reference level in terms of the strength of electric and mag-
netic fields. The reference levels are obtained from the basic restrictions by
mathematical modeling.
beyond 8 T.
d Because of potential indirect adverse effects, ICNIRP recognizes that
2009b). The acute exposure of the general public should not exceed 400 mT
(any part of the body). The term “general public” refers to the entire pop-
ulation including individuals of all ages and varying health status. For
occupational exposure, the ICNIRP has a 2-T limit for the head and trunk
and an 8-T limit for limbs in controlled situations.
Occupational Exposure
CNS tissue of the head 1–10 Hz 0.5/f
10–25 Hz 0.05
25–400 Hz 2 × 10–3f
400 Hz–3 kHz 0.8
3 kHz–10 MHz 2.7 × 10–4f
All tissues of head and body 1 Hz–3 kHz 0.8
3 kHz–10 MHz 2.7 × 10–4f
range between 100 kHz and 10 MHz, basic restrictions are provided on
both current density and SAR (Table 9.4). As shown in Table 9.5, basic
restrictions between 10 and 300 GHz are provided on power density to
prevent excessive heating in tissue at or near the body surface (ICNIRP
1998).
Reference levels for occupational exposure are shown in Table 9.6.
In the frequency range between 100 kHz and 300 GHz, the exposure
basic restriction and reference level are still based on the 1998 ICNIRP
guidelines until there is a revision of the guidelines for the high-
frequency portion of the electromagnetic spectrum between 100 kHz
and 300 GHz.
According to guidelines (ICNIRP 1998, 2010), when the reference levels
are exceeded, assessments with measurement and calculation are neces-
sary to estimate whether the basic restrictions are exceeded.
TABLE 9.4 Basic Restrictions for Time-Varying Electric, Magnetic, and Electromagnetic
Fields up to 10 GHz
Current Density Localized Localized
for Head and Whole-Body SAR (Head SAR
Exposure Frequency Trunk (mA m–2) Average SAR and Trunk) (Limbs)
Characteristics Range (rms) (W kg–1) (W kg–1) (W kg–1)
Occupational Up to 1 Hz 40 – – –
exposure 1–4 Hz 40/f – – –
4 Hz–1 kHz 10 – – –
1–100 kHz f/100 – – –
100 kHz– f/100 0.4 10 20
10 MHz
10 MHz– – 0.4 10 20
10 GHz
General Up to 1 Hz 8 – – –
public 1–4 Hz 8/f – – –
exposure 4 Hz–1 kHz 2 – – –
1–100 kHz f/500 – – –
100 kHz– f/500 0.08 2 4
10 MHz
10 MHz– – 0.08 2 4
10 GHz
Source: Reproduced from ICNIRP, Health Physics 74(4): 494–522, 1998.
Note: 1. f is the frequency in hertz.
2. Because of electrical inhomogeneity of the body, current densities should be aver-
aged over a cross section of 1 cm2 perpendicular to the current direction.
3. For frequencies up to 100 kHz, peak current density values can be obtained by
multiplying the rms value by √2 (~1.414). For pulses of duration tp, the equiva-
lent frequency to apply in the basic restrictions should be calculated as f = 1/(2 tp).
4. For frequencies up to 100 kHz and for pulsed magnetic fields, the maximum cur-
rent density associated with the pulses can be calculated from the rise/fall times
and the maximum rate of change of magnetic flux density. The induced current
density can then be compared with the appropriate basic restriction.
5. All SAR values are to be averaged over any 6-min period.
6. Localized SAR averaging mass is any 10 g of contiguous tissue; the maximum
SAR so obtained should be the value used for the estimation of exposure.
7. For pulses of duration tp, the equivalent frequency to apply in the basic restric-
tions should be calculated as f = 1/(2 tp). Additionally, for pulsed exposures in the
frequency range 0.3 to 10 GHz and for localized exposure of the head, in order to
limit or avoid auditory effects caused by thermoelastic expansion, an additional
basic restriction is recommended. This is that the SA should not exceed 10 mJ kg–1
for workers and 2 mJ kg–1 for the general public, averaged over 10 g tissue.
282 ◾ Biomagnetics
With respect to the application of the SAR levels defined in Table 9.8,
the following points should be taken into account (ICNIRP 2004):
TABLE 9.9 Basic Restrictions for Body Temperature Rise and Partial-Body Temperature
Spatially Localized Temperature Limits
Rise of Body Core
Operating Mode Temperature (°C) Head (°C) Trunk (°C) Extremities (°C)
Normal 0.5 38 39 40
Controlled 1 38 39 40
Restricted >1 >38 >39 >40
Source: Reproduced from ICNIRP, Health Physics 87: 197–216, 2004.
Safety Aspects of Magnetic and Electromagnetic Fields ◾ 287
TABLE 9.10 Summary of Recommended Limits for Patient Exposure for Static,
Gradient, and Radiofrequency during MRI Procedure
Gradient Radiofrequency Field
dB/dt as a Temperature Limits (°C)
Static Percentage of Maximum
Magnetic Median Core
Operating Field in Perception Temperature
Mode Bore (T) Threshold (%) Rise (°C) Head Trunk Extremities
Normal <4 80 0.5 38 39 40
Controlled <8 100 1 38 39 40
Experimental >8 100 >1 >38 >39 >40
288 ◾ Biomagnetics
the body and neck, (2) MRI examination of the body or extremities, (3)
escort or assist patient, and (4) other work contents. The highest exposure
was detected during MRI examination of the head or neck. The maximum
stray fields were 628 ± 28 mT at the edge of the bed from one institute
and 373 ± 2 mT at the edge of the bed from another institute. The reason
why the stray magnetic field is different between the two institutes is that
the sizes of the magnet and bore are different because they are from dif-
ferent suppliers.
Kännälä et al. examined the assessment of occupational exposure to
gradient magnetic fields and time-varying magnetic fields generated by
motion in nonhomogeneous static magnetic fields of two MRI scanners
(1 T open and 3 T conventional) (Kännälä et al. 2009). These magnetic
components can be measured simultaneously with an induction coil set
up detecting the time rate of change of magnetic flux density (dB/dt). As
a result, the highest motion-induced dB/dt was 0.7 T/sec for the 1-T MRI
scanner and 3 T/s for the 3-T MRI scanner when only the static magnetic
field was present.
Andreuccetti et al. proposed and tested the procedure for assessing
occupational exposure due to MRI gradient magnetic fields and movement-
induced effects in the static magnetic field (Andreuccetti et al. 2013). Their
procedure considered two exposures of low-frequency switched gradient
magnetic field and movement-induced effects in the static magnetic field.
The radio frequency electromagnetic field was not taken into consideration.
They tested in two 1.5-T whole-body MRI scanners and one 3-T head-only
MRI scanner and evaluated exposure due to switched gradient magnetic
fields in the location inside the magnet room where operators usually stay
during those particular medical procedures. Movement-induced effects
were evaluated considering the actual movements of volunteer operators
during work activity by measuring the perceived time-varying magnetic
field. The result was analyzed based on the ICNIRP 1998 and 2010 guide-
lines. Exposure to switched gradient magnetic fields in 1.5-T MRI scanner
mostly resulted in noncompliance with the ICNIRP 1998 occupational
levels, while at the same time, it showed as always compliant with the
ICNIRP 2010 ones. Movement-induced effects resulted potentially non-
compliant only in the case the operator moved the head inside the bore of
1.5-T MRI scanner.
There have been no epidemiological studies related to MRI operations.
It is needed and very important from the point of long-term human health
effects to conduct well-designed epidemiological studies of MRI workers.
Safety Aspects of Magnetic and Electromagnetic Fields ◾ 291
The ICNIRP published two guidelines, one for static magnetic field and
one for low-frequency electromagnetic fields (up to 100 kHz) (ICNIRP
2009b, 2010). Using these guidelines, the European Commission finally
proposed a Directive replacement. This newly proposed EU Directive
2013/35/EU covers all known direct biophysical effects and other indirect
effects caused by electromagnetic fields (up to 300 GHz) (EU Directive
2013/35/EU 2013). The Directive addresses short-term effects and does
not cover suggested long-term effects. It does not cover the risks resulting
from contact with live conductors. Finally, Member States are required to
have transposed EU Directive 2013/35/EU into their respective national
legislation framework by July 1, 2016.
The EU Directive 2013/35/EU covers exposure limit value and action
levels of static electric, static, magnetic, and time-varying electric, mag-
netic, and electromagnetic fields with frequencies up to 300 GHz. The fol-
lowing discussions are focused on the issues related to MRI.
The basic restrictions in the ICNIRP become “exposure limit values”
in the 2013 Directive and reference levels in the ICNIRP become “action
levels.” The Directive introduces two tiers of ELVs: sensory effects ELV
and health effects ELV.
1. Exposure limit values (ELVs) and action levels (ALs) in the frequency
range from 0 Hz to 10 MHz
a. ELVs
– ELVs below 1 Hz are limits for static magnetic field. For static
magnetic field, the 2013 Directive sets exposure limit values
identical to those in guidelines (ICNIRP 2009b).
– ELVs for external magnetic flux density from 0 to 1 Hz
– Sensory effects ELV is the ELV, 2 T for normal working
conditions and is related to vertigo and other physiologi-
cal effects related to disturbance of the human balance,
resulting from the movement in a static magnetic field.
– Health effects ELV is 8 T for controlled working conditions.
– ELVs for frequencies between 1 Hz and 10 MHz are limits
for induced electric fields in the body from exposure to time-
varying electric and magnetic fields.
Safety Aspects of Magnetic and Electromagnetic Fields ◾ 293
and for frequencies above 400 Hz, from the health effects
ELVs for internal electric field.
– High action levels (high ALs) are derived from the health effects
ELV for internal electric fields related to electric stimulation of
peripheral and autonomous nerve tissues in head and trunk.
2. ELVs and ALs in the frequency range from 100 kHz to 300 GHz.
a. ELVs
– Health effects ELVs between 100 kHz and 6 GHz are limits
for energy and power absorbed per unit mass of body tissue
generated from exposure to electric and magnetic fields.
– Sensory effects ELVs from 0.3 to 6 GHz are limits on absorbed
energy in a small mass tissue in the head from exposure to
electromagnetic fields.
– Health effects ELVs between 6 and 300 GHz are limits for
power density of an electromagnetic wave incident on the
body surface.
– Sensory effect ELVs from 0.3 to 6 GHz are related to avoid-
ing auditory effects caused by exposure of the head to pulsed
microwave radiation.
b. ALs for exposure to electric and magnetic fields shown in Table 9.12.
– ALs (B) are derived from the SAR or power density and ELVs
based on the thresholds related to internal thermal effects
caused by exposure to (external) electric and magnetic fields.
TABLE 9.12 ALs for Exposure to Electric and Magnetic Fields from 100 kHz
to 300 GHz
Electric Field Strength Magnetic Flux Density Power Density
Frequency Range ALs(E) [V m–1] (RMS) ALs(B) [μT] (RMS) ALs(S) [Wm–2]
100 kHz ≤ f < 1 MHz 6.1 × 102 2.0 × 106/f –
1≤ f < 10 MHz 6.1 × 108/f 2.0 × 106/f –
10 ≤ f < 400 MHz 61 0.2 –
400 MHz ≤ f < 2 GHz 3 × 10–3 f ½ 1.0 × 10–5 f ½ –
2 ≤ f < 6 GHz 1.4 × 102 4.5 × 10–1 –
6 ≤ f ≤ 300 GHz 1.4 × 102 4.5 × 10–1 50
Source: Reproduced from EU Directive 2013/35/EU, Official Journal of the European
Union 56: 1–21, L179, 2013.
Note: 1. f is the frequency expressed in hertz.
2. [ALs(E)]2 and [ALs(B)]2 are to be averaged over a 6-min period. For RF pulses,
the peak power density averaged over the pulse width shall not exceed
1000 times the respective ALs(S) value. For multifrequency fields, the analysis
shall be based on summation, as explained in the practical guides referred to in
Article 14.
3. ALs(E) and ALs(B) represent maximum calculated or measured values at the
workers’ body position. This results in a conservative exposure assessment and
automatic compliance with ELVs in all non-uniform exposure conditions. In
order to simplify the assessment of compliance with ELVs, carried out in accor-
dance with Article 4, in specific non-uniform conditions, criteria for the spatial
averaging of measured fields based on established dosimetry will be laid down in
the practical guides referred to in Article 14. In the case of a very localized source
within a distance of a few centimeters from the body, compliance with ELVs shall
be determined dosimetrically, case by case.
4. The power density shall be averaged over any 20 cm2 of exposed area. Spatial
maximum power densities averaged over 1 cm2 should not exceed 20 times the
value of 50 W m–2. Power densities from 6 to 10 GHz are to be averaged over any
6-min period. Above 10 GHz, the power density shall be averaged over any 68/
f 1.05-min period (where f is the frequency in gigahertz) to compensate for pro-
gressively shorter penetration depth as the frequency increases.
2. Given the state of the art, all technical and/or organizational mea-
sures have been applied.
3. Circumstances duly justify exceeding the ELVs.
4. Characteristics of the workplace, work equipment, or work practices
have been taken into account.
5. Employer demonstrates that workers are still protected against
adverse health effects and against safety risks.
296 ◾ Biomagnetics
9.6 CONCLUSION
The MRI is the most successful technique in medicine, particularly, in
clinical practices. In this chapter, we provided an overview and evaluation
of the biological and health effects resulting from three kinds of electro-
magnetic fields (static magnetic field, gradient [time-varying] magnetic
field, and radio frequency electromagnetic field) with the newest results,
highlighting ICNIRP guidelines of occupational exposures, and European
EMF Directives and its impact for MRI community. The electromagnetic
fields associated with MRI have been used in medical technology without
a concomitant risk of adverse health effects. Our scientific understanding
of the biological and health effects of electromagnetic field has gradually
improved from much research.
The WHO produced and planned three documents of EHC on the
possible health effects of exposure to (1) static, (2) extremely low fre-
quency (ELF), and (3) radio frequency electromagnetic fields. After
publication of EHC documents, the WHO opened their research rec-
ommendations of MRI-related electromagnetic fields (WHO 2006b,
2010). The available data on possible effects of MRI examinations rel-
evant to the safety of patients and worker are not sufficient to draw
conclusions.
Since the introduction of MRI, patients, MRI staff, and other work-
ers exposed to electromagnetic field associated with MRI have increased.
Protection of workers requires occupational exposure assessment in MRI-
working environments. In addition, in MRI environments, there has been
no epidemiological research to assess the possible long-term health effects
in patients, volunteers, and medical and MRI staffs. Because the health
risk assessment for MRI procedures is incomplete, it is needed to address
Safety Aspects of Magnetic and Electromagnetic Fields ◾ 297
ACKNOWLEDGMENTS
We thank Dr. Ikehata of Railway Technical Research Institute, Dr. Okano
of Hakuju Institute for Health and Science, Dr. Sakurai of Gifu Pharma
ceutical University, Dr. Yamazaki of Central Research Institute of Electric
Power Industry, and Dr. Watanabe of NICT for their kind support on this
manuscript.
298 ◾ Biomagnetics
REFERENCES
Abe, Z., Tanaka, K., Hotta, K., and Imai, M. 1974. Noninvasive measurements of
biological information with application of nuclear magnetic resonance. In
Biological and Clinical Effects of Low Magnetic and Electric Fields, Llaurado,
J. G., Sances, A., and Battocletti, J. H., eds. Charles C. Thomas, Springfield,
Ill.: 295–317.
Ahlbom, A., Green, A., Kheifets, L., Savitz, D., and Swerdlow, A. 2004.
Epidemiology of health effects of radiofrequency exposure. Environmental
Health Perspectives 112: 1741–1754.
Andreuccetti, D., Contessa, G. M., Falsaperia, R., Lodato, R., Pinto, R., Zoppetti,
N., and Rossi, P. 2013. Weighted-peak assessment of occupational exposure
due to MRI gradient fields and movement in a nonhomogeneous static mag-
netic field. Medical Physics 40(1): 011910.
Australian Radiation Protection and Nuclear Safety Agency (ARPANSA). 2014.
Review of radiofrequency health effects research—Scientific literature 2000–
2012. Technical Report Series No. 164:1–68.
Baan, R., Gross, Y., Lauby-Secretan, B., El Ghissassi, F., Bouvard, V., Benbrahim-
Tallaa Guha, N., Islami, F., Galicht, L., and Straif, K. 2011. Carcinogenicity
of radio-frequency electromagnetic fields. The Lancet Oncology 12(7): 624–626.
Barregard, L., Jarvholm, B., and Ungethum, E. 1985. Cancer among workers
exposed to strong static magnetic fields. The Lancet 11(8460): 892.
Bongers, S., Christopher, Y., Engels, H., Slottje, P., and Kromhout, H. 2014.
Retrospective assessment of exposure to static magnetic fields during pro-
duction and development of magnetic resonance imaging systems. Annals of
Occupational Hygiene 58(1): 85–102.
Börner, F., Brüggemeyer, H., Eggert, S., Fischer, M., Heinrich, H., Hentschel, K.,
and Neuschulz, H. 2011. Electromagnetic fields at workplaces: A new sci-
entific approach to occupational health and safety. Bundesministeriums für
Arbeit und Soziales.
Bradley, J. K., Nyekiova, M., Price, D. L., Lopez, L. D., and Grawley, T. 2007.
Occupational exposure to static and time-varying gradient magnetic fields
in MR units. Journal of Magnetic Resonance Imaging 26(5): 1204–1209.
Brix, G., Seebass, M., Hellwig, G., and Griebel, J. 2002. Estimation of heat trans
fer and temperature rise in partial-body regions during MR procedures:
An analytical approach with respect to safety considerations. Magnetic
Resonance Imaging 20: 65–76.
Cason, A. M., Kwon, B., Smith, J. C., and Houpt, T. A. 2009. Labyrinthectomy
abolishes the behavioral and neural response of rats to a high-strength static
magnetic field. Physiology & Behavior 97: 36–43.
Damadian, R. 1971. Tumor detection by nuclear magnetic resonance. Science 171:
1151–1153.
De Vocht, F., van Drooge, H., Engles, H., and Kromhout, H. 2006. Exposure,
health complaints and cognitive performance among employees of an MRI
scanners manufacturing department. Journal of Magnetic Resonance Imaging
23: 197–204.
Safety Aspects of Magnetic and Electromagnetic Fields ◾ 299
De Vocht, F., Stevens, T., Glover, P., Sunderland, A., Gowland, P., and Kromhout,
H. 2007. Cognitive effects of head-movements in stray fields generated by a 7
Tesla whole-body MRI magnets. Bioelectromagnetics 28(4): 247–255.
De Vocht, F., Muller, H., Engels, H., and Kromhout, H. 2009. Personal exposure
to static and time-varying magnetic fields during MRI system test procedure.
Journal of Magnetic Resonance Imaging 30: 1223–1228.
EU Directive 2004/40/EC. 2004. EU Directive 2004/40/EC of the European
Parliament and of the Council of 29 April 2004 on the minimum health and
safety requirements regarding the exposure of workers to the risks arising
from physical agents (electromagnetic fields). Official Journal of the European
Union 47: 1–9, L184.
EU Directive 2008/46/EC. 2008. EU Directive 2008/46/EC of the European Par
liament and of the Council of 23 April 2008 amending Directive 2004/40/
EC on the minimum health and safety requirements regarding the exposure
of workers to the risks arising from physical agents (electromagnetic fields).
Official Journal of the European Union 51: 81–89, L114.
EU Directive 2013/35/EU. 2013. Directive 2013/35/EU of the European Parlia
ment and of the Council of 29 June 2013 on the minimum health and safety
requirements regarding the exposure of workers to the risks arising from
physical agents (electromagnetic fields) (20th individual Directive within the
meaning of Article 16(1) of Directive 89/391/EEC) and repealing Directive
2004/40/EC. Official Journal of the European Union 56: 1–21, L179.
Febles Santana, V. M., Hernandez Armas, J. A., Martin Diaz, M. A., de Miguel
Bibao, S., de Aldecoa Fernandez, J. C., and Ramos Gonzalez, V. 2014.
Assessment of magnetic field in the surroundings of magnetic resonance
systems: Risks for professional staff. IFMBE Proceedings 41: 190–193.
Fuentes, M. A., Trikic, A., Wilson, S. J., and Crozier, S. 2008. Analysis and measure-
ments of magnetic field exposure for healthcare workers in selected MR envi-
ronment. IEEE Transactions on Biomedical Engineering 55(4): 1355–1364.
Gaffey, C. T., and Tenforde, T. S., 1981. Alterations in the rat electrocardiogram
induced by stationary magnetic fields. Bioelectromagnetics 2: 357–370.
Glover, P. M., Cavini, I., Qian, W., Bowtell, R., and Gowland, P. A. 2007. Magnetic-
field-induced vertigo: A theoretical and experimental investigation. Bio
electromagnetics 28: 349–361.
Hansson Mild, K., Hand, J., Hietanen, M., Gowland, P., Karpowicz, J., Keevil,
S., Van Rongen, E., Scarfi, M. R., and Wilen, J. 2013. Exposure classifi-
cation of MRI workers in epidemiological studies. Bioelectromagnetics
34(1): 81–84.
Health Protection Agency (HPA). 2008. Static magnetic fields. RCE-6 Report of
the independent Advisory Group on Non-ionising Radiation. Chilton, UK.
Heinrich, A., Szostek, A., Nees, F., Meyer, P., Semmler, W., and Flor, H. 2011.
Effects of static magnetic fields on cognition, vital signs, and sensory per-
ception: A meta-analysis. Journal of Magnetic Resonance Imaging 34(4):
758–763.
300 ◾ Biomagnetics
Heinrich, A., Szostek, A., Meyer, P., Nees, F., Rauschenberg, J., Groebner, J.,
Gilles, M., Paslakis, G., Deuschle, M., Semmler, W., and Flor, H. 2013.
Cognition and sensation in very high static magnetic fields: A randomized
case-crossover study with different field strengths. Radiology 266(1): 236–245.
Houpt, T. A., and Houpt, C. E. 2010. Circular swimming in mice after exposure to
a high magnetic field. Physiology & Behavior 100: 284–290.
Houpt, T. A., Carella, L., Gonzalez, G., Janowitz, I., Mueller, A., Mueller, K., Neth,
B., and Smith, J. C. 2011. Behavioral effects on rats of motion within a high
static magnetic field. Physiology & Behavior 102: 338–346.
HPA. 2012. Health effects from radiofrequency electromagnetic fields. RCE-20
Report of the independent Advisory Group on Non-ionising Radiation.
Chilton, UK.
IARC. 2002. Non-ionizing radiation. Part 1: Static and extremely low frequency
(ELF) electric and magnetic fields. IARC Monographs on the Evaluation
of Carcinogenic Risks to Humans, Volume 80. International Agency for
Research Cancer, Lyon.
IARC. 2013. Non-ionizing radiation. Part 2: Radiofrequency electromagnetic
fields. IARC Monographs on the Evaluation of Carcinogenic Risks to
Human, Volume 102. International Agency for Research Cancer, Lyon.
ICNIRP. 1998. Guidelines for limiting exposure to time-varying electric, magnetic,
and electromagnetic fields (up to 300 GHz). Health Physics 74(4): 494–522.
ICNIRP. 2003a. Exposure to static and low frequency electromagnetic fields, biolog-
ical effects and health consequences (0-100kHz). Matthes, R, McKinlay, A. F.,
Bernhardt, J. H., Vecchia, O., and Veyret, B., eds. ICNIRP, Oberschleissheim,
Germany.
ICNIRP. 2003b. Guidance on determining compliance of exposure to pulsed and
complex non-sinusoidal waveforms below 100 kHz with ICNIRP guidelines.
Health Physics 84: 383–387.
ICNIRP. 2004. Medical magnetic resonance (MR) procedures: Protection of
patients. Health Physics 87:197–216.
ICNIRP. 2009a. ICNIRP statement on the “Guidelines for limiting exposure to
time-varying electric, magnetic, and electromagnetic fields (up to 300 GHz).”
Health Physics 93: 257–258.
ICNIRP. 2009b. Guidelines on limits to exposure from static magnetic field. Health
Physics 96: 504–514.
ICNIRP. 2009c. ICNIRP statement on amendment to the ICNIRP “Statement
on medical magnetic resonance procedures: Protection of patients.” Health
Physics 97: 259–261.
ICNIRP. 2009d. Exposure to high frequency electromagnetic fields, biological
effects and health consequences (100 kHz–300 GHz). Review of the Sci
entific Evidence and Health Consequences. ICNIRP, Oberschleissheim,
Germany.
ICNIRP. 2010. Guidelines or limiting exposure to time-varying electric, mag-
netic, and electromagnetic fields (up to 100 kHz). Health Physics 99: 818–
836. (Erratum. 2011). 100: 112.
Safety Aspects of Magnetic and Electromagnetic Fields ◾ 301
Ueno, S., and Shigemitsu, T. 2007. Biological effects of static magnetic fields. In
Handbook of Biological Effects of Electromagnetic Fields: Bioengineering
and Biophysical Aspects of Electromagnetic Fields, 3rd ed. Barnes, F. S., and
Greenebaum, B., eds. CRC Press. Boca Raton, Fla.: 203
Ueno, S., and Okano, H. 2012. Static, low-frequency, and pulsed magnetic fields in
biological systems. In Electromagnetic Fields in Biological Systems. Lin, J. C.,
ed. CRC Press, Boca Raton, Fla.: 115–196.
Van Nierop, L. E., Slottje, P., van Zandvoort, M. J., de Vocht, F., and Kromhout,
H. 2012. Effects of magnetic stray fields from a 7 Tesla MRI scanner on neu-
rocognition: A double-blind randomized crossover study. Occupational &
Environmental Medicine 69(10): 759–766.
Van Nierop, L. E., Slottje, P., Kingma, H., and Kromhout, H. 2013. MRI-related
static magnetic stray fields and postural body sway: A double-blind random-
ized crossover study. Magnetic Resonance in Medicine 70: 232–240.
Wang, H., Trakic, A., Liu, F., and Crozier, S. 2008. Numerical field evaluation of
healthcare workers when bending towards high-field MRI magnets. Magnetic
Resonance in Medicine 59: 410–422.
WHO. 1987. Magnetic fields. Environmental Health Criteria 69. World Health
Organization, Geneva.
WHO. 2006a. Static fields. Environmental Health Criteria 232. World Health
Organization, Geneva.
WHO. 2006b. WHO research agenda for static fields. World Health Organization,
Geneva.
WHO. 2007a. Extremely low frequency fields. Environmental Health Criteria
238. World Health Organization, Geneva.
WHO. 2007b. WHO Research agenda for extremely low frequency fields. World
Health Organization, Geneva.
WHO. 2010. WHO research agenda for radiofrequency fields. World Health
Organization, Geneva.
WHO. 2014. The International EMF project. Progress Report June 2013–2104.
http://www.WHO-int/peh-emf/project/IAC_2014_Progress_Report
.pdf?ua=1.
Wilén, J., and de Vocht, F. 2011. Health complaints among nurses working near
MRI scanners- A descriptive pilot study. European Journal of Radiology
80(2): 510–513.
Yamaguchi-Sekino, S., Sekino, M., and Ueno, S. 2011. Biological effects of elec-
tromagnetic fields and recently updated safety guidelines for strong static
magnetic fields. Magnetic Resonance in Medical Science 10(1): 1–10.
Yamaguchi-Sekino, S., Nakai, T., Imai, S., Izawa, S., and Okuno, T. 2014. Occupa
tional exposure levels of static magnetic field during routine MRI examina-
tion in 3 T MR system. Bioelectromagnetics 35(1): 70–75.
Zahedi, Y., Zaun, G., Maderwald, S., Orzada, S., Putter, C., Scherag, A.,
Winterhager, E., Ladd, M. E., and Grummer, R. 2014. Impact of repetitive
exposure to strong static magnetic fields on pregnancy and embryonic devel-
opment of mice. Journal of Magnetic Resonance and Imaging 39(3): 691–699.
304 ◾ Biomagnetics
Zaun, G., Zahedi, Y., Maderwald, S., Orzada, S., Putter, C., Scherag, A.,
Winterhager, E., Ladd, M. E., and Grummer, R. 2014. Repetitive exposure
of mice to strong static magnetic fields in utero does not impair fertility in
adulthood but may affect placental weight of offspring. Journal of Magnetic
Resonance and Imaging 39(3): 683–690.
Zhao, G., Chen, S., Wang, L., Zhao, Y., Wang, J., Wang, X., Zhang, W., Wu, R., Lu,
L., Wu, Y., and Xu, A. 2011. Cellular ATP content was decreased by a homo-
geneous 8.5 T static magnetic field exposure: Role of reactive oxygen species.
Bioelectromagnetics 32(2): 94–101.
Chapter 10
New Horizons
in Biomagnetics
and Bioimaging
Shoogo Ueno and Masaki Sekino
CONTENTS
10.1 Advances in Biomagnetics 306
10.1.1 Deep Brain Stimulation for Therapeutic Application 306
10.1.2 Combination of TMS with DTI and EEG 307
10.1.3 Biomagnetic Approaches to Treatments of Cancer
and Other Diseases 309
10.2 Advances in Bioimaging 309
10.2.1 MEG and Multimodal Integration 309
10.2.2 Combination of SQUIDs and Ultra-Low Magnetic
Field MRI 309
10.2.3 Advances in MRI 310
10.2.4 Magnetic Particle Imaging and Microwave Imaging 311
10.2.5 Molecular Imaging 312
References 312
305
306 ◾ Biomagnetics
10.1 ADVANCES IN BIOMAGNETICS
10.1.1 Deep Brain Stimulation for Therapeutic Application
A method of deep brain stimulation (DBS) with implanted electrodes has
been used for the treatments of brain dysfunctions. The DBS has brought
good news1 for patients who are afflicted with, for example, unbearable
chronic pain, resistant movement, and affective disorders including major
depression. Although DBS is a powerful tool, it is an invasive method to
insert and implant electrodes inside the deeper parts of the brain.
If an alternative noninvasive method for deep brain stimulation is
introduced, new horizons will be opened in this area. The transcranial
magnetic stimulation (TMS) is a promising potential tool for the realiza-
tion of deep brain stimulation. In researching deep TMS (dTMS), several
attempts have been reported. There are several coil configurations poten-
tially suitable for dTMS; double cone, H-coil, and Halo coils.
The double-cone coil2–4 operates on the same principle as the figure-
eight coil configuration where a pair of opposing directed time-varying
magnetic fields around a target increase induced electric fields in the tar-
geted areas. The two circular coils have a fixed angle between them, and
their diameter is larger than that of the figure- eight coil. The double-cone
coil induces a greater electric field intensity to stimulate the deeper brain
regions compared with other coils. However, the surrounding large areas
in the brain are also stimulated.
Another coil configuration for dTMS is called the H-coil.5,6 The H-coils
have complex winding patterns based on numerical calculations of three-
dimensional brain phantom models to achieve effective stimulation of
deep brain structures and are used for the treatment of a variety of psychi-
atric and neurological disorders. Although the H-coils are used for clini-
cal applications, nontargeted areas of the brain are also stimulated.
A family of dTMS coil designs, called the Halo coil, was proposed to
increase the induced electric fields at depth in the brain.7 The Halo coil
system is composed of a small coil on the top of the head and a large cir-
cular coil around the head.
Computational studies were carried out in order to evaluate the focal-
ity of these three types of coil configurations.8,9 Three-dimensional dis-
tributions of the induced electric fields in realistic head model by dTMS
New Horizons in Biomagnetics and Bioimaging ◾ 307
coils were calculated by an impedance method, and the results were com-
pared with that of a standard figure-eight coil.10,11 Simulation results show
that double cone and H-coils have deep field penetration at the expense
of induced higher and wider spread electric fields in superficial cortical
regions. The combination of the Halo coil with a circular coil produce
deeply penetrating electric fields the same as double cone and H-coils, but
the stimulation in superficial brain tissues are much higher.10,11 Further
studies are needed for better dTMS with better focality.
The studies on TMS and reward circuits in the brain are important
and interesting for the understanding of the neuronal connectivity in the
brain as well as for the potential treatments of brain dysfunctions such as
depression and Parkinson’s disease. For example, Wassermann et al. have
studied reward-related activity in the human motor cortex,12 modulation
of corticospinal excitability by reward,13 reward processing abnormalities
in Parkinson’s disease,14 and so on. The study concludes that TMS of the
human primary motor cortex M1 may be useful as a quantitative measure
of reward-related activity.12
It seems that repetitive transcranial magnetic stimulation (rTMS), tran-
scranial direct current stimulation (tDCS), and other brain stimulation
techniques are promising for treatments of depression and other psychi-
atric disorders.15,16
In 2011, Fitzgerald reviewed the current state of development and appli-
cation of a wide range of brain stimulation approaches in the treatment of
psychiatric disorders.17 The brain stimulation methods that he reviewed
include vagus nerve stimulation, rTMS, tDCS, electroconvulsive therapy,
and magnetic seizure therapy. The review concludes that it appears likely
that the range of psychiatric treatments available for patients will grow
over the coming years to progressively include a number of novel brain
stimulation techniques.17
In 2013, Fitzgerald et al. examined a double-blind, randomized trial
of deep rTMS for autism spectrum disorder and came to the following
conclusion: Deep rTMS to bilateral dorsomedial prefrontal cortex yielded
a reduction in social relating impairment and socially related anxiety.18
Further studies are needed in the area of deep transcranial magnetic
stimulation.
10.2 ADVANCES IN BIOIMAGING
10.2.1 MEG and Multimodal Integration
With its high temporal resolution, magnetoencephalography (MEG) has
been a powerful tool to study and diagnose noninvasively the human
brain function in the millisecond time scale. A variety of studies have been
reported, for example, gamma phase locking,48 resting cortical dynam-
ics,49 intra- and interfrequency brain network structure.50 entrainment
of alpha oscillations,51 real-time MEG neurofeedback training,52 visual
object recognition related to feedback and feedforward inputs,53 and lan-
guage processing in atypical development.54
The combination of MEG and electroencephalographic (EEG) mea-
surements can reduce ambiguities in data from usage of only one of the
modalities.55 The combination of MEG with fMRI56,57 has accelerated
functional brain studies such as the study on human emotion cogni-
tion58 and gamma and alpha oscillations related to working memory per-
formance.59 Combining MEG with other noninvasive tools will further
developments to neuroscience and clinical medicine.
fields (ULFs) with readout magnetic fields as low as 132 μT and using
pulsed pre-polarization.60 In ULF-MRI, precession signals are detected
at ~μT magnetic fields using highly sensitive SQUID sensors. Compared
with conventional MRI with high magnetic fields of a few Teslas order, sig-
nal acquisition times or imaging times in ULF-MRI tend to be long. Many
studies on SQUID-detected ULF-MRI have been reported in the 2000s
and 2010s regarding distinct imaging with a shorter imaging time.62–69
10.2.3 Advances in MRI
In contrast with ULF-MRI, projects of ultra-high magnetic field MRI
(UHF-MRI) are also very important to get ultra-clear and distinct imag-
ing. Big projects for UHF-MRI have been promoted and are now being pro-
moted by different institutions, for example, by a group at the University
of Minnesota in the United States conducted by Ugurbil70 and by a group
at NeuroSpin in France conducted by Le Bihan.71 Larmor frequency, or
resonant frequency (RF), proportionally increases with increase in mag-
netic field. Many important issues need to be resolved in UHF-MRI such
as safety, superconducting coils, gradient coils, and RF coils related to the
inhomogeneity in RF field distributions and coil design, etc. The issues on
parallel transmission and RF coil design were studied72–74 to overcome the
problem in RF field inhomogeneity.
Diffusion tensor MRI (DTI) and in vivo fiber tractography invented by
Basser et al.19,20 have become powerful tools in the imaging of ultra-fine ana-
tomical structures of neuronal fibers in the human brain. Functional MRI
(fMRI) based on blood oxygenation level dependent (BOLD) effects invented
by Ogawa et al.56,57 has become a powerful tool to reveal the functional orga-
nization of the human brain. Data obtained by these two tools have enabled
us to study anatomical and functional connectivity in the brain. Polonara
et al. studied the topography and connectivity of the corpus callosum in
patients with partial callosal resection using both fMRI and DTI.75
Thomas et al. evaluated the inherent limit for anatomical accuracy
of brain connections derived from DTI tractography by comparing the
DTI data with the data obtained from tracer studies in the macaque.76
Anatomical accuracy is highly dependent upon parameters of the tractog-
raphy algorithm, with different optimal values for mapping different path-
ways. The results by Thomas et al. suggest there is an inherent limitation in
determining long-range anatomical projections based on boxel averaged
estimates of local fiber orientation obtained from DTI data that is unlikely
to be overcome by improvements in data acquisition and analysis alone.76
New Horizons in Biomagnetics and Bioimaging ◾ 311
10.2.5 Molecular Imaging
Studies of molecular imaging have been rapidly accelerated for medical appli-
cations in the 2000s through the 2010s. Brain research through advancing
innovative neurotechnologies (BRAIN) initiative was introduced by President
Obama in the United States, on April 2, 2013, and a National Institutes of
Health (NIH) Workshop on clinical translation of molecular imaging probes
and technology was held on August 2, 2013, in Bethesda, Maryland.88,89
Pettigrew, director of the national institute of biomedical imaging and bio
engineering (NIBIE) at the NIH, stated that “This high-priority BRAIN
initiative on next-generation human imaging technologies is a striking call
for teams of our best and brightest minds from multiple disciplines to think
big, aim high, and reach far beyond previous boundaries in human imaging
science.”89
Studies on in vivo imaging using rodents are promising and closest to
human imaging, and new strategies for fluorescent probe design in medi-
cal diagnostic imaging were reviewed by Kobayashi et al.90 Among them,
Urano et al., for example, developed a method of selective molecular
imaging of viable cancer cells with a pH-activatable fluorescence probe.91
Urano’s group also developed a unique method of rapid cancer detection
by topically spraying a γ-glutamyltranspeptidase (GGT)-activated fluores-
cent probe.92 This probe is practical for clinical application during surgical
or endoscopic procedures because of its rapid and strong activation upon
contact with GGT on the surface of cancer cells.
The future of science and medicine will be rapidly progressing with the
integration of advanced technologies, including biomagnetics, bioimaging,
drug delivery system, and nanobioscience.93
Biomagnetics and bioimaging are thus leading medicine and biology
into new horizons through their novel applications of magnetism and
imaging technology. With the increasing integration of medicine and
engineering, biomagnetics and bioimaging are merging into a new science
that encompasses a wide range of fields including their diverse cultures to
create a new and original discipline.
REFERENCES
1. Kringelbach, M. L., Jenkinson, N., Owen, S. L. F., and Aziz, T. Z. 2007.
Translational principles of deep brain stimulation. Nature Reviews
Neuroscience 8(8): 623–635.
2. Ugawa, Y., Uesaka, Y., Terao, Y., Hanajima, R., and Kanazawa, I. 1995. Magnetic
stimulation over the cerebellum in humans. Annals Neurology 37: 703–713.
New Horizons in Biomagnetics and Bioimaging ◾ 313
3. Roth, Y., Zangen, A., and Hallett, M. 2002. A coil design for transcra-
nial magnetic stimulation of deep brain regions. Journal of Clinical
Neurophysiology 19: 361–370.
4. Lontis, E. R., Voigt, M., and Struijk, J. J. 2006. Focality assessment in
transcranial magnetic stimulation with double and cone coils. Journal of
Clinical Neurophysiology 23: 462–471.
5. Zangen, A., Roth, Y., Voller, B., and Hallett, M. 2005. Transcranial mag-
netic stimulation of deep brain regions; evidence for efficacy of the H-coil.
Clinical Neurophysiology 116:775–779.
6. Roth, Y., Amir, A., Levkovitz, Y., and Zangen, A. 2007. Three-dimensional
distributions of the electric fields induced in the brain by transcranial
magnetic stimulation using figure-8 and deep H-coils. Journal of Clinical
Neurophysiology 24: 31–38.
7. Crowther, L. J., Marketos, P., Williams, P. I., Melikhov, Y., and Jiles, D. C.
2011. Transcranial magnetic stimulation: Improved coil design for deep
brain investigation. Journal of Applied Physics 109: 07B314.
8. Lu, M., and Ueno, S. 2013. Calculating the electric fields in the human
brain by deep transcranial magnetic stimulation. Proceedings of the IEEE
Engineering in Medicine and Biology Society, 376–379.
9. Lu, M., and Ueno, S. 2013. Numerical simulation of deep brain transcra-
nial magnetic stimulation. In Transcranial Magnetic Stimulation, Lucia
Alba-Ferrara, L., ed. Nova Science Publishers, Hauppauge, New York:
41–60.
10. Ueno, S., Tashiro, T., and Harada, S. 1988. Localized stimulation of neu-
ral tissues in the brain by means of a paired configuration of time-varying
magnetic fields. Journal of Applied Physics 64: 5862–5864.
11. Ueno, S., Matsuda, T., and Fujiki, M. 1990. Functional mapping of the
human motor cortex obtained by focal and vectorial magnetic stimulation
of the brain. IEEE Transactions on Magnetics 26: 1539–1544.
12. Kapogiannis, D., Campion, P., Grafman, J., and Wassermann, E. M. 2008.
Reward-related activity in the human motor cortex. European Journal of
Neuroscience 27: 1836–1842.
13. Mooshagian, E., Keisler, A., Zimmermann, T., Schweickert, J. M., and
Wassermann, E. M. 2015. Modulation of corticospinal excitability by
reward depends on task framing. Neuropsychologia 68: 31–37.
14. Kapogiannis, D., Mooshagian, E., Campion, P., Grafman, J., Zimmermann,
T. J., Ladt, K. C., and Wassermann, E. M. 2011. Reward processing abnor-
malities in Parkinson’s disease. Movement Disorders 26: 1451–1457.
15. Fitzgerald, P. B., and Daskalakis, Z. J. 2012. A practical guide to the use of
repetitive transcranial magnetic stimulation in the treatment of depression.
Brain Stimulation 5: 287–296.
16. Fitzgerald, P. B., Hoy, K. E., Singh, A., Gunewardene, R., Slack, C., Ibrahim,
S., Hall, P. J., and Daskalakis, Z. J. 2013. Equivalent beneficial effects of
unilateral and bilateral prefrontal cortex transcranial magnetic stimula-
tion in a large randomized trial in treatment-resistant major depression.
International Journal of Neuropsychopharmacology 16: 1975–1984.
314 ◾ Biomagnetics
17. Fitzgerald, P. B. 2011. The emerging use of brain stimulation treatments for
psychiatric disorders. Australian and New Zealand Journal of Psychiatry 45:
923–938.
18. Enticott, P. G., Fitzgibbon, B. M., Kennedy, H. A., Arnold, S. L., Elliot, D.,
Peachey, A., Zangen, A., and Fitzgerald, P. B. 2013. A double-blind, ran-
domized trial of deep repetitive transcranial magnetic stimulation (fTMS)
for autism spectrum disorder. Brain Stimulation 6: 1–6.
19. Basser, P. J., Mattiello, J., and Le Bihan, D. 1994. Estimation of the effec-
tive self-diffusion tensor from the NMR spin echo. Journal of Magnetic
Resonance B 103: 247–254.
20. Basser, P. J., Pajevic, S., Pierpaoli, C., Duda, J., and Aldroubi, A. 2000. In vivo
fiber tractography using DT-MRI data. Magnetic Resonance in Medicine 44:
625–632.
21. De Geeter, N., Crevecoeur, G., Dupre, L., Van Hecke, W., and Leemans,
A. 2012. A DTI-based model for TMS using the independent impedance
method with frequency-dependent tissue parameters. Physics in Medicine
and Biology 57: 2169–2188.
22. De Geeter, N., Crevecoeur, G., Leemans, A., and Dupre, L. 2015. Effective
electric fields along realistic DTI-based neural trajectories for modelling
the stimulation mechanisms of TMS. Physics in Medicine and Biology 60:
453–471.
23. Roth, B. J., and Basser, P. J. 1990. A model of the stimulation of a nerve fiber
by electromagnetic induction. IEEE Transactions on Biomedical Engineering
37: 588–597.
24. Ueno, S., Matsuda, T., and Hiwaki, O. 1991. Estimation of structures of
neural fibers in the human brain by vectorial magnetic stimulation. IEEE
Transactions on Magnetics 27: 5387–5389.
25. Basser, P., Wijesinghe, R. S., and Roth, B. J. 1992. The activating function for
magnetic stimulation derived from a three-dimensional volume conductor
model. IEEE Transactions on Biomedical Engineering 39: 1207–1211.
26. Nagarajan, S. S., Dutand, M., and Warman, E. N. 1993. Effects of induced
electric fields on finite neuronal structures: A simulation study. IEEE
Transactions on Biomedical Engineering 40: 1175–1188.
27. Hyodo, A., and Ueno, S. 1996. Nerve excitation model for localized magnetic
stimulation of finite neuronal structures. IEEE Transactions on Magnetics
32: 5112–5114.
28. Liu, R., and Ueno, S. 2000. Calculating the activating function of nerve
excitation in inhomogeneous volume conductor during magnetic stimu-
lation using finite element method. IEEE Transactions on Magnetics 36:
1796–1799.
29. Lu, M., Ueno, S., Thorlin, T., and Persson, M. 2008. Calculating the acti-
vating function in the human brain by transcranial magnetic stimulation.
IEEE Transactions on Magnetics 44: 1438–1441.
30. Hiwaki, O., and Inoue, T. 2009. A method for estimation of stimulated
brain sites based on columnar structure of cerebral cortex in transcranial
magnetic stimulation. Journal of Applied Physics 105: 07B303.
New Horizons in Biomagnetics and Bioimaging ◾ 315
31. Ilmoniemi, R. J., Virtanen, J., Karhu, J., Aronen, H. J., Naatanen, R., and
Katila, T. 1997. Neuronal responses to magnetic stimulation reveal cortical
reactivity and connectivity. Neuroreport 8: 3537–3540.
32. Ilmoniemi, R. J., and Karhu, J. 2008. TMS and electroencephalography:
Methods and current advances. In The Oxford Handbook of Transcranial
Stimulation, Wasserman E. M., Epstein, C. M., Ziemann, U., Paus, T., and
Lisanby, S. H., eds. Oxford University Press, New York: 593–608.
33. Raij, T., Karhu, J., Kicic, D., Lioumis, P., Julkunen, P., Lin, F. H., Ahveninen,
J., Ilmoniemi, R. J., Makela, J. P., Hamalainen, M., Rosen, B. R., and
Belliveau, J. W. 2008. Parallel input makes the brain run faster. NeuroImage
40: 1792–1797.
34. Ilmoniemi, R. J., and Dubravko, K. 2010. Methodology for combined TMS
and EEG. Brain Topography 22: 233–248.
35. Maki, H., and Ilmoniemi, R. J. 2010. EEG oscillations and magnetically
evoked motor potentials reflect motor system excitability in overlapping
neuronal populations. Clinical Neurophysiology 121: 492–501.
36. Iramina, K., Maeno, T., Kowatari, Y., and Ueno, S. 2002. Effects of transcra-
nial magnetic stimulation on EEG activity. IEEE Transactions on Magnetics
38: 3347–3349.
37. Iramina, K., Maeno, T., Nonaka, Y., and Ueno, S. 2003. Measurement of
evoked electroencephalography induced by transcranial magnetic stimula-
tion. Journal of Applied Physics 93: 6718–6720.
38. Iwahashi, M., Koyama, Y., Hyodo, A., Hayami, T., Ueno, S., and Iramina,
K. 2009. Measurements of evoked electroencephalograph by transcranial
magnetic stimulation applied to motor cortex and posterior parietal cortex.
Journal of Applied Physics 105: 07B321.
39. Torii, T., Sato, A., Nakahara, Y., Iwahashi, M., Itoh, Y., and Iramina, K. 2012.
Frequency-dependent effects of repetitive transcranial magnetic stimula-
tion on the human brain. Neuroreport 19: 1065–1070.
40. Marshall, T. R., O’Shea, J., Jensen, O., and Bergmann, T. O. 2015. Frontal
eye fields control attentional modulation of alpha and gamma oscillations
in contralateral occipitoparietal cortex. Journal of Neuroscience 35: 1–10.
41. Fox, M. D., Halko, M. A., Eldaief, M. C., and Pascual-Leone, A. 2012.
Measuring and manipulating brain connectivity with resting state func-
tional connectivity magnetic resonance imaging (fcMRI) and transcranial
magnetic stimulation (TMS). Neuroimaging 62: 2232–2243.
42. Ogiue-Ikeda, M., Sato, Y., and Ueno, S. 2003. A new method to destruct
targeted cells using magnetizable beads and pulsed magnetic force. IEEE
Transactions on Nanobioscience 2: 262–265.
43. Yamaguchi, S., Ogiue-Ikeda, M., Sekino, M., and Ueno, S. 2006. The effect
of repetitive magnetic stimulation on tumor and immune functions in
mice. Bioelectromagnetics 27: 64–72.
44. Kotani, H., Kawaguchi, H., Shimoaka, T., Iwasaka, M., Ueno, S., Ozawa, H.,
Nakamura, K., and Hoshi, K. 2002. Strong static magnetic field stimulates
bone formation to a definite orientation in vivo and in vitro. Journal of Bone
and Mineral Research 17: 1814–1821.
316 ◾ Biomagnetics
45. Eguchi, Y., Ohtori, S., and Ueno, S. 2015. The effectiveness of mag-
netically aligned collagen for neural regeneration in vitro and in vivo.
Bioelectromagnetics 36: 233–243.
46. Cespedes, O., and Ueno, S. 2009. Effects of radio frequency magnetic fields
on iron release from cage proteins. Bioelectromagnetics 30: 336–342.
47. Cespedes, O., Inmoto, O., Kai, S., Nibu, Y., Yamaguchi, T., Sakamoto, N.,
Akune, T., Inoue, M., Kiss, T., and Ueno, S. 2010. Radio frequency mag-
netic field effects on molecular dynamics and iron uptake in cage proteins.
Bioelectromagnetics 31: 311–317.
48. Han, J., Mody, M., and Ahlfors, S. P. 2012. Gamma phase locking modu-
lated by phonological contrast during auditory comprehension in reading
disability. Cognitive neuroscience and neuropsychology 24: 851–856.
49. Shriki, O., Alstott, J., Carver, F., Holroyd, T., Henson, R. N. A., Smith, M. L.,
Coppola, R., Bullmore, E., and Plenz, D. Neuronal avalanches in the resting
MEG of the human brain. Journal of Neuroscience 33: 7079–7090.
50. Siebenhuner, F., Weiss, S. A., Coppola, R., Weinberger, D. R., and Bassett,
D. S. 2013. Intra- and inter-frequency brain network structure in health and
schizophrenia. PLoS One 8: 1–13.
51. Spaak, E., de Lange, F. P., and Jensen, O. 2014. Local entrainment of alpha
oscillations by visual stimuli causes cyclic modulation of perception.
Journal of Neuroscience 34: 3536–3544.
52. Okazaki, Y., Horschig, J. M., Luther, L., Oostenveld, R., Murakami, I., and
Jensen, O. 2015. NeuroImage 107: 323–332.
53. Ahlfors, S. P., Jones, S. R., Ahveninen, J., Hamalainen, M. S., Belliveau,
J. W., and Bar, M. 2015. Direction of magnetoenchephalography source
associated with feedback and feedforward contributions in a visual object
recognition task. Neuroscience Letters 585: 149–154.
54. Mody, M. 2014. Language processing in atypical development: looking
below the surface with MEG. In Magnetoencephalography, Supek, S., and
Aine, C. J., eds. Springer-Verlag, Berlin: 579–597.
55. Ahlfors, S. P. 2014. MEG and multimodal integration. In Magnetoence
phalography, Supek, S., and Aine, C. J., eds. Springer-Verlag, Berlin: 183–198.
56. Ogawa, S., Lee, T. M., Nayak, A. S., and Glynn, P. 1990. Oxygenation-
sensitive contrast in magnetic resonance image of rodent brain at high
magnetic field. Magnetic Resonance in Medicine 14: 68–78.
57. Ogawa, S., Lee, T. M., Kay, A. R., and Tank, D. W. 1990. Brain magnetic reso-
nance imaging with contract dependent on blood oxygenation. Proceedings of
the National Academy of Sciences of the United States of America 87: 9868–9872.
58. Jabbi, M., Kohn, P. D., Nash, T., Ianni, A., Coutlee, C., Holroyd, R., Carver,
F. W., Chen, Q., Cropp, B., Kippenhan, J. S., Robinson, S. E., Coppola, R.,
and Berman, K. F. 2014. Convergent BOLD and beta-band activity in supe-
rior temporal sulcus and frontolimbic circuitry underpins human emotion
cognition. Cerebral Cortex, in press.
59. Lozano-Soldevilla, D., ter Huurne, N. Cools, R., and Jensen, O. 2014.
GABAergic modulation of visual gamma and alpha oscillations and its con-
sequences for working memory performance. Current Biology 24: 2876–2887.
New Horizons in Biomagnetics and Bioimaging ◾ 317
60. Clarke, J., Hatridge, M., and Mossle. M. 2007. SQUID detected magnetic
resonance in microtesla fields. Annual Reviews of Biomedical Engineering 9:
389–413.
61. Mossle, M., Han, S. I., Myers, W. R., Lee, S. K., Kelso, N., Pines, A., and
Clarke, J. 2006. SQUID-detected microtesla MRI in the presence of metal.
Journal of Magnetic Resonance 179: 146–151.
62. Matlachov, A. N., Volegov, P. L., Espy, M. A., George, J. S., and Kraus Jr.,
R. H. 2004. SQUID detected NMR in microtesla magnetic fields. Journal of
Magnetic Resonance 170: 1–7.
63. Espy, M., Matlashov, A., and Volegov, P. 2013. SQUID-detected ultra-low
field MRI. Journal of Magnetic Resonance 229: 127–141.
64. Nieminen, J. O., Burghoff, M., Trahms, L., and Ilmoniemi, R. J. 2010.
Polarization encoding as a novel approach to MRI. Journal of Magnetic
Resonance 202: 211–216.
65. Vesanen, P. T., Nieminen, J. O., Zevenhoven, K. C., Dabek, J., Parkkonen,
L. T., Zhdanov, A. V., Luomahaara, J., Hassel, J., Penttila, J., Simonla, J.,
Ahonen, A. I., Makela, J. P., and Ilmoniemi, R. J. 2013. Hybrid ultra-low-
field MRI and magnetoenchephalography system based on a commercial
whole-head neuromagnetometer. Magnetic Resonance in Medicine 69:
1795–1804.
66. Nieminen, J. O., Zevenhoven, K. C. J., Vesanen, P. T., Hsu, Y. C., and
Ilmoniemi, R. J. 2014. Current-density imaging using ultra-low-field MRI
with adiabatic pulses. Magnetic Resonance Imaging 32: 54–59.
67. Vesanen, P. T., Nieminen, J. O., Zevenhoven, K. C. J., Hsu, Y. C., and
Ilmoniemi, R. J. 2014. Current-density imaging using ultra-low-field MRI
with zero-field encoding. Magnetic Resonance Imaging 32: 1–5.
68. Espy, M. A., Magnelind, P. E., Matlashov, A. N., Newman, S. G., Sandin,
H. J., Schultz, L. J., Sedillo, R., Urbaitis, A. V., and Volegov, P. L. 2015.
Progress toward a deployable SQUID-based ultra-low field MRI system
for anatomical imaging. IEEE Transactions on Applied Superconductivity
25: 1601705.
69. Woo, T., Nagase, M., Ohashi, H., and Sekino, M. 2015. Development of a
SQUID system for ultralow-field MRI measurement. International Journal
of Applied Electromagnetics and Mechanics, in press.
70. Ugurbil, K. 2014. Magnetic resonance imaging at ultrahigh fields. IEEE
Transactions on Biomedical Engineering 61: 1364–1379.
71. Le Bihan, D. 2014. Diffusion MRI: What water tells us about the brain.
EMBO Molecular Medicine 6: 569–573.
72. Pruessmann, K. P. 2006. Encoding and reconstruction in parallel NMR.
NMR in Biomedicine 19: 288–299.
73. Katscher, U., and Bornert, P. 2006. Parallel RF transmission in MRI. NMR
in Biomedicine 19: 393–400.
74. Sekino, M., Boulant, N., Luong, M., Amadon, A., Ohsaki, H., and Le Bihan,
D. 2009. Effects of relaxation during RF pulses on the homogeneity of signal
intensity in parallel transmission. Proceedings of the ISMRM 17th Scientific
Meeting and Exhibition 2568.
318 ◾ Biomagnetics
75. Polonara, G., Mascioll, G., Foschi, N., Salvolini, U., Pierpali, C., Manzoni,
T., Fabri, M., and Barbaresi, P. 2014. Further evidence for the topography
and connectivity of the corpus callosum: An fMRI study of patients with
partial callosal resection. Journal of Neuroimaging, in press.
76. Thomas, C., Ye, F. Q., Irfanoglu, M. O., Madi, P., Saleen, K. S., Leopold,
D. A., and Pierpaoli, C. 2015. Anatomical accuracy of brain connec-
tions derived from diffusion MRI tractography is inherently limited.
Proceedings of the National Academy of Sciences of the United States of
America 111: 16,574–16,579.
77. Sekino, M., Sano, M., Ogiue-Ikeda, M., and Ueno, S. 2006. Estimation of
membrane permeability and intracellular diffusion coefficient. Magnetic
Resonance in Medical Sciences 5: 1–6.
78. Imae, T., Shinohara, H., Sekino, M., Ueno, S., Ohsaki, H., Mima, K., and
Ohtomo, K. 2008. Estimation of cell membrane permeability of the rat
brain using diffusion magnetic resonance imaging. Journal of Applied
Physics 103: 07A311.
79. Imae, T., Shinohara, T., Sekino, M., Ueno, S., Ohsaki, H., Mima, K., and
Ohtomo, K. 2009. Estimation of cell membrane permeability and intracel-
lular diffusion coefficient of the human gray matter. Magnetic Resonance in
Medical Science 8: 1–7.
80. Krishnan, K. M. 2010. Biomedical nanomagnetics: A spin through new
possibilities in mapping, diagnostics and therapy. IEEE Transactions on
Magnetics Advances in Magnetics 46: 2523–2558.
81. Shiozawa, M., Kobayashi, S., Sato, Y., Maeshima, H., Hozumi, Y., Lefor,
A. T., Kurihara, K., Sata, N., and Yasuda, Y. 2012. Magnetic resonance lym-
phography of sentinel lymph nodes in patients with breast cancer using
superparamagnetic iron oxide: A feasibility study. Breast Cancer 21(4):
394–401.
82. Minamiya, Y., Ito, M., Katayose, Y., Saito, H., Imai, K., Sato, Y., and Ogawa,
J. 2006. Intraoperative sentinel lymph node mapping using a new steriliz-
able magnetometer in patients with nonsmall cell lung cancer. Annals of
Thoracic Surgery 81: 327–330.
83. Ookubo, T., Inoue, Y., Kim, D., Ohsaki, H., Masahiko, Y., Kusakabe, M.,
and Sekino, M. 2013. Characteristics of magnetic probes for identifying
sentinel lymph nodes. Proceedings of the IEEE Engineering in Medicine and
Biology Society 2013: 5485–5488.
84. Gleich, B., and Weizenecker, J. 2005. Tomographic imaging using the non-
linear response of magnetic particles. Nature 435: 1214–1217.
85. Satttel, T. F., Knopp, T., Biederer, S., Gleich, B., Weizenecker, J., Borgert, J.,
and Buzug, T. M. 2009. Single-sided device for magnetic particle imaging.
Journal of Physics D: Applied Physics 42: 022001.
86. Morishige, T., Mihaya, T., Bai, S., Miyazaki, T., Yoshida, T., Matsuo, M., and
Empuku, K. 2014. Highly sensitive magnetic nanoparticle imaging using
cool-Cu/HTS-superconductor pickup coils. IEEE Transactions on Applied
Superconductivity 24: 1800105.
New Horizons in Biomagnetics and Bioimaging ◾ 319
87. Persson, M., Fhager, A., Dobsicek, H., Yu, Y., McKelvey, T., Pegenius, G.,
Karlsson, J. E., and Elam, M. 2014. Microwave-based stroke diagnosis mak-
ing global pre-hospital thrombolytic treatment possible. IEEE Transactions
on Biomedical Engineering 61: 2806–2817.
88. Liu, C. H., Sastre, A., Conroy, R., Seto, B., and Pettigrew, R. I. 2014. NIH work-
shop on clinical translation of molecular imaging probes and technology—
Meeting report. Molecular Imaging and Biology 16: 595–604.
89. Pettigrew, R. I. 2014. BRAIN initiative to transform human imaging.
Neuroscience 6: 1–2.
90. Kobayashi, H., Ogawa, M., Alford, R., Choyke, P. L., and Urano, Y. 2010.
New strategies for fluorescent probe design in medical diagnostic imaging.
Chemical Reviews 110: 2620–2640.
91. Urano, Y., Asanuma, D., Hama, Y., Koyama, Y., Barrett, T., Kamiya, M.,
Nagano, T., Watanabe, T., Hasegawa, A., Choyke, P. L., and Kobayashi, H.
2009. Selective molecular imaging of viable cancer cells with pH-activatable
fluorescence probes. Nature Medicine 15: 104–109.
92. Urano, Y., Sakabe, M., Kosaka, N., Ogawa, M., Mitsunaga, M., Asanuma, D.,
Kamiya, M., Young, M. R., Nagano, T., Choyke, P. L., and Kobayashi, H. 2011.
Rapid cancer detection by topically spraying a γ-glutamyltranspeptidase-
activated fluorescent probe. Science Translational Medicine 3: 1–10.
93. Ueno, S., Ando, J., Sugawara, T., Jimbo, Y., Itaka, K, Kataoka, K., and Ushida,
T. 2006. The state of the art of nanobioscience in Japan. IEEE Transactions
on Nanobioscience 5: 1–12.
Index
321
322 ◾ Index
SAR, See Specific absorption rate (SAR) signal inhomogeneities and, 157
Scalar potential finite difference (SPFD) from square of electric field strength,
method, 49–50 265
Schwann cells Spectroscopic techniques, in
magnetic orientation in, 187, 198, 198f, determination of iron contents,
203, 205 224–225
in medical and tissue engineering, 203 SPFD, See Scalar potential finite difference
nerve regeneration and (SPFD) method
magnetic field on, 203–205, 204f Spin echo (SE) signal, 127, 127f, 128, 130,
in PNS, 202–203 133, 133f
S-COL (collagen gel mixed with Schwann Spinning nuclei, 125–126
cells), 205 Spin-warp imaging technique, 127–137
SE, See Spin echo (SE) signal image contrast, 135–137, 136f, 137f
Sekino, M., 11 k space, 133–135, 134f, 135f
Sensitivity, to weak electric currents, 170–174 nuclei distributed in first direction,
experimental evaluation, 174 128–129, 129f
neuronal current dipole, 170–171 nuclei distributed in second direction,
theoretical evaluation, 171–173, 173f 129–131, 130f, 131f
Sentinel lymph nodes, imaging, 311 nuclei distributed in third direction,
Shielding effects, of eddy currents, 159 131–133, 132f, 133f
Shimming, 126 SQUID, See Superconducting quantum
Signals, MEG interference devices (SQUID)
beamformer, 107–108, 108f Standing for integrated parallel
distributed source imaging, 105–108 acquisition techniques square
equivalent current dipole, 101–105, (iPAT2), 137
106f Static magnetic fields, 13, 14
generation, 96–100 biological effects related to MRI,
inverse problem and, 100, 101f 266–272
minimum-norm estimates, 106–107 human experimental and
neural source reconstruction from, epidemiological studies, 270–272
100–108 in vitro studies, 269–270
“Silent” source, magnetically, 100 in vivo studies, 266–269
Single ECD model, 101 biological interaction in, 262–263
Singlet-triplet intersystem crossing, on biological systems, 188–194
196–197, 196f magnetic force, 189–194, 190f
Singular value decomposition (SVD), 103 magnetic torque in homogeneous
Skin effect, 154, 155f magnetic field, 191
Slice-selective excitation of nuclei, overview, 188f
128–129, 129f bone growth acceleration by, 198–202
Smooth muscle cells, magnetic orientation background, 198–199
in, 187, 198, 198f magnetic control, 200–202, 200f, 201f
Soma (cell body), 94, 95f exposure guidelines related to MRI,
Somatosensory-evoked potentials, 174 277–278, 277t
Spatial resolution, brain mapping, 64–66, in MRI, 260
65f, 66f in nerve regeneration, 202–205
Specific absorption rate (SAR), 13 background, 202–203, 202f
calculation, 236 Schwann cells, magnetic field on,
quantity, 265 203–205, 204f
342 ◾ Index
Torque, magnetic V
on biological systems in homogeneous
Varied frequencies, RF distribution in
magnetic field, 191
human head at, 157–158, 158f
defined, 188, 189
Virtual lesion, 69
Transcranial direct current stimulation
Visual evoked responses, 109–110, 109f,
(tDCS), 307
110f
Transcranial magnetic stimulation (TMS),
Visual perception, 3–4, 25
6, 25, 26, 26f, 46–47
Volume conductor, dipolar current in,
applications of, 195
97–98, 97f
for DBS, 306–307
DTI and EEG, combination, 307–308
dTMS, 306–307 W
EEG evoked by TMS, 66, 67f, 68
Wallerian degeneration, defined, 202, 203
FDA clearance for, 63
Water
rTMS, 307
relative permittivity of, 149
stimulation site, 46–47, 47f
Weak electric currents, sensitivity to,
vs. ECT, 51–53, 52f
170–174
Transmitters, RF, 165–166, 166f
experimental evaluation, 174
TSE image, 137
neuronal current dipole, 170–171
Tumor growth inhibition
theoretical evaluation, 171–173, 173f
by pulsed magnetic stimulation,
Weizmann Institute of Science, 4
209–212
WHO (World Health Organization), 188,
background, 209–210, 210t
266, 268, 270, 271
magnetic control, 211–212, 211f, 212f
William, Gilbert, 2–3, 2f
Tumor necrosis factor (TNF-α)
Worker exposure assessment, 289–291
production, 211–212
World Health Organization (WHO), 188,
T1-weighted images, 136–137, 136f
266, 268, 270, 271
T2-weighted images, 136–137, 136f, 137f
U Y