vvEPA
              United States       Health Effects Research
              Environmental Protection   Laboratory
              Agency          Research Triangle Park NC 27711
                             September 1984
                             Final Report
              Research and Development
Biological Effects of
Radiofrequency
Radiation
                                            600883026F

-------

-------
                                      EPA-600/8-83-026F
                                        September 1984
     Biological  Effects of
Radiofrequency Radiation
                 Edited by

          Joe A. Elder and Daniel F. Cahill
         Health Effects Research Laboratory
      Research Triangle Park, North Carolina 27711
       Health Effects Research Laboratory
       Office of Research and Development
      U.S. Environmental Protection Agency
   Research Triangle Park, North Carolina 27711

-------
                            Disclaimer
This document has been subjected to the Agency's required peer and policy
review and approved for publication. Mention of trade names or commercial
products does not signifiy official endorsement.

-------
                               Foreword


The many benefits of our modern, developing industrial society are accompanied
by certain hazards. Careful assessment of the risk of existing and new man-
made environmental hazards is necessary to establish sound regulatory policy.
Environmental regulations enhance the quality of our environment in order to
promote the public health and welfare and the productive capacity  of our
nation's population.

The Health Effects Research Laboratory conducts a coordinated environmental
health research program in toxicology and clinical studies. These studies
address problems in air pollution,  radiofrequency radiation, environmental
carcinogenesis, and the toxicology  of pesticides, as well as other chemical
pollutants. The Laboratory participates in the development and revision of air
quality criteria documents on pollutants for which national ambient air quality
standards exist or are proposed,  provides  the  data for registration of new
pesticides or proposed suspension of those already in use, conducts research on
hazardous and toxic materials, and  is primarily responsible for providing the
health basis  for radiofrequency radiation guidelines. Direct support to the
regulatory function of the Agency is provided in the form of expert testimony and
preparation of affidavits as well as expert advice to the Administrator.

The intent of this document is to provide a comprehensive review of the scientific
literature on the biological effects of radiofrequency radiation. The purpose of
this effort is  to evaluate critically  the current state of knowledge for its
pertinence and applicability  in developing radiofrequency-radiation exposure
guidelines for the general public.


                                        F. G. Hueter, Ph.D.
                                        Director
                                        Health Effects Research Laboratory

-------
                              Abstract
This document presents a critical review of the available literature on the
biological effects of radiofrequency (RF)  radiation. The objective was to
summarize and  evaluate the existing database for use  in developing RF-
radiation exposure guidance for the general public.

The frequency range of concern in this document is 0.5 MHz to 100 GHz, which
includes nearly all the significant sources of population exposure to RF radiation,
except 60-Hz electrical power systems. Research reports that are judged to be
credible according to a set of objective criteria are examined for the relation
between the RF energy absorbed and the presence or absence of biological
effects. The reported consequences of the interaction between RF radiation and
biological systems are examined from two perspectives: whole-body-averaged
specific  absorption rate (SAR)  and RF-energy-induced core-temperature
increases.

The existing database provides sufficient evidence about the relation between
RF-radiation exposure and biological effects to permit development of exposure
limits to protect the health of the general public. It has been concluded from this
review that biological effects occur at an SAR of about 1  W/kg; some of them
may be significant under certain environmental conditions.
                                  IV

-------
                              Contents

                                                                  Page

Foreword	iii

Abstract	iv

Figures	ix

Tables	xiii

Acknowledgment	  xvi

 1.  Introduction	 1-1
     (Daniel F. Cahill and Joe A. Elder)

 2.  Approach	2-1
     (Joe A. Elder and Daniel F. Cahill)

     2.1   General Approach	2-1
     2.2   Specific Approach	2-1
     2.3   Major Sections	2-1

 3.  Physical Principles of Electromagnetic Field Interactions	3-1

     3.1   Elements of Electromagnetic Field Theory	3-1
          (William T. Joines)
          3.1.1  Electromagnetic spectrum	3-1
          3.1.2  Wave propagation  	3-1
          3.1.3  Wave modulation	3-4

     3.2   RF-Field Interactions with Biological Systems	3-6
          (Claude M. Weil and James R. Rabinowitz)
          3.2.1  Scattering  and absorption of electromagnetic
                waves	3-6
          3.2.2  RF dosimetry definitions	3-11
          3.2.3  Analytical and numerical RF-electromagnetic
                interaction models	3-13
          3.2.4  Mechanisms of RF interaction with biological
                systems	3-19

     3.3   Experimental Methods	3-25
          (Claude M. Weil and Joseph S. Ali)
          3.3.1  Exposure methods used  in biological
                experimentation	3-25
          3.3.2  Animal holders	3-40
          3.3.3  Densitometric instrumentation	3-42

     3.4   Dosimetric Methods	3-47
          (James B. Kinn)
          3.4.1  Whole-body dosimetry 	3-47
          3.4.2  Regional dosimetry 	3-48
          3.4.3  Unresolved issues	3-49

-------
                       Contents (Continued)

                                                                 Page

4.  Effect of RF-Radiation Exposure on Body Temperature	4-1

    4.1   Thermal Physiology	4-1
         (Christopher J. Gordon)
         4.1.1  Temperature regulation	4-1
         4.1.2  Effect of RF radiation on thermophysiological
               effectors	4-13
         4.1.3  Body temperature regulation during
               RF-radiation exposure	4-18
         4.1.4  Effect of body size on thermoregulatory
               sensitivity to RF radiation	4-22
         4.1.5  Unresolved issues	4-27

    4.2   Numerical Modeling of Thermoregulatory Systems in Man ... 4-29
         (Ronald J. Spiegel)
         4.2.1  Heat-transfer models	4-29
         4.2.2  RF-radiation/heat-transfer models	4-30
         4.2.3  Numerical  results	4-32
         4.2.4  Unresolved issues	4-33

5.  Biological Effects of RF Radiation 	5-1

    5.1   Cellular and Subcellular Effects	5-1
         (John W. Allis)
         5.1.1  Effects on molecular systems	5-3
         5.1.2  Effects on subcellular organelles	5-4
         5.1.3  Effects on single cells	5-6
         5.1.4  Unresolved issues	5-9

    5.2   Hematologic and Immunologic Effects	5-13
         (Ralph J. Smialowicz)
         5.2.1  Hematology	5-13
         5.2.2  Immunology	5-18
         5.2.3  Unresolved issues  	5-27

    5.3   Reproductive Effects	5-29
         (Ezra Berman)
         5.3.1  Teratology	5-29
         5.3.2  Reproductive efficiency	5-37
         5.3.3  Testes	5-38
         5.3.4  Unresolved issues	5-41

    5.4   Nervous System	5-43
         (Michael I. Gage and Ernest N. Albert)
         5.4.1  Morphological and physiological observations ........ 5-44
         5.4.2  Blood-brain barrier studies 	5-46
         5.4.3  Pharmacological effects	5-48
         5.4.4  Effects on neurotransmitters	5-49
         5.4.5  Unresolved issues	5-50

    5.5   Behavior	5-53
         (Michael I. Gage)
         5.5.1  Naturalistic behavior	5-54
         5.5.2  Learned behavior	5-55
         5.5.3  Interactions with other stimuli	5-60
         5.5.4  Unresolved issues	5-61

                                   vi

-------
                       Contents   (Continued)

                                                                  Page

     5.6  Special Senses	5-64
          (Joe A. Elder)
          5.6.1   Cataractogenic effects	5-64
          5.6.2   Unresolved issues	5-69
          5.6.3   Auditory effects	5-70
          5.6.4   Unresolved issues	5-75
          5.6.5   Human cutaneous perception	5-76
          5.6.6   Unresolved issues	5-78

     5.7  Endocrine, Physiological and Biochemical Effects	5-79
          (Charles G. Liddle and Carl F. Blackman)
          5.7.1   Endocrine effects	5-79
          5.7.2   Clinical chemistry and metabolism	5-81
          5.7.3   Growth and development	5-84
          5.7.4   Cardiovascular system	5-85
          5.7.5   Biological effects of low-frequency modulation of
                 RF radiation	5-88
          5.7.6   Unresolved issues	5-92

     5.8  Genetics and Mutagenesis	5.94
          (Carl F. Blackman)
          5.8.1   Effects on genetic material of cellular and
                 subcellular systems	'.	5.95
          5.8.2   Effects on genetic material of higher-order
                 biological systems	5.99
          5.8.3   Unresolved issues	5-100

     5.9  Life Span and Carcinogenesis 	5-106
          (William P. Kirk)
          5.9.1   Life span	5-106
          5.9.2   Carcinogenesis	5-108
          5.9.3   Unresolved issues	-5-111

     5.10 Human Studies 	5-112
          (Doreen Hill)
          5.10.1   Occupational surveys/clinical studies	5-112
          5.10.2   Mortality studies	.5-114
          5.10.3   Ocular effects	5-116
          5.10.4   Congenital anomalies and reproductive
                  effects	5-119
          5.10.5   Unresolved issues	5-121

 6.  Summary and Conclusions	6-1
     (Joe A. Elder)

     6.1   Major Conclusions and Generalizations	6-1

     6.2   Specific Absorption Rate	6-3

     6.3   Core Temperature	6-7

     6.4   Summary	6-9

References	R-1

Glossary	G-1
                                        VII

-------
                      Contents  (Continued)
                                                                Page
Appendix
 A.  Science Advisory Board Subcommittee on the Biological
     Effects of Radiofrequency Radiation	A-1
 B.  Authors	B-1
                                 viii

-------
                               Figures
Number                                                          Page
 3-1   Far-field electromagnetic wave at a particular instant in time	3-3
 3-2   Power density vs. distance along axis from antenna aperture	3-4
 3-3   Interaction of RF radiation with electrical conductors,
       biological tissue, and electrical insulators	3-6
 3-4   Energy distribution in proximity to man at 1 GHz at the chest
       plane contour presentation	3-7
 3-5   Dielectric data for tissues in RF range 0.01 to 10 GHz	3-8
 3-6   Illustration of object size vs. wavelength dependence	3-9
 3-7   Calculated whole-body average SAR vs. frequency for three
       polarizations in a prolate spheroidal model of a human	3-10
 3-8   Absorption dependence on various ground and multipath
       factors	3-12
 3-9   Penetration depth as a function of frequency	3-16
3-10   ARD distribution in core of 6-cm radius multilayered
       sphere at 1650 MHz	3-17
3-11   Curve fitting of SAR data for a prolate spheroidal model of
       man for three basic orientations 	3-18
3-12   Effect of a capacitive gap on average SAR between the man
       model and the ground plane	3-18
3-13   A realistic block model of man	3-19
3-14   Absorption for man block model standing on ground plane	3-20
3-15   The real component of the complex permittivity of
       muscle as a function of frequency	3-21
3-16   EPA 2450-MHz anechoic chamber facility or 2.45-GHz
       far-field exposure facility	3-26
3-17   EPA 2450-MHz anechoic chamber facility: diagram of the
       microwave exposure facility	3-27
3-18   Diagram of absorber-lined horn	3-28
3-19   Diagram of a point-source compact range	3-28
3-20   Miniature anechoic chamber facility	3.30
3-21   Photograph of tapered exposure chamber at EPA facility	3-31
3-22   Facility for simultaneous exposure of 10 animals with
       minimal inter-animal interaction	 3-33
                                   IX

-------
                        Figures (Continued)

Number                                                          Page

3-23   Monopole-over-ground plane irradiation facility	3-33

3-24   Coaxial air-line system for high power exposures of
       cell cultures	3.35

3-25   Parallel-plate (microstrip) exposure system	3-36

3-26   Block diagram of complete RF near-field synthesizer	3-36

3-27   EPA 100-MHz rectangular strip line or Crawford cell	3-37

3-28   Exposure chamber of the circularly polarized 915-MHz
       waveguide facility  	3.37

3-29   Exposure chamber with associated instrumentation for the
       970-MHz circularly polarized waveguide facility at EPA	3-38

3-30   VHP resonant cavity facility	3-39

3-31   SRI multimodal cavity facility for primate irradiation 	3-40

3-32   Water-supply system for exposure chamber	3.42

3-33   Samples of commercially available survey meters for
       measuring RF electric-field strength	3-44

3-34   Microprocessor-controlled twin-well calorimeter	3-48

 4-1   Simple neural  model of thermoregulation in a mammal for
       predicting the motor responses to short-term changes in
       ambient or body temperature	4-2

4-2A   Current view of the principal systems employed in the
       regulation of body temperature (from Werner 1980)	4-4

4-2B   Overall relationship of the sensory, controller, effector, and
       passive system components of the the thermoregulatory
       system. Note that the overall structure of Figure 4-2A is
       retained (modified from Werner 1980)	4-4

 4-3   Effect of body mass on thermal.conductance of various
       mammals	  4-6
 4-4   Blood flow vs. skin temperature	4-7

 4-5   Effect of age on the preferred ambient temperature of mice	4-8

 4-6   Effect of Ta on  evaporation, conductance, metabolic  rate, and
       rectal temperature of dogs acclimatized to summer and winter	4-8

 4-7   Effect of T. on  evaporation, conductance, metabolic  rate, and
       rectal temperature of the rhesus monkey under unanesthetized
       and ketamine-anesthetized conditions	4-8

 4-8   Temperature distributions in the human body at Ta's of
       20 and 35°C 	4-9

 4-9   Effect of skin temperature on sweating	4-10

4-10   Effect of calorimetric temperature on rectal and skin
       temperature, heat loss, heat production, evaporation, and
       conductance in men and women	4-10

-------
                       Figures  (Continued)

Number                                                         Page

4-11   Effect of exercise on heat loss, metabolic rate, and tympanic
       temperature of humans	,... 4-11

4-12   Relationship between body temperature and set point during
       pyrogenesis	4-12

4-13   Power densities at 2450 MHz necessary to raise the rectal
       temperature by 1 °C in 60 min for the rat, squirrel monkey, and
       rhesus monkey	4-12

4-14   Examples of change in tail skin temperature in restrained mice
       exposed to 2450 MHz at 25°C and SAP of 10.6 W/kg
       or 20.0 W/kg	4-14

4-15   Evaporative water loss of mice exposed to 2450 MHz
       for 90 min 	4-15

4-16   Change in the rate of sweating from the foot of a squirrel monkey
       exposed for 10 min to 2450 MHz; the parameter is ambient
       temperature	4-15

4-17   Oxygen consumption and carbon dioxide production of rats
       immediately after a 30-min exposure to 2450 MHz at a
       T. of 24°C	4-16

4-18   Ventilatory frequency and preferred Ta of mice during exposure
       to 2450 MHz inside a waveguide-temperature gradient
       system	4-17

4-19   Mean Ta selected by one squirrel monkey exposed to
       2450-MHz RF radiation  	4-17

4-20   Effects of 5-HT injections on mice for various ambient
       temperatures and various power densities at 2450 MHz	4-19

4-21   Example of the triphasic rectal temperature response of a
       dog exposed to 2790-MHz RF radiation at a power density
       of 165 mW/cm2	4-21

4-22   Effect of an increasing THI on the lethal dose of RF radiation
       (2450 MHz) in mice	4-23

4-23   Relationship between the surface area:body mass ratio and
       the body mass of various mammals	4-24

4-24   Relation between metabolic rate and body mass of resting
       mammals under thermoneutral conditions	4-25

4-25   Relation between mass and percent increase in whole-body
       heat loss necessary to maintain normothermia in mammals
       exposed to RF radiation at three SARs	4-26

4-26   Effect of body mass on the maximum rate of nonevaporative
       heat loss relative to resting metabolism in mammals during
       exercise at a Ta of 20°C  	4-26

4-27   Relationship between SAR and Ta on the activation of
       various thermoregulatory effectors and elevation in body
       temperature in five animal species	4-26

4-28   Block diagram for one segment of the thermal model 	4-31

-------
                       Figures  (Continued)

Number                                                          Page

 4-29  Incident power density vs. exposure duration to obtain
       a hot spot	4-33

  5-1  Arrhenius plot of Na* efflux	 5-8

  5-2  Summed incidence of abnormal and nonviable chick embryo
       eggs exposed to 2450-MHz radiation	5-30

  5-3  Cross-sectional sketch of the human and the rabbit eye	5-66

  5-4  Time and power-density threshold for cataractogenesis in
       rabbits exposed to near-field 2450-MHz radiation	5.57

  5-5  Distribution of energy absorption rate per mW/cm2 incident
       power density in the rabbit's eye and head exposed to
       2450-MHz radiation  	5-67

  5-6  Distribution of energy absorption rate per mW/cm2 incident
       power density in the rabbit's eye and head exposed to
       918-MHz radiation	,	5-67

  5-7  Linear regressions of eye score on age for workers exposed to
       microwave radiation and controls	5-118
                                   XII

-------
                                Tables
Number                                                         Page
 3-1   Radiofrequency Bands 	3-2
 3-2   Proposed System of RF Dosimetric Quantities, Definitions,
       and Units	3-14
 3-3   Range of Resonant Frequencies of Man and Animals Irradiated
       by Plane Waves in Free Space at 1 mW/cm2 with Long Axis
       Parallel to Electric Field 	3-15
 3-4   Dielectric Permittivities for Various Tissues	3-21
 3-5   Energy Units for RF Radiation	3-22
 4-1   Partitioning of Heat Loss in Humans as a Function of
       Ambient Temperature Under Still Air Conditions	4-10
 4-2   Approximate Normal and Lethal Core Temperatures of
       Some Birds and Mammals	4-22
 4-3   Steady-State Temperatures in a Human Body After
       Exposure to 80- and 200-MHz RF Fields	4-32
 5-1   Classification of Cellular and Subcellular Experiments	5-1
 5-2  ' Summary of Studies Concerning RF-Radiation Effects on
       Molecular Systems  	5-3
 5-3   Summary of Studies Concerning RF-Radiation Effects on
       Subcellular Systems	5-5
 5-4   Summary of Studies Concerning RF-Radiation Effects on
       Single Cells	5-7
 5-5   Summary of Studies Concerning Hematologic Effects of
       RF-Radiation Exposure	5-15
 5-6   Summary of Studies Concerning Immunologic Effects
       (In Vivo) of  RF-Radiation Exposure	 5-19
 5-7   Summary of Studies Concerning Immunologic Effects
       (In Vitro) of RF-Radiation Exposure	5-26
 5-8   Conversion of J/g to W/kg	5-31
 5-9   Summary of Studies Concerning Teratologic Effects of
       RF-Radiation Exposure	5-37
5-10   Summary of Studies Concerning Reproductive Effects of
       RF-Radiation Exposure	5-38
5-11   Summary of Studies Concerning Effects of RF-Radiation
       Exposure in Testes of Mice and Rats	5-41
5-12   Summary of Studies Concerning RF-Radiation Effects on
       the Nervous System 	5-50

-------
                       Tables  (Continued)
Number                                                         Page
5-13  Summary of Studies Concerning RF-Radiation Effects on
      Behavior	5-62
5-14  Summary of Studies Concerning Ocular Effects of
      Near-Field Exposures 	5-65
5-15  Summary of Studies Concerning Ocular Effects of
      Far-Field Exposures	5-65
5-16  Summary of Studies Concerning Auditory Effects of
      RF Radiation in Humans	5-71
5-17  Summary of Studies Concerning Threshold Values for
      Auditory-Evoked Potentials in Laboratory Animals	5-76
5-18  Summary of Studies Concerning Human Cutaneous
      Perception of RF Radiation	5.77
5-19  Summary of Studies Concerning RF-Radiation Effects on
      Endocrinology	5-80
5-20  Summary of Studies Concerning RF-Radiation Effects on
      Clinical Chemistry and Metabolism	5-83
5-21  Summary of Studies Concerning RF-Radiation Effects on
      Growth and  Development 	5-86
5-22  Summary of Studies Concerning RF-Radiation Effects on
      Various Aspects of Cardiac Physiology	5-87
5-23  Summary of Studies Concerning Biological Effects of
      Low Frequency Modulation of RF Radiation	5-93
5-24  Summary of Studies Concerning Genetic and Mutagenic
      Effects of RF-Radiation Exposure	5-103
5-25  Summary of Studies Concerning RF-Radiation Exposure
      Effects on Life  Span/Carcinogenesis	,	5-107
5-26  Summary of Prausnitz and Susskind Data	5-107
5-27  Summary of Selected Human Studies Concerning Effects of
      RF-Radiation Exposure	5-113
5-28  Distribution of  Years of Exposure for 226 Radar Workers	5-113
5-29  Age Distribution of 226 Microwave Workers and 88 Controls ... 5-113
5-30  Microwave Exposure Levels at the U.S. Embassy in Moscow	5-115
5-31  Number of Deaths from Disease and Mortality Ratios by Hazard
      Number:  U.S. Enlisted Naval Personnel Exposed to Microwave
      Radiation During the Korean War Period	5-116
5-32  Classification by Military Occupation of World War II and
      Korean War Veterans With and Without Cataracts, Based on
      Discharges from VA Hospitals 	5-117
5-33  Estimated Relative Risk of Cataracts Among Army and
      Air Force Veterans by Age Group and Occupation 	5-117
5-34  Paternal Radar Exposure Before Conception of Index Child	5-120
                                 XIV

-------
                        Tables   (Continued)
Number                                                        Page
6-1  Selected Studies Reporting "No Effects" at SARs
     < 10 W/kg Grouped by Biological Variable	 6-5
6-2  Selected Studies with Reported "Effects" at SARs
     < 10 W/kg Grouped by Biological Variable	6-6

-------
                         A cknowledgment


The authors thank Jan Parsons and Carole Moussalli, Northrop Services, Inc., for
editorial  review of  this document. The authors wish to express special
appreciation to Dr. Don R. Justesen for his editorial comments on an earlier draft
and Dr. Charles Susskind for his editorial assistance on the final draft.
                                 XVI

-------
                                           Section 1
                                          Introduction
Daniel F. Cahill
Joe A. Elder
The goal and purpose of this document is review and
evaluation of the available scientific information on the
biological effects of radiofrequency (RF) radiation. To
address this  broad topic  most  effectively  and to
present the information in the most useful form,
several guidelines and simplifying assumptions were
adopted. It is in the selection and employment of
these guidelines and assumptions that differences of
opinion are most likely to arise. These guidelines and
assumptions are as follows:
1. The frequency range of interest is  defined as 0.5
   MHz to 100 GHz. This range includes nearly all the
   frequencies that  serve  as  significant sources of
   population exposure (e.g., AM, FM, land  mobile,
   and amateur radio; UHF and VHFTV; and air traffic
   control radar).
2. The general  public is the population of concern.
   Therefore, far-field exposures are the more usual
   condition,  rather than  the near-field exposures
   commonly associated with  the workplace.
   Wherever  possible,  the impact on  all  ages is
   considered, not only on  healthy adults, who are of
   principal concern in occupational-exposures.
3. Dose rate is defined by the specific absorption rate
   (SAR), which is used to normalize  the rate of RF
   energy input  into biological systems across this
   frequency range. The SAR is the mass-normalized
   rate  at which the energy of an electromagnetic
   (EM) field is  coupled into an absorbing body; the
   units are watts per kilogram (W/kg). Inherent in
   the use of a single, whole-body-averaged  SAR
   value is a disregard of nonuniform RF  energy
   deposition patterns and possible local SAR values
   in excess of the average. However, the data  that
   describe the  distribution patterns of RF energy in
   biological systems over a frequency band as broad
   as 0.5 MHz to 100 GHz are far too incomplete at
   this time to be directly useful.
4. In reports of pulsed  RF-radiation exposure
   experiments the time-averaged SAR is considered,
   although controversy exists as to whether some
   biological effects are functions of the temporal
   pattern of energy delivery or only the result of total
   energy input.
5. We have accepted "no effects" data  in credible
   reports as highly valuable in our review of the state
   of knowledge in this area. Both "effects" and "no
   effects" studies are used to determine the extent
   of the biological interactions of RF radiation.
6. We are assuming that data derived from the use of
   experimental mammalian systems have relevance
   for the human situation.
7. This document presents only the biological effects
   of RF radiation. The benefits of RF radiation are
   not considered,  and therefore  no benefit/risk
   analysis is undertaken.
A draft of this  report (Cahill and Elder  1983) was
published in June 1983 and transmitted to EPA's
Science  Advisory Board (SAB)  for  review of its
scientific and  technical merit  (Federal Register
1983a). The initial review by the SAB Subcommittee
of the Biological Effects of Radiofrequency Radiation
was held  on September  22-23, 1983 (Federal
Register  1983b);  the second review occurred  on
January 24-25, 1984 (Federal Register 1984). A list
of the Subcommittee members is given in Appendix
A.
The  June 1983 draft was a critical  review of the
literature on the biological effects of radiofrequency
radiation through 1980. The present document is a
revision of the June 1983 draft and is based on the
comments  and suggestions  of the Subcommittee,
and includes a number of post-1980 references that
were considered important to the conclusions of the
present review of the biological effects of RF radiation
by both  the authors  and by the Subcommittee
members.
                                               j-j

-------

-------
                                           Section 2
                                           Approach
Joe A. Elder
Daniel F. Cahill

2.1 General Approach
Although a comprehensive literature review  is
useful,  it is even more desirable  if the body  of
literature is consolidated, analyzed, and synthesized
into a statement or statements  that relate the
presence or absence of biological effects to  a
meaningful exposure parameter such as  dose rate
(SAR). To this end, our general approach is essentially
as follows:

1. The  reports are  evaluated  for their  scientific
   quality and utility. Acceptable  reports contain
   adequate descriptions of appropriate physical and
   biological systems and tests.
2. The credible reports are then examined for the
   relation between the RF energy absorbed and the
   presence or absence of biological effects in the
   experimental systems.

2.2 Specific Approach
The literature evaluated for this document includes
English-language  publications,  numerous English
translations of Soviet research reports  obtained
through the U.S. Joint Publications Research Service,
and selected technical reports translated from Polish,
French, Italian, and Russian.
The  extant  literature base numbers  over 5000
citations, but considerably fewer  are  valuable
in developing exposure guidelines for the  following
reasons.
1. A large fraction of the literature is available only in
   Slavic languages.
2. The research reports are  uneven in quality and
   usefulness because authors often  failed to
   include sufficient experimental details to allow
   reviewers to estimate critical exposure parameters
   such as incident field intensity, dose rate, or dose.
3. In many review articles, equal currency is given to
   the conclusions  from properly  designed  and
   executed studies and to those from less stringently
   conducted  research. Because these uncritical
   reviews have contributed to the general confusion
   over the health risks associated with RF radiation.
   we have relied on original research papers rather
   than review articles and have chosen to be highly
   selective in our review.
 In  reviewing  the literature  we  first  evaluate
 descriptions of frequency, exposure parameters, RF
source, experimental species, age, sex environmental
 and biological controls, and the statistics employed;
 and whether  actual data are displayed or merely
 referred to in the text. Many pre-1970 reports are
 deficient in one or more of these key areas and are
 either rejected outright if the flaws are judged fatal, or
 are  segregated  into the category of reports with
 "unresolved issues." Reports  that provide adequate
 descriptions  of these  parameters are further
 scrutinized  for  the appropriateness of biological
 systems, tests, sample sizes, controls, and statistics
 employed,  as  well  as  substantiation  of  their
 conclusions. Reports that clear this second hurdle are
 credible reports, but are considered usable only if
 biological  results are linked explicitly or implicitly to
 SAR data from the description  of the  exposure
 parameters.  Reports that fail  to  provide  these
 parameters  are also  assigned to the "unresolved
 issues" category.


2.3 Major Sections
Four major sections follow. They are  Physical
Principles  of Electromagnetic Field Interactions (Sec.
3),  Effect of RF-Radiation  Exposure on  Body
Temperature  (Sec. 4),  Biological  Effects of RF
Radiation  (Sec. 5), and Summary and Conclusions
(Sec. 6).
The  first part of  Sec. 3 presents some introductory
information on electromagnetic field theory. Next,
RF-field interactions with both simple and complex
biological  objects,  such as the  human body, are
discussed  and, most  important, definitions  of RF
dosimetric terms are given. The mechanisms of RF
interactions with biological systems, particularly
molecular systems, are discussed. The remaining two
major subsections described experimental methods
and  dosimetric methods used in state-of-the-art
research on the biological effects of RF radiation.

Since the absorption  of RF  energy by biological
species can lead to an increase in the temperature of
body  tissues,  it is  important to  understand how
                                               2-1

-------
animals, including man, regulate the  additional
thermal input of RF-radiation  exposure,  both
physiologically and behaviorally.  Section 4 is  an
introduction to the subject of thermoregulation in
both animals and human beings  and, in addition,
discusses the specific effects of RF radiation  on
thermoregulatory processes. Also, a description is
included of the mathematical models that are being
used to predict increases in temperature and
activation of thermoregulatory effectors  in  human
beings in simulated RF fields.

Section 5 is a review of the main body of literature on
the biological effects  of  RF radiation. The  section
contains ten subsections, each of which represents a
biological discipline or major research area, ranging
from subcellular systems to human beings. In each of
the ten areas, the  conclusions and generalizations
that can be drawn from the review of the literature are
presented.

In Sec. 6, the major conclusions and generalizations
of Sees. 3, 4, and 5 are presented.  Next, many of the
reports are tabulated by biological variable and dose
rate (SAR).  In summary, the reported consequences
of the interaction between RF radiation and biological
systems are examined from two perspectives (whole-
body-averaged SAR and RF-energy-induced core
temperature increases) in order to analyze, synthesize,
and consolidate the review data into statements that
relate biological effects to  a  meaningful exposure
parameter (dose rate or SAR).
                       2-2

-------
                                          Section 3
                Physical Principles of Electromagnetic Field Interactions
3.1  Elements of Electromagnetic Field
Theory
William T. Joines
3.1.1 Electromagnetic Spectrum
3.1.1.1 Frequency, Wavelength, and Velocity
Oscillating electric charges induce an electromagnetic
(EM) field within the region surrounding the charge
source. In turn, this oscillating induction field—often
called the near field—of the source generates an EM
wave that radiates energy from the region surrounding
the charges. The radiated wave consists of coupled
electric and magnetic fields that oscillate at the same
frequency as the source,  and the wave  propagates
outward from the source at the velocity of light in the
medium. In free space, this velocity is~ 3 x 108 m/s,
whereas in a medium with low EM energy dissipation
the velocity is this value divided by the square root of
the material's dielectric constant relative to that of
free space. In a low-loss material such as fatty tissue,
with a relative dielectric constant of 4, the velocity of
light is 1.5 x 108 m/s. For an EM wave traveling at
velocity v, the  wavelength in the medium  is the
distance the wave travels in one time period or T = 1 /f,
where f  is the oscillation  frequency of the wave in
cycles per second, or hertz (Hz). Therefore, the
wavelength ^ is expressed as
                    * =v/f
The EM spectrum extends from zero to  1025 Hz (or
cycles/s) and includes the ranges of visible light and
ultraviolet (UV), infrared (IR), X, and gamma rays. The
region from  zero to 3000 GHz (lower IR range) is
referred to as the RF region (Sams & Co. 1981). The
region of interest here is the RF band from 0.5 MHz to
100  GHz. The wavelengths in free  space for this
frequency band range from 600 m to  3 mm.

3.1.1.2 Ionizing and Nonionizing
Electromagnetic Radiation
EM waves of all frequencies carry energy. According
to quantum mechanics, they can also be thought of as
packets of energy called photons. The energy of a
photon is given by
                  Energy = hf

where h =  6.63 x  10"34  joule  seconds (Planck's
            constant)
       f  = frequency
The energy of a photon is thus directly proportional to
the frequency of the radiation. When the frequency
approaches or exceeds 3 x 10'6 Hz (UV), the photon
energies equal or exceed 2x10"18 J, or 12.4 eV, and
become comparable to the binding  energy  of
electrons to atoms. This  high-frequency radiation (X
rays,  gamma rays, etc.) is referred to  as ionizing
radiation.  Since  even  the weakest  chemical bonds
have energies that are several orders of magnitude
greater than those of RF  or microwave photons (10~3
eV or less), RF waves are referred to as nonionizing
radiation.
3.1.1.3 Designation of Microwave and
Radiofrequency Bands

Bands of radiofrequencies have been assigned
designations according to frequency or wavelength
as shown in Table 3-1 (Sams & Co.  1981). Typical
uses  of  the frequencies within a band  are also
indicated. Certain radiofrequencies—notably the
industrial, scientific, and medical (ISM) frequencies—
have been assigned by the Federal Communications
Commission for  specific applications  (Sams & Co.
1981). The ISM frequencies are

             13.56 MHz ±6.78 kHz
             27.12 MHz± 160kHz
             40.68 MHz ± 20 kHz
               915 MHz ±25 MHz
              2450 MHz ± 50 MHz
              5800 MHz ± 75 MHz
            22,125 MHz± 125 MHz


3.1.2 Wave Propagation

Power Density, Electric Field, and Magnetic Field

In free space, EM waves spread uniformly  in all
directions from a theoretical point source. The
wavefront, or the surface joining all points of identical
phase, is spherical in this case. As the distance from
the point source increases, the area of the wavefront
surface increases as a square of the distance, so that
the source power is spread over a larger area. If power
density is defined as the ratio of the  total radiated
power to the spherical surface area enclosing the
source, the power density W is inversely proportional
to the square of the distance from the source, and can
be expressed as
                                              3-1

-------
Table 3-1.    Radiofrequency Bands*

Frequency                  Wavelength
                        Band Designation
                             Typical Uses
300-3000 GHz


30-300 GHz



3-30 GHz



0.3-3 GHz





30-300 MHz




3-30 MHz





0.3-3 MHz



30-300 kHz



3-30 kHz

0.3-3 kHz

30-300 Hz


0-30 Hz
1-0.1 mm


10-1 mm


10-1 cm



100-10 cm





10-1 m




100-10 m





103-102m



10-1 km


100-10 km

103-10'km

104-103km


oe-104 km
Supra EHF
(extremely high
frequency)
Extremely high
frequency (EHF)

Super high
frequency (SHF)

Ultra high
frequency (UHF)
Very high
frequency (VHF)
High frequency
(HF)
Medium frequency
(MF)
Low frequency
(LF)

Very low
frequency (VLF)
Voice frequency
(VF)

Extremely low
frequency (ELF)
Sub-ELF
Not allocated
Satellite communications,
radar, microwave relay.
radionavigation, amateur radio
Satellite communications, radar.
amateur, taxi, police, fire,
airborne weather radar, ISM
Microwave point to point,
amateur, taxi, police, fire
radar, citizens band, radio-
navigation, UHF-TV, microwave
ovens, medical diathermy. ISM
Police, fire, amateur, FM,
VHF-TV, industrial RF equipment
diathermy, emergency medical
radio, air traffic control
Citizen band, amateur, medical
diathermy. Voice of America,
broadcast, international
communications, industrial RF
equipment
Communications, radio-
navigation, marine radiophone,
amateur, industrial RF
equipment, AM broadcast
Radionavigation, marine
communications, long-range
communications
Very long range communications,
audiofrequencies, navigation
Voice, audiofrequencies

Power lines, audiofrequencies,
submarine communications
Direct-current power lines
*Sams& Co. 1981.
                    W = P/477T2               (3-1)
 where P = transmitted power
        r= distance from the source
 An  isotropic source  radiates  uniformly in  all
 directions in space. In practice, there is no source that
 has this property, although there are close approxi-
 mations. The inverse-square  law (Equation 3-1)
 applies  also when  the  source is anisotropic
 (directional), as long as the medium is homogeneous
 and isotropic (e.g., free space or a medium in which
 the velocity of wave propagation does not change
 with direction or distance).
 An EM wave is also characterized by its electric-field
 (E) intensity and its magnetic-field (H) intensity. In an
 unbounded medium, the product of E and H  is the
 power density W, or
                      W = EH                 (3-2)
                             where E is in volts per meter, H  is in amperes per
                             meter, and W is in watts per square meter. If the peak
                             instantaneous  values of  E and  H  are  used in
                             Equation 3-2, then W is the peak power density; if
                             effective or root-mean-square (RMS) values of E and
                             H are used, then W is the effective or time-averaged
                             power density.  The ratio of E to H is the intrinsic
                             impedance of the medium  rj (in ohms), or
                                                   /7 = E/H                  (3-3)
                             where rj is 120w ohms in free space. From Equations
                             3-1 through 3-3, one can express E in free space at a
                             distance r from an  isotropic source radiating total
                             power P as
                                                  E =V3OP/r               (3-4)
                             The radiated waves that propagate outward from an
                             RF  or  microwave source can be  confined to travel
                             along  a two-conductor  transmission  line (coaxial.
                          3-2

-------
 stripline, microstrip, twin lead), or inside a hollow
 metal pipe (waveguide).  The waves can  also be
 propagated outward into the space that surrounds a
 transmitting antenna, as in communications broad-
 casting.  In this case, the RF or microwave source
 forces electrons to oscillate on the surface of a metal
 transmitting antenna and thereby produces EM
 waves. This is  a reciprocal process, for when these
 waves strike a receiving antenna, they force electrons
 to oscillate, which typically produces current in the
 receiver.  This  reciprocity phenomenon forms the
 basis of EM communication systems. Information can
 be placed on an EM carrier in several ways, including
 amplitude and  frequency  modulation of the carrier
 wave.

 3.1.2.1  Sources of Radiofrequency Radiation
 There is widespread interest in possible new
 applications of RF energy, especially at microwave
 frequencies. The growing number of  commercial
 applications has led to an increased awareness of
 potential  hazards due to the  energy sources used.
 Typical sources of RF  or microwave  energy are
 klystrons,  magnetrons,  planar  triodes,  backward-
 wave oscillators, and semiconductor devices. In all
 these sources, energy is imparted to charged particles
 or electrons from a convenient supply, usually a direct
 voltage and current. A portion of the energy is then
 given up by the charged  particles  in the  form of
 oscillations in  a tuned circuit.  Such sources may
 operate to produce (1) continuous (CW) radiation, as
 in the case of  some communications systems; (2)
 intermittent  radiation,  as in  microwave ovpns,
 induction heating equipment,  and diathermy equip-
 ment; or (3) radiation in the pulsed(PW) mode in radar
 systems.

 3.1.2.2 Plane-Wave, Far-Field, and Near-Field
 Concepts
 Usually  the region  that is more than a few
 wavelengths from the transmitting antenna is called
 the far field. In this region, the  spatial relationship
 between  the electric and magnetic fields in the EM
 wave is that shown in Figure 3-1. In this diagram, an
 EM wave is propagating in a direction perpendicular
 to the motion  of the electrons  in the transmitting
 antenna. The electric field E is always perpendicular
 to the direction of propagation and lies in the plane
 formed by the line of propagation and the motion of
 electrons in the antenna. The direction of the electric
 vector periodically changes along the direction of
 propagation, and its magnitude  forms a  sine-wave
function.  The  magnetic field  H, which  is  always
 perpendicular to both the electric field  E  and the
 direction  of propagation,  traces out a  sine-wave
function in the same manner as the electric field E.
 Figure 3-1 depicts the situation at a particular instant
 in time. The entire wave can be pictured as moving in
the direction of  propagation at ~ 3 x 108  m/s.
 Figure 3-1.  Far-field electromagnetic wave at a particular
           instant in time.
                                  Velocity, v
 At some distance from the transmitting antenna
 (within the far field), the radius of curvature of the
 generally spherical  wavefronts is large compared
 with the diameter of any receiving or detecting object
 within the field. At this distance from the antenna the
 wavefronts may be treated (for all practical purposes)
 as plane surfaces over which the intensity of E and H
 is constant. In such regions, E and H form plane
 waves. The terms "plane-wave field" and "far field"
 are often  used interchangeably,  although the
 distinction between the two terms depends on the
 size of the receiving object, as stated above.
 At distances less than  a  few wavelengths  from a
 transmitting antenna, in the near field, the situation
 is somewhat complicated  because the  maxima and
 minima of E and  H do  not occur at the same points
 along the direction of propagation as they do in the
 far-field case  (Figure 3-1).  Near-field exposures
 become particularly important when one is consider-
 ing  radiation  from microwave ovens,  microwave
 diathermy equipment, RF sealers, broadcast antennas,
 and microwave oscillators under test. For investiga-
 tions into the biological effects of RF-radiation
 exposure, studies of exposures in the  far field are
 usually preferable; field strengths in the nearfield are
 more difficult to specify because both E and H must be
 measured and  because the field patterns are more
 complicated.

Incident power  density at a given distance from an
antenna may be  calculated from the measured
power transmitted by the  antenna,  the  known
antenna gain G, and effective area A of the antenna.
Antenna gain is defined as the power density at a
point in front of the antenna divided by the  power
density  at  the  same point  if the  antenna were
radiating the same total power as an isotropic source.
For any impedance-matched antenna (i.e., one where
all the energy fed to it is transmitted), the ratio of G to
A is
                 G/A = 47T/M2
(3-5)
                                                                       3-3

-------
The far-field power density is calculated from the Friis
free-space transmission formula (Mumford 1 961 ) as

             W = GP/477T2 = APA*2r2         (3-6)

where ^ = wavelength
       P = power output of the antenna
       W = power density on a surface at distance r
       from the antenna

It is  convenient to express the far-field free-space
power density in terms of the power density at the
antenna aperture, i.e., W0 = P/A, and hence P = WoA.
Substituting this expression for Pin Equation 3-6 and
dividing both  sides by W0 yields

                W/Wo = (A/^r)2            (3-7)

Equation 3-7  may be rewritten as
         Figure 3-2.  Power density vs. distance along axis from
                  antenna aperture.
This expression applies  in the far-field region, and
the following simple modification makes it applicable
to the near-field region as well:

             W/W0 = 4sin2( A/2,* r)          (3-9)

This formula applies to uniform irradiation of square,
round, or rectangular apertures or to irradiation that
is tapered in amplitude for round apertures. For other
shapes and tapers, a more complicated analysis is
necessary (Mumford 1961).

Since the power density at the antenna aperture (Wo =
P/A) is greater than the output power P divided by the
cross-sectional area of  the aperture,  the effective
area A is less than the actual area. For the waveguide
horn and dish antennas commonly used at microwave
frequencies, the effective area ranges from about 50
to 80 percent of the actual area of the cross section.
For a given P, with A determined for the particular
antenna used, W0  = P/A is used in Equation 3-9 to
compute W at  any  distance  r from  the  antenna
(Mumford 1961).

Equation 3-9 is plotted in Figure 3-2 along with lines
representing the maximum values in the near-field
and the far-field approximations as given by Equation
3-7. In this graph, the relative power density is plotted
in decibels [dB = 1 0logio (W/Wo)] on the ordinate, and
^r/A is plotted logarithmically on the abscissa. Note
the alternate maxima and minima in the near field.
The maxima all are 6 dB above (4 times) the power
density at the aperture. For- a conservative estimate
in possible near-field human exposure situations, the
power density can be approximated as the maximum
value or
                W = 4W0 = 4P/a
(3-10)
Figure 3-2 and Equation 3-9 also show that for ^r/A =
1, W = W0, the  power  density at the antenna
aperture. Hence, a convenient point of demarcation
  10


€  5
2f
        O
        0  0
        I  "5
        J5
        i.o
                                           I3'8'  t.15
              III!
                                          -.Iv/M'-H I f
                                         * * , ** v< <. Mi.,, "* "^
                                      Locus of Maxima
                                      -^  W/W0 =
                                   Far Field
                               W(r)/W0 = (A/Mr)2 j
          -20l I  I I I
                  I	I I  I III
                0.10
                 0.2  0.4 0.6 1.0
4.0  6.0  10
 between the near field and far field is the distance
 from the antenna where

                      r=A/*                (3-11)
 The gain of an antenna as used in Equations 3-5 and
 3-6 is usually taken to be the gain in the direction of
 maximum  radiated power. In general, the antenna
 will have a radiation pattern that is at a maximum in
 front of the antenna on an axial line perpendicular to
 the aperture surface. On a surface of radius r from the
 antenna, the power density typically decreases as the
 off-axis angle increases. The relative power density is
 expressed  as W(0)/Wma«, where Wmax is the power
 density at a given distance along the axis of maximum
 radiation, and W(0) is the power density at the same
 distance from the center of the aperture but at an off-
 axis  angle 0. Equations 3-6  through  3-9 apply to
 angles off-axis if their right-hand sides are multiplied
 by the  radiation pattern  W(0)/Wma«. While the
 radiation pattern may be known or determined, it
 does depend upon the type of antenna that is used.


 3.1.3 Wave Modulation

 3.1.3.1. Amplitude and Frequency Modulation
 With an EM wave of a particular frequency serving as
 a carrier, information signals at lower frequencies are
 superposed onto the carrier through the process of
 modulation. Modulation of the  carrier can  be
 accomplished either by continuous variation of one of
 the parameters  of the  carrier (amplitude, frequency,
 or phase) or by interruption of the carrier by a process
 in which it is chopped into segments (pulse wave or
 PW modulation). If the impression of the information
                       3-4

-------
signal on the carrier causes its amplitude to vary, the    Long-range and short-range radars may thus be of
process is called amplitude  modulation (AM). If the    comparable average transmitted power.
information signal causes a frequency variation of the
carrier at constant amplitude, the process is called
frequency modulation (FM). If the information signal
causes a  time-phase variation  of the carrier,  the
process is called  phase  modulation (PM). Phase
modulation and  frequency  modulation  are closely
related; both cause a change in the time phase angle
of the carrier (Froehlich 1969).
Since the purpose of modulation is to improve the
overall efficiency of transmission, the carrier
frequency is chosen for its suitability tothe medium in
which it is to be transmitted. Transmitting unmodified
audio signals is theoretically possible; however, the
antenna required to transmit  and  receive such
signals  would be  too large to  be  practical.  More
important, unmodulated signals  are not transmitted
because no other method exists  to separate several
information signals on the same transmission path to
provide multichannel capability. This flexibility is
obtained if carriers of higher frequencies are chosen
and  then spaced  in frequency so  as  to  prevent
overlap.  High-frequency carriers  are  also  more
suitable for radio propagation. One must then shift
the information  signal  to a high-frequency  range
through  the  process  of  modulation. Introducing
frequency translation to information signals provides
a way of separating information channels and
thereby improving the transmission efficiency.

3.1.3.2 Pulse Modulation
One of the  more familiar applications of  PW
modulation is in radar. Radar is possible because EM
waves at some frequencies  are  reflected by certain
materials, including metal surfaces and water vapor
in clouds. A radar antenna can transmit and receive,
and it is directional; that is, it transmits and receives
signals in a narrow beam. For most  purposes, the
radar antenna sends out short pulses of EM waves
and later receives the reflected pulses. The interval
between  the outgoing  and incoming pulse is
correlated with the beam direction to pinpoint the
position of the reflecting objects, such as airplanes or
cloud formations. The incoming reflected pulse is
smaller than the outgoing pulse by a factor of millions
and sometimes billions. To detect an incoming pulse
in the presence of atmospheric  noise, the power of
the outgoing pulse is made as large as possible.
The duty factor (DF) of a transmitter is the ratio of the
width of  the transmitted pulse to the  time between
the leading  edges of  consecutive  pulses. Hence,
average power (Pav) is determined from peak power
     ) by
                 Pav = (DF)Ppeak

In general,  long-range radars have a  larger
transmitted pulse power and a longer waiting time
between pulses (lower DF) than short-range radars.
                                                                       3-5

-------
3.2 RF-Field Interactions with Biological
Systems
Claude M. Weil
James R. Rabinowitz

Consideration of the coupling of RF radiation into
various biological objects (human, animal, etc.) is
important to any study of the biological effects of RF-
radiation exposure. From  measurements  and
predictions of RF-energy  absorption based on
experimentation with laboratory animals, one can
extrapolate data for estimating effects of RF radiation
on humans. Furthermore, an adequate understanding
of the  interaction of RF  radiation with humans has
been crucial in determining exposure limits recom-
mended in the American National Standards Institute
(ANSI)  C95.1 standard (ANS11982). Relating a health
effect  to a  maximum permissible incident  power
density requires an in-depth understanding of the
interaction of RF energy with humans.

3.2.1 Scattering and Absorption of
Electromagnetic Waves
The interaction of  EM  waves with  any irradiated
object is a complex event that forms a large part of the
study  of electromagnetism. When an  EM wave
propagating  in air impinges normal to the surface of
any material with dielectric properties different from
those  of air, a  reflected  wave  created at the air
dielectric interface propagates in a direction opposite
to that of the incident wave.  This interaction  is
illustrated  in Figure 3-3 for three materials  of
differing electrical properties. At the top of the figure,
an  EM wave  is shown incident  on  a metallic
conductor.  In this  case,  virtually  all the incident
energy is  reflected back  to  its source by the
conductive plate. At the bottom of the figure, the
energy is shown  incident on a slab of  low-loss,
nonconductive  material  (dielectric insulator) where
almost no energy is dissipated within the material. In
that case part of the incident energy is transmitted
through the slab, part is reflected back at the first air-
dielectric interface,  and  a  part is absorbed—a small
amount, since this material exhibits almost no
dielectric losses. An intermediate case, shown in the
middle of Figure  3-3,  is the  interaction with  a
"complex" dielectric. Biological tissue falls between
the conductor and the nonconductor in Figure 3-3;
i.e., it exhibits values of relative permittivity (a
measure  of how strongly the. material reflects
energy) and conductivity (how  well the material
conducts electrically) that fall between the very high
values for a metallic conductor and the near-zero
values for a dielectric insulator. If the tissue slab is of
sufficient thickness, all incident energy is reflected or
absorbed and no energy is transmitted through the
slab. The mechanism by which RF energy is absorbed
or dissipated within-the tissue slab is discussed in
more detail  below.
Figure 3-3.  Interaction of RF radiation with electrical
          conductors, biological tissue, and electrical
          insulators (modified from Sher 1970).
Incident
Microwave
Power
Incident

Microwave
Power
                         Total
                     Reflection
                               Tissue
                         Part
                     Reflection
Incident
                     Reflection
                            Nonconductor
                         Part k\\\\NNS* Part
Microwave
Power
                                     Transmission
In general, the reflected and transmitted waves that
surround a dielectric nonplanar object irradiated by
an  incident RF  wave together  comprise what is
termed the scattered field. Scattering by a dielectric
object and any absorption within the object represent
the two  basic components of the EM  interaction
phenomenon. The degree, to which any object
interacts with an EM field when irradiated  by a
uniform plane wave is defined  in terms of scattering
and absorption cross sections. These cross sections
represent the cross-sectional area in square meters
of an equivalent flat  plate placed normal to the
direction of propagation that totally reflects (in the
case of the scattering cross section) or totally absorbs
(in the  case of  the absorption cross section) the
radiant energy incident on the plate.

In general, the EM cross section is smaller than the
geometric (optical) cross  section that the object
creates when illuminated by a distant light source.
However, under certain conditions—termed reson-
ance—when the wavelength of the incident radiation
is comparable to the physical dimensions of the
object, the scattering and absorption  effects are
enhanced,  so that both respective cross sections
exceed the geometric cross section. Resonance is a
significant phenomenon that  will be discussed in
more detail  in  the following sections;  it can  be
explained by the apparent ability of the resonant
                        3-6

-------
Figure 3-4.  Energy distribution in proximity to man at 1 GHz at  the chest plane contour presentation; A) vertical.
            B) horizontal (Reno 1974).
                                                         Radiation
                                                         Radiation
                                                      Field Amplitude
          0.0-0.1      0.1-0.2      0.2-0.3      0.3-0.5     0.5-0.7      0.7-1.0     1.0-1.3      1.3-1.6      1.6-2.0
                                                          nj/m3
                                                                                         3-7

-------
object to intercept or "pull in" considerably more of
the energy incident upon it relative to the optical case.
Figure 3-5.  Dielectric data for tissues in RF range 0.01 to 10
          GHz. a: permittivity, b: conductivity.
The terms "scattering coefficient" and "absorption
coefficient," referring to measures of the efficiency
with which an object scatters and absorbs energy, are
defined as the ratio of the scattering to optical cross
sections and the absorption to optical cross sections,
respectively. Of the two basic measures of interaction,
the absorption  coefficient is considerably more
prominent in  the interaction of EM waves with
biological bodies. Potentially adverse biological effects
are related to the energy that the body is absorbing,
not to what is scattered. Because absorption rate is
related to internal field strength, it is conceivable that
a biological effect could be related to the field strength
within the tissue. For this reason,  most of the
remaining material in this section deals only with the
absorptive aspect of the interaction phenomenon.

Some  measurements have been   made  of  the
scattered fields that surround both human subjects
and life-size  phantom models exposed to  incident
microwave fields (Reno 1974). Although this work
illustrates well the complex scattered fields
surrounding the irradiated subject (see Figure 3-4),
no attempts have yet been made to relate the field
patterns to what the subject is absorbing. Although it
is theoretically possible to determine the  energy
being absorbed and the  internal distribution from
measurements  of the distribution  of externally
scattered fields,  in  practice  this  determination
remains a difficult task and has  not yet been
attempted. Interest in this area is  growing, as
evidenced by a recent symposium (IEEE 1980)  that
addressed methods of obtaining high-resolution
images of the internal dielectric  structure of a
biological target. The  methods  are based on
techniques of probing the scattered  fields created
when the target is irradiated by an RF field. The scope
for future work in this area appears large, particularly
in diagnostic applications.

3.2.1.1 Factors Affecting the Absorption of RF
Energy by an Irradiated Subject
The greater the efficiency of EM energy coupling into
a biological subject the greater the  overall whole-
body absorption.  The  factors that  influence the
degree of absorptive coupling are listed below and
will be discussed in turn:
a. Dielectric composition of subject
b. Object size relative to wavelength of incident field
c. Shape or geometry of subject and  its orientation
   with respect to polarization of incident field
d. Complexity of  incident radiation

a. Dielectric composition of subject—As mentioned
   previously,  absorptive coupling to the irradiated
   10H
 3  102
at
a.
S5
*
cc
   10°.
         Brain
                                              a
         Bone
      10'
0*            103
 Frequency (MHz)
                                              104
    10H
 £   10°-
G
>-
§
o
   10-'-
   10-2
      10'
                                 103
                          10*
                    Frequency (MHz)
   subject can  take place only  if the subject is
   composed of dissipative dielectrics of  finite
   conductivity value. The dielectric properties of all
   biological tissues fall into this category, primarily
   because their major constituent is water (average
   of 70 to  80 percent by mass), which contains
   electrically polarizable or dipolar molecules as
   well  as free ions.  The basic mechanism  of
   dissipation of incident  RF energy within  these
   tissues is that 1) the free rotations induced in the
   dipolar water  molecules  by the externally
                        3-8

-------
Figure 3-6. Illustration of object size vs. wavelength
          dependence. (Object is composed of tissue-like
          dispersive dielectric.)
     A  A
         \J\J
           Incident wavelength £> object size.

               Little power absorption,
               uniform internal distribution.
     A   A
        WT7
           Wavelength and object comparable in magnitude.

               Resonant case with significant absorption;
               internal distribution very nonuniform with
               "EM hot spots" present.

           Penetration
     A   A
         V/T7
            Incident wavelength <*object size.

               Intermediate absorption with internal
               energy deposition largely confined to
               object surface.

            Poor penetration
   impressed electric  field are  damped out  by
   collisions  with surrounding  molecules, and 2)
   conduction currents are induced within the tissue
   medium because of the presence of free ions.
   These phenomena are discussed extensively in
   Sec. 3.2.4, Mechanisms of  RF Interaction with
   Biological  Systems.

Over the past 40 years, many measurements have
been made of the dielectric properties of biological
tissues. These data  have  been collected in  the
Radiofrequency  Radiation Dosimetry Handbook
(Durney et al. 1978, Tables 9 and 10), as well as in a
more recent paper by Stuchly andStuchly (1980). The
dielectric properties of biological tissues depend to a
considerable  extent  on the water content of  the
tissue; i.e., tissues of high water content (blood, skin,
muscle, brain, etc.) exhibit much higher permittivity
(cf. Sec. 3.2.4.1) and conductivity values  than  do
tissues of low water content (fat, bone, etc.). As a
result, most incident RF energy tends to pass through
the fatty surface tissues of the body and be deposited
in the deeper tissues such  as muscle and brain.
Figure 3-5 illustrates some dielectric data for various
types of tissues in the RF range 0.01 to 10 GHz. Note
that the dielectric  properties  of  tissue  are not
constant,  but change with  frequency. This is an
example of what is termed a dispersive dielectric. For
the  case of  biological tissues,  the dispersive
characteristic above 1  GHz is created by the Debye
relaxation phenomenon in water molecules (discussed
further in Sec. 3.2.4, Mechanisms of RF Interaction
with Biological Systems). In part (b) of Figure 3-5, it is
apparent that conductivity increases  by almost an
order of magnitude for all tissues in the frequency
range from 1 to 10 GHz. Because of this fundamental
dielectric property of tissue, microwaves  in the
frequency range above about 5 GHz  become strongly
attenuated in the increasingly lossy tissue  medium.
Consequently, microwave energy of such frequencies
cannot penetrate deeply  into body tissues and  is
deposited near the body surface. As a result of their
poor  penetration capability, microwaves in the
centimetric wavelength region (<6 cm approximately)
are less capable of inflicting potential damage on
internal tissue than are those of longer wavelength (>
10cm).
b. Object size  relative  to  wavelength of incident
   field—The second major factor on which  energy
   absorption depends is the size of the irradiated
   object relative to the wavelength of the incident
   radiation.  Figure 3-6 illustrates  that  when the
   wavelength is much  greater than the object size
   (see  case A at top of figure),  the absorptive
   coupling  is  inefficient and little energy is
   deposited in the object. For this case, called the
   "subresonant" condition, values of the absorption
   coefficient  lie  in the range 0 to 0.5.  Case  B
   illustrates the resonant case, which occurs when
   the  wavelength and object dimensions are
   comparable.  For this case,  the  absorptive
   efficiency is  markedly improved so that  signifi-
   cantly more RF energy is coupled into the object
   and creates much greater deposition. Absorption
   coefficient values  in the  range  1.5 to 4 are
   obtained under resonant conditions. Furthermore,
   the incident energy penetrates into the object and
   is  deposited internally in  a characteristically
   nonuniform  manner, with   localized regions of
   enhanced energy deposition ("EM hot spots") at
   or  near the  object's center. These effects are
   discussed further  in  later sections  of this
   document. In case C shown at the bottom of Figure
   3-6,  the wavelength is much shorter than the
   object size. This represents the quasi-optical case,
   where the absorptive efficiency is similar to that at
   optical wavelengths. The absorption  coefficient
                                                                         3-9

-------
Figure 3-7.  Calculated whole-body average SAR vs. frequency
          for three polarizations in a prolate spheroidal model
          of a human; incident power density = 1 mW/cm2.
 Case        Field Polarization

                       I  -

A] E//a    Vertical


B] H//a*  Horizontal
                                         Body
                                       Orientation
   10°_ „.-,,-
                  102           103

                    Frequency (MHz)
                                       10"
   approaches 0.5  under these conditions, which
   gives intermediate absorption  values  that  fall
   between those achieved with the conditions of
   cases A and B. Case C is also characterized by the
   inability of the incident energy to penetrate much
   beyond the  surface of the  object (note  the
   dielectric factors discussed earlier); consequently,
   the energy coupled into the object under these
   conditions is confined to the object surface.

 c. Geometry and orientation of subject—The third
   factor on which energy absorption depends is the
   shape of the object and its orientation with respect
   to the electric-field vector of the incident field.
   Most biological  subjects  used in studying  the
   biological effects of RF  radiation have a charac-
   teristic shape that is significantly elongated along
   one axis  (similarly to humans).  Although this
   observation may appear to be simplistic, it is of
   considerable significance in the discussion of the
   interaction of RF waves with such objects. When
   the electric-field vector of the incident radiation is
   oriented parallel to the major or long axis of the
   irradiated subject (e.g., a vertically polarized field
   that is incident on a  standing human subject),
   then the subject absorbs as much as 10 times
   more energy at resonance than if the electric-field
   vector is oriented parallel to either of the two
   minor axes.  Figure 3-7  shows absorption
   characteristics of a prolate spheroid model of an
   average human. The ordinate represents the
   whole-body-averaged specific absorption rate
  •(SAR; see Sec. 3.2.2, RF Dosimetry Definitions, for
   discussion of units), and the abscissa represents
   frequency.  From Figure 3-7, it  is evident that
   when the electric-field vector is oriented parallel
   to the long axis of the prolate spheriod (electric-
   field polarization case), the absorption curve has a
   sharp peak at the  resonant frequency of ~ 80
   MHz.  For  the  other orientations (H-  and  k-
   polarization cases), it  is apparent that the
   absorption  peak is  much broader and occurs at
   somewhat higher frequencies. Furthermore, the
   resonant absorption values for the H and k cases
   are considerably  lower than  that for the E
   polarization case.  These basic  differences  in
   absorption values hold true in both the subreson-
   ant and resonant regions, but not  in the above-.
   resonant (or quasi-optical) region, where absorp-
   tion  in  the E-  and  H-  polarization cases  is
   approximately the same.
d. Complexity of incident radiation—The fourth and
   final  factor on which RF-energy absorption
   depends involves the complexity of the incident
   radiation. Up to this point, most of the discussions
   of RF coupling in biological targets have assumed
   the simplest form of free-field exposure, which
   involves a unipath plane wave emanating from a
   distant source and  incident on a subject
   suspended in space. (See Sec. 3.1 for definition of
   plane  wave.)  Virtually all  the  RF-radiation
   protection guides and standards in use throughout
   the world  are based  on this  rather  idealized
   assumption. Furthermore, most of the biological
   experimentation involving free-field exposures in
   anechoic  chambers  attempts to simulate this
   idealized concept.  However,  real-world, actual
   exposure conditions are far more complex.
One complicating  circumstance  arises in  the near
field. It is readily apparent that the most intense target
exposure with the resulting greatest potential for
injury would occur close to the RF-radiation source.
However,  in  this near field, the plane-wave
assumptions  that  are true for  the  far field of the
source's radiating antenna no longer hold. As
discussed in Sec. 3.1,  the near field  is characterized
by complex EM field properties: the E and H fields are
no  longer  in  space quadrature (E  and  H vectors
separated by 90°), and  the  value of the E/H ratio
(termed the wave impedance) differs greatly from the
constant value 377 Q that characterizes the far field.
                        3-10

-------
In the near  field, the  power density concept  is
meaningless  in its usual sense. For an object placed
in the near field of  an  antenna the interaction  is
exceedingly complex  because every type of antenna
possesses near-field  characteristics that are unique
to that type of radiator.  Consequently,  every
phenomenon involving near-field exposures must be
individually analyzed  and characterized. Although it
is possible to predict the spatial field distribution  in
the near field of an antenna, this distribution cannot
be  used  as a  basis for predicting the absorptive
properties of an object placed in this field because the
object will substantially alter that field distribution, as
well  as alter the radiating charateristics of the
antenna itself. When an object is placed in the near
field,  it interferes with the basic purpose for which
most antennas are designed: the efficient transfer of
RF energy from source to far field. An object in the
near  field tends to load the antenna capacitively,
which creates an impedance mismatch if the antenna
was originally well matched to free space. Asa result
the antenna reflects some of the transmitter energy
output back to the source. In general, for an antenna
that radiates well in free space, the closer the object is
to the antenna, the greater   is  the mismatch
condition, and the poorer is the transfer of radiated
energy to the far field.  Consequently, it is reasonable to
conclude that whole-body-averaged SAR values in
the near field do not greatly exceed those existing at
the beginning of the far field, although there may be
regions within the object of high energy deposition.
The limited data available from EM models involving
near-field exposure (see Sec. 3.2.3) seem to confirm
this conclusion.

A secondary aspect of near-field exposures involves
the unwanted exposure to leakage fields emanating
from RF devices, such as microwave ovens, RF heat
sealers, and diathermy units, which are not intended
to radiate energy beyond their immediate surroundings.
Such  leakage  fields are complex, and exposure
invariably takes place in the near zone of the leakage
source. Nevertheless, the limited data available from
near-field exposures of EM models to leakage fields
seem  to show that whole-body-averaged SAR values
are lower than might  otherwise be expected because
of the leakage-source and  near-field interaction
effects.

Another aspect of real-world exposures is the so-
called multipath problem. People  usually stand on
the ground and are not suspended in space as in the
idealized exposure  concept; furthermore,  they
frequently  stand close  to buildings or to other
reflective  surfaces. This  situation creates the
multipath problem, in which a subject is exposed to
scattered as well as  to direct energy. Some of the
effects are illustrated  in Figure 3-8, which shows that
there is  a  considerable enhancement in energy
absorption occurring for various multipath conditions.
These effects have been illustrated by means of the
prolate-spheroid representation of a 70-kg man. To
compute  the  SAR of man  for the nonresonant
condition, the  optical  cross-sectional area of the
prolate spheroid is first multiplied by  the power
density (10 mW/cm2), and then the resultant 38 W is
divided by 70 kg to give an SAR of 0.54 W/kg. Note
that when the model is in contact with a conducting
ground plane, there is a doubling  of the resonant
whole-body energy absorption compared to that of
free-space (Case II  vs. Case I). Furthermore, if the
model is  standing  at  a distance  in front of the
reflecting  plane that corresponds to an eighth of the
incident wavelength (Case V, 0.125,*),  there  is a
tenfold increase in absorption relative to that of free
space. Even greater increases are recorded when the
model is placed at a critical point inside a 90° corner
reflector (Cases IV and VI). The enhanced absorption
is a result  of the reflector producing an enhanced field
strength at the position of the object and not due to
interaction between object and reflector.

In addition to these multipath effects, enhancement
of energy  absorption can occur because  of proximity
effects that are created when two or more subjects are
simultaneously  irradiated. Gandhi et at. (1979) have
found that when subjects are spaced by a critical
separation of 0.65,*, a  50-percent enhancement in
absorption can  occur. Coupling of RF energy  into
biological systems is discussed  further in the
Radiofrequency Radiation Dosimetry  Handbook
(Durney et al. 1978, pp. 27-40).

3.2.2 RF  Dosimetry Definitions
Dosimetry is the measurement or estimation of RF
energy or  power deposition in an irradiated subject,
including  the internal distribution of that deposited
energy. In recent years, the terminology describing
dosimetric assessment has been evolving. Prior tothe
mid-seventies most investigators used terminology
that appears to have  been conveniently borrowed
from the ionizing-radiation field and that employed
the same  cumulative dosage concept (Youmans and
Ho  1975;  Justesen 1975).  Many investigators felt
that this terminology was inappropriate for use with
nonionizing or RF radiation because energy absorption
in  this case   is not  considered  a cumulative
phenomenon in a way that  has been accepted for
ionizing radiation  (Susskind  1975; Guy 1975).
Consequently, a new terminology was proposed that
did not use the word "dose." The now widely
accepted term "specific absorption rate" (SAR) was
first reported in the literature by Johnson (1975);
however,  the term evolved from joint discussion by
members  of Scientific Committee 39 of the  National
Council on Radiation Protection and Measurement
(NCRP). (This meeting, chaired by George Wilkening,
took place at the Battelle Research Center in Seattle,
Washington, on June 23,1975, with members Frank
Barnes,  Curtis Johnson, Arthur  Guy,  Charles
                                                                      3-11

-------
Figure 3-8.  Absorption dependence on various ground and multipath factors (Gandhi etal. 1977). Note that these data are based on
            experimental measurements.
                                    Rate of Whole-Body Absorption (WBA) in Watts for a
                       Prolate Spheroid Model of Man (weight = 70 kg, height = 1.75 m) at 10 mW/cm3
               WBA
                                          SAR
Nonresonant conditions

              38 W
          (no reflectance)
              19W
         (50% reflectance)


Case I.    At resonance for free space.
                                        0.54  W/kg
                                       0.27  W/kg
                           0
              151 W
                                       2.1 6 W/kg
                                     f = 62-68 MHz
Case II.   At resonance for conditions of electrical
         contact with the ground plane.

         2 x 151= 302 W
                                        4.31 W/kg
                                       f = 31-34 MHz
Case III.  At resonance for placement in front
         of a flat reflector.
                             O
                         d = 0.125

         4.7 x 151= 710 W
                                         10.14 W/kg
                                        f = 62-68 MHz
                                                                       WBA
                                                                                                     SAR
                                                              Case IV.   At resonances for placement in a 90° corner reflector.
                                                                                    d= 1.5A
                                                                27x151 =4077W
                                                                                                 58.24W/kg
                                                                                                f = 62-68 MHz
Case V.   At resonance in electrical contact with
         ground plane, in front of a flat reflector.
                                                                                   d =0.125 A
                                                                2 x 710= 1420 W
                                                                                                 20.28 W/kg
                                                                                                f = 31-34 MHz
                                                             Case VI.  At resonance in electrical contact with
                                                                      ground plane in a 90° corner reflector.
                                                                                      d = 1.5
                                                                 2 x 4077 = 81 54 W
                                                                                                 11 6.48 W/kg
                                                                                                f = 31 -34 MHz
                           3-12

-------
Susskind, Saul Rosenthal, Karl Illinger, and Ronald
Bowman attending. The committee members agreed
to publicize the new term as much as possible in their
publications to promote wide acceptance. Members
Johnson, Susskind, and Guy  published comments
concerning SAR in 1975.) NCRP (1981) has formally
recommended adoption of the term  SAR  in  RF
dosimetry. NCRP has also endorsed  use of the term
"specific absorption" (SA) for  the absorbed energy
per unit mass (referred to earlier as  "dose").
Table 3-2 lists the various RF dosimetric quantities in
general use, together with their definitions and units.
Although the terms (particularly SA and SAR)  for
some of these quantities have now gained widespread
acceptance, other terms, particularly  dealing with the
volume-normalized power or energy absorption, have
only recently been proposed and have not yet gained
wide acceptance.

The mass-normalized and  volume-normalized
absorption rates are directly  related through the
localized or whole-body-averaged tissue mass
density p, that  is, absorption-rate  density, ARD =
p(SAR), where p is in kilograms per cubic meter. Most
biological tissues are composed largely of water, so
that it is reasonable to assume a unity tissue density
value. In this case, the two quantities are numerically
equivalent. This approach has been adopted by many
investigators engaged in predictive EM modeling (see
Sec. 3.2.3, Analytical  and  Numerical  RF-Electro-
magnetic Interaction Models), where the model is
composed of some homogeneous tissue with average
physical  and dielectric characteristics that are
representative of the various tissues of the body being
modeled. For example, all the SAR data given in the
familiar Radiofrequency Radiation Dosimetry Hand-
book were derived this way; the assumption of a unity
average tissue density is specifically stated (Durney
et al. 1978).

A subset of two additional values exists for both the
mass-normalized and volume-normalized quantities
listed  in Table  3-2: (a) the whole-body-averaged
value,  which represents the overall absorption or
absorption rate divided by the total mass or volume of
the  subject;  and (b) the  localized value,  which
describes the absorption or absorption  rate in  an
incrementally small mass or volume at some given
point within the subject.  The localized ARD is
mathematically defined by the expression 
-------
Table 3-2.    Proposed System of RF Dosimetric Quantities. Definitions, and Units*
                                                          SI Unit
Quantity and Definitions
Name
Symbol
Other Units Used
Whole-Body Absorption
(formerly integral dose)
  The total EM energy absorbed
  by the irradiated subject
Whole-Body Absorption Rate
(formerly integral dose rate)
  The time rate of total EM energy or
  total power absorption by the
  irradiated subject
Specific Absorption
(formerly dose)
  The mass-normalized EM energy absorbed
  by the irradiated subject
Specific Absorption Rate
(formerly dose rate)
  The mass-normalized rate of energy
  or mass-normalized power absorbed by
  the irradiated subject
Absorption Density
(also absorbed energy density
or energy density, dissipated)
  The volume-normalized EM energy
  absorbed by the irradiated subject
Absorption Rate Density
(also absorbed power density,
density of absorbed power, heating
potential, etc.)
  The volume-normalized rate of EM energy
  or volume-normalized power absorbed by
  the irradiated subject
joule
watts
joules
per
kilogram

wans
per
kilogram
joules
per
cubic
meter


watts
per
cubic
meter
                      W
J/kg
W/kg
J/m3
W/m3
                     millijoules (mJ)
                     (1 mJ = 10'3J)
                     calories
                     (1 cal = 4.187 J)
                     Milliwatts (mW)
                     (1 mW = 10~3W)
                     cal/min
mj/g

cal/g

mW/g

cal/g-min



mJ/cm3
mW/cm3
(1 mW/cm3 =
103 W/m3)
"Throughout this table, "absorption" refers to RF-energy absorption.
corresponds to a whole-body-averaged SAP of ~ 1 to
1.4 W/kg. To determine the power density level of
incident radiation that will create this exogenous heat
load in man,  a  detailed knowledge of human  RF-
absorption characteristics is required. At the time the
first RF-radiation protection guide was proposed in
the early 1950s, virtually no data existed on the RF-
absorption characteristics of humans; the assumption
had to be made that the EM absorption cross section
for humans was equivalent to man's geometric cross
section, which is ~  1  m2 (i.e., a unity absorption
coefficient was assumed). From this calculation came
the  well-known exposure limit  of 10 mW/cm2.
During the past  10 years or so,  a considerable
accumulation of  data based  largely on human
interaction modeling has been developed (discussed
later in this section). These data showed that under
resonant conditions, the RF-absorption cross section
of humans can be 4 to 8 times greater  than the
geometric cross section. Thus, at resonance, the RF-
absorption rate for humans can potentially exceed the
BMR by this  factor when exposure is to incident
radiation  levels  of 10  mW/cm2.  Such a  situation
clearly constitutes an excessive thermal burden for
exposures of indefinite duration. Thus, EM  modeling
has played a significant role in demonstrating that,
under  certain conditions, the originally recommended
     exposure limit is potentially unsafe, and in providing
     an objective basis for the revision of that guideline.
      The  second reason for the importance  of EM
      modeling is that it provides  us with a method of
      extrapolating absorption data from animal  experi-
      ments for application to humans. From the  prolate
      spheroid representation for humans and animals, a
      considerable amount of data has been compiled on
      whole-body-averaged SAR vs. frequency characteris-
      tics for several species. These data are given in the
      Radiofrequency Radiation Dosimetry Handbook
      (Durney et al.  1978); highlights of this material are
      presented in Table 3-3, which shows the resonant
      frequency range (long axis oriented parallel to E field)
      and  the  whole-body-averaged SAR  at resonance,
      normalized to an incident field level of 1  mW/cm2 for
      five different species,  including humans. The data
      given in Table 3-3 enable a  researcher readily to
      determine  for a particular  animal species the
      resonance   frequency  range  to  be  used  during
      experimentation.  For  example, to simulate  human
      resonance in adult rats, exposure would have to be
      in the frequency range of 600 to 700 MHz. This is an
      example of  "frequency scaling," or adjusting the
      frequency  at  which an  experimental animal is
                        3-14

-------
 exposed  in  order to approximate  certain  given
 conditions of human exposure. In addition, differences
 in the whole-body-averaged SAR between  the two
 species must be considered; from the last column of
 Table 3-3, it is evident that, at resonance, the whole-
 body SAR for the rat is 3 times that of a human. Thus,
 if a certain  SAR  value is obtained in a rat  when
 exposed at a 10-mW/cm2  level, roughly 3 times that
 level, or 30 mW/cm2, is needed  to obtain the same
 SAR value in humans (resonant condition for both).

 The data shown in Table 3-3 can  be likewise used to
 determine whether a given exposure frequency falls
 in, above, or below the resonance range for humans.
 As was discussed earlier, these three categories
 lead to fundamental differences in absorption as well
 as patterns of internal energy deposition. Most likely
 a corresponding difference in biological effects will be
 seen in several frequency  ranges, with the resonant
 range  representing the potentially  most hazardous
 condition  because of the high absorption efficiency
 and highly nonuniform internal deposition of energy.
 As already emphasized, proper simulation of a given
 human exposure scenario  by means of experimental
 animals requires the use of (1) frequency scaling to
 ensure that the type of interaction taking place is
 roughly comparable (i.e., subresonant, resonant, or
 supraresonant); and (2) appropriate adjustment of the
 incident exposure  level to  obtain  a specified whole-
 body-averaged SAR  or, where  possible, a given
 localized  SAR  in some specified target  organ. To
 check for  the possibility of any effects that might be
 specific to a  particular frequency to which humans
 are  being exposed,  animals must  necessarily be
 exposed  to  that same frequency.  Given such an
 experiment, the use of frequency scaling is obviously
 inappropriate.
3.2.3.1 Model Details
Gandhi (1980) and Durney (1980) published excellent
reviews on EM modeling, summarizing most of the
major research contributions to 1980 and detailing
the status of the EM modeling field at that time. The
material that follows does  not duplicate these
reviews, but instead traces the historical development
of EM models with emphasis on  the  increasing
complexity of such models and their application. Only
a few relevant publications in the field are cited; the
reader is referred to Gandhi's and Durney's reviews
for more complete details,  as well  as  to the
Radiofrequency Radiation Dosimetry Handbook
(Durney et at. 1980, pp. 42-44), which  contains a
comprehensive literature survey of all theoretical and
experimental modeling work performed to 1 979. The
reader is also referred to the original publications for
further details of the mathematical techniques used
to solve  the  various problems that are discussed
below.
 Table 3-3.    Range of Resonant Frequencies of Man and
            Animals Irradiated by Plane Waves in Free Space
            at 1  mW/cm2 with Long Axis Parallel to the
            Electric Field*
Species
Man (av)
Woman (av)
Child (10 years
old}
Rhesus monkey
(seated)
Dog (beagle)
Guinea pig
Rat (medium)
Mouse
Average
Weight
(kg)
70
61

322

35
135
0.58
032
002
Average
Length
(m)
1 75
1.61

1 38

004
006
0.02
0.02
0.007
Resonant
Frequency
(MHz)
70-80
80-90

90-100

300-350
200-250
550-600
600-700
2400-2600
RMR
(W/kgl
1 26
1.15

2.00

2.36
1.72
3.83
4.8
6.5
Average
SAR
(W/kg)
0.25
025

035

030
016
055
075
072
 •Data derived from Durney et al (1978).

 Planar  and spherical models—The earliest model
 considered involved the one-dimensional solution to
 the  planar tissue slab irradiated by a plane wave
 (Schwan and Li 1956). The slab is infinitely wide and
 can contain  several  layers of various tissues.
 Although the  solution to this problem is straight-
 forward, the model  is restricted in  its application
 because it does not account for the obvious closed-
 form shape of all human and infrahuman bodies.
 This particular  model  is  valid  only when the
 wavelength  is small compared to the system size, as
 depicted earlier in Figure 3-6, case C; for human
 irradiation,  such wavelengths correspond to fre-
 quencies above approximately 3 GHz. The concept of
 "penetration depth" has been developed on the basis
 of this one-dimensional model. The penetration depth
 is the depth within the tissue slab at which the RF
 power has been attenuated to a value of 1 /e2 and is
 shown  as a function of frequency in Figure  3-9.
 Strictly  speaking, this concept applies only to this
 particular planar model; however, if prudently applied
 the concept  has an approximate qualitative meaning
 for  other  configurations.  It is  unfortunate that the
 concept has been extensively borrowed and is often
 inappropriately used in discussions of RF absorption
 in nonplanar models where the wavelength is either
 longer than  or comparable to the object  size. The
 penetration depth concept can be accurately applied
 only when the wavelength is small compared to the
 object size.

The next level of complexity involves two-dimensional
solutions of the infinitely long cylinder with circular,
elliptical, and arbitrary cross sections; some of these
models  are of  homogeneous tissue and others are
multilayered (Ho 1975). These models are primarily
intended to  simulate  RF heating in  human limbs
under conditions  of plane-wave irradiation from  a
rectangular applicator in contact with the cylindrical
model, which is of particular interest  to diathermy
practitioners.  Ruppin  (1979) examined  a lossy-
dielectric cylinder model placed in front of a perfectly
conducting plane (i.e., a multipath configuration). His
data show  both  enhancement  and reduction of
                                                                      3-15

-------
average SAR relative to the isolated cylinder,
depending on the conditions of irradiation. Cylindrical
models were  also used extensively for SAR
calculations  in the second edition of the Radiofre-
quency Radiation Dosimetry Handbook (Durney et al.
1978).
The first attempts at three-dimensional solutions
involved homogeneous spheres of tissue-equivalent
dielectrics. Although it is only a poor approximation to
the actual shape and  heterogeneous  dielectric
composition of humans and experimental animals,
the sphere model has been used extensively because
of the  ready availability  of  solutions based on the
classical Mei theory involving spherical wave
functions (Stratton 1941).  These methods can be
readily programmed  on  high-speed machines that
yield data on the EM absorption cross section and on
the distribution of localized ARD for homogeneous
spheres (Kritikos  and Schwan  1975). (Some
approximations made by Kritikos and Schwan [1975]
can be used in the subresonance or Rayleigh region
and the supra-resonance  or  quasi-optical region,
which  greatly simplifies  these computations.) This
work  demonstrated that  a  highly  nonuniform
absorption distribution is obtained under resonant
conditions, including the  formation of  relatively
intense "EM  hot  spots" near the sphere's center
(Figure 3-10). Other workers (Weil 1975) have used a
multilayered spherical model that consists of several
concentric layers of different tissues representing an
idealized model of an adult, child, or monkey head
exposed to plane-wave radiation. However, this
concept has obvious  limitations because the
spherical model is spatially  isolated  rather than
attached by the neck to the body's trunk. Results for
the multilayered sphere  were  similar to those
obtained for  the homogeneous  sphere with the
exception  that,  in the  supraresonance region,
enhanced absorption was noted in the multilayered
model. This enhancement is caused by the presence
of the surrounding layers of skin  and fat, which
appear to  provide an impedance  transformation
mechanism by which energy transfer into the sphere
is increased. A recent paper by Barber et al. (1979)
has shown that a planar model  can accurately predict
this enhancement effect for any three-dimensional
nonplanar shape, and that the layering enhancement
factor can be applied to any nonlayered object, such
as a  human model, to predict  trve absorption
characteristics of the layered object.

The effects  of altering irradiation from plane wave to
near  field  on the  absorption characteristics of
multilayered sphere  models were  investigated by
Hizal and Baykal (1978). They found that EM hot spots
continue to occur when the model is excited near
resonance in the near field of a loop antenna, but not
when similarly excited by a dipole antenna.
Figure 3-9.  Penetration depth as a function of frequency (data
          from Table 3, Oumey et al. 1978).
  100
  0.1
    10
100           1000

   Frequency (MHz)
10000
Prolate spheroidal and ellipsoidal models—Since
prolate  spheroids and  ellipsoids are much better
approximations to actual human and animal shapes,
use of these models was a logical next step to solving
this problem. Although one can in principle obtain
solutions  to  the prolate spheroid and ellipsoid
problem in spheroidal and ellipsoidal wave functions in
the manner already accomplished for the sphere, in
practice this approach has not been possible because
of many mathematical difficulties. Consequently,
workers have resorted to using  several different
approximation techniques, further details of which
may be found in Durney's review (1980). Such
methods have provided accurate data on  absorption
in the subresonant and  near-resonant  region,  as
well as in the supraresonant or quasi-optical region.
These data have been compiled in tbeRad/ofrequency
Radiation Dosimetry Handbook (Durney et al. 1978,
pp. 75-106)  in  whole-body-averaged SAR vs.
frequency  plots for many  animal species. However,
under some conditions, there is a gap at the resonant
and immediate post-resonant  regions, where these
approximations no longer hold. Using approximate
data derived from antenna theory and experimental
observations, investigators have bridged this gap with
the aid of curve-fitting techniques. This approach is
illustrated in Figure 3-11, which shows SAR data for
a prolate spheroidal model of man for the three basic
orientations of the model's long axis relative to the E,
                      3-16

-------
Figure 3-10.
6.0 cm
1650 MHz
E-Plane
ARD distribution in core of 6-cm radius
multilayered sphere at 1650 MHz (Weil 1975).
            IS	170
  -70,
H, and k vectors. Note the strong orientation effect
discussed earlier. Durney et al. (1978,  1979) have
also developed a useful empirical formulation that
describes the E-field polarization curve shown in
Figure 3-11.  It can be  programmed  on a hand
calculator and yields reasonably accurate SAR data
for a prolate  spheroidal  model of any size of
eccentricity.

As discussed earlier, an isolated model, regardless of
shape, suspended  in  space  is an inappropriate
representation of the real-life conditions under which
humans are exposed. A prolate spheroidal model in
contact with or in  close proximity to a conductive
ground plane  is an obvious improvement over the
isolated model. Iskander et al. (1979) have analyzed
such a model by both antenna  and circuit theory. In
this model,  the finite separation between feet  and
ground created by the usual presence of footwear can
be accounted for by a lossy capacitor in the equivalent
electrical circuit  of the prolate spheroid.  From
Iskander's data (Figure 3-12), it is apparent that the
primary effect of the ground plane is to shift the SAR
curve for the isolated model to the left  with  only a
slight evident  increase  in peak SAR. The resonant
frequency of the prolate spheroidal model of man is
shifted from ~ 75 MHz to 45 MHz when the model is
in perfect contact with a conducting ground plane; the
intermediate case shown, in which resonance occurs
at 55 MHz, represents imperfect contact between
model and ground plane due to a 3-cm lossy gap. The
latter case is probably the most realistic  of the three
considered. Iskander er al. also concluded that when
the model is separated from the ground plane by a gap
of 7.5 cm or more, the whole-body SAR values are
essentially the same as those for the isolated model
It is evident from oral presentations at recent (1978-
1980) scientific meetings that much effort is being
devoted  to analyses of  the absorption of  prolate
spheroid and other  models in the near field of  an
antenna. Iskander et al. (1980) have studied such a
problem for a prolate spheroid model of man in the
near field of a short electric dipole at 27 MHz. They
found that average SAR values in the near field tend
to oscillate about  the  constant value predicted for
plane-wave  irradiation. There are  also significant
alterations in the  internal distribution of deposited
energy in the near field.

As  discussed previously, the data developed on
prolate spheroidal and ellipsoidal models have played
an important role in extrapolating RF absorption data
from lower animals to man. Yet these data have
limitations:  the solutions are  not exact  but are
approximations, and yield data  that are  accurate
enough (±10 percent)for most practical applications.
Furthermore, with the  exception of the subresonant
region  (long-wavelength analysis. Figure 3-11),
existing solutions  provide data on only the whole-
body absorption characteristics of this model. No data
are  yet  available  on  the  three-dimensional ARD
distribution and EM "hot spot" formation potential in
the  resonant and  immediate supraresonant region.
Such data  would be  of potential  value  to  those
engaged in experimentation with rodents because of
the close similarity between the rodent shape and the
prolate spheroid.

Human block models—Although  prolate spheroidal
and  ellipsoidal  models are reasonably accurate
representations of  many laboratory animals, partic-
ularly rodents, that is not  true  of the primates,
including man, because the appendages(head, arms,
and legs) are much larger and more significant. The
importance of these  appendages is underscored by
experimentation on  saline-filled  dolls and  scaled-
down human phantoms  (Guy et al. 1977), which
showed relatively intense localized absorption in the
ankles,  legs, and neck. Since it  was impossible  to
confirm  the localized effects  with the  prolate
spheroidal model, attempts were made to develop a
more realistic  human  model  in  which numerous
cubical cells form  a so-called block model of man
(Figure 3-13). The  EM  solution to this  type  of
sophisticated problem is realized through the solution
of a large  system of simultaneous equations and
matrix inversion or iterative techniques.  Further
mathematical details are given in Durney's review
(1980). Such solutions can be obtained only by use of
dedicated  large-scale computing  facilities with
expensive memory capability. Furthermore, because
of finite limits in the average memory capacity of all
but the largest computers, the number of cubical cells
                                                                      3-/7

-------
Figure 3-11.  Curve fitting of SA R data for a prolate spheroidal
           model of man for the three basic orientations of
           the model's long axis relative to the E, H, and k
           vectors of an incident plane wave (1 mW/cm2
           power density). Taken from Ourney et al. (1978).
   10°H
                                                   Figure 3-12.  Effect of a capacitive gap on average SAR
                                                              between the man model and the ground plane
                                                              (Iskander et al. 1979).
   10-'
cc

W  ID'2

S,
2
0)
  10-3
                                            H//r  =
                                            E//a
                                            k//a
                        	Long wavelength analysis
                        -«- EBCM

                        	Empirical relation
                        	Estimated values
                         • • Cylindrical model
                        -*—*• Geometrical optics
                            technique
     10'
               102        103        104

                     Frequency (MHz)
                                             106
that can be used to make up the model is limited. The
accuracy and resolution of the solutions depend on
cell size. In the work of Chen and  Guru (1977), a
maximum of 108 cells was used, whereas Hagmann
et al. (1979a) used 180 cells. These  limits inevitably
reduce the spatial resolution with which the localized
absorption  distribution can  be determined  in the
model, and they place an upper limit on the frequency
range over which solutions are valid. (This range does
include the resonant and immediate supraresonant
region.)  Since this  technique readily  allows for
differences in the dielectric properties of individual
cells in the models, it is possible to approximate the
inhomogeneous tissue properties of the human body,
but there isa limit to howfinely such inhomogeneities
can be modeled.

Figure 3-14 shows some data obtained for the human
block model (Figure 3-13) when it stands on a ground
plane and is irradiated by a vertically polarized plane
wave. The  curve labeled "whole body" gives the
whole-body-averaged SAR  throughout the model,
IV






"5
>
2
c
0)
1 io-i

£
«


1
E
_o
5"
5
of 10-2

C/5

0>
ID
1
<

10-?.
1











M
{Vj
^*SJ
X^/'
_p X
<£'' 4>'
«>c/w// « -^

J" r- '
rf* IT / /
. ^ /. /
' C^/^ ^
*>'/&/
/
-------
 human and infrahuman heads (Rukspollmuang and
 Chen 1979). The Hagmann er a/, data show strong
 localized absorption at the front and back of the neck
 and in the center-based region of the skull; localized
 SAR values are  two and four times the whole-body
 average.
 RF absorption in block models when irradiated under
 near-field  conditions  is being  studied  by some
 workers. Chatterjee et at. (1980) reported a study
 involving the near-field interaction of the  human
 block  model (Figure  3-13) with the measured
 leakage-field distribution of an industrial RF heat
 sealer operating at 27 MHz. Their data show that
 averaged and localized SAR values are much lower
 for near-field exposure than  is the case for plane-
 wave exposure at equivalent power density levels.

 3.2.3.2 Accuracy
 As discussed already, this document and  the 1982
 ANSI guideline (ANSI 1982) recommend frequency-
 dependent  exposure limits that have been derived
 mostly from the EM modeling data just described. The
 validity of this approach therefore depends strongly
 on the accuracy of these data. One may conclude that
 the accuracy of the whole-body-averaged absorption
 data is good (certainly within ±10 percent). The good
 agreement  obtained between data for the  human
 prolate spheroidal and block models appears to
 support this contention. However, the accuracy of the
 localized absorption data is probably inferior. This is
 particularly true of the block models where accuracy
 is probably no better than ± 50 percent, because of
Figure 3-13.  A realistic block model of man (Hagmann era/.
           1979a).
 the limitations on numbers of blocks available. On the
 other hand, localized absorption data for the spherical
 model,  as well  as those available for the prolate
 spheroid, are more accurate, since they were derived
 by more rigorous techniques.
 3.2.3.3 Unresolved Issues
 The need for additional modeling has been discussed
 by Durney (1980). Improvements and refinements in
 existing block models are needed, especially in spatial
 resolution and accuracy; the same is also true  in
 modeling tissue  inhomogeneities.  In coming years, a
 new generation  of "superpower" computers with
 greatly  enhanced computational speeds—an order-
 of-magnitude increase over existing machines—will
 undoubtedly become available for scientific studies.
 Such computers will almost certainly be used  to
 refine existing EM absorption models, particularly in
 spatial  resolution problems  and in solving  more
 complex  problems that hitherto have  not  been
 attempted because of computer limitations.
 As Durney (1980) emphasized, there is still a need
 for more solutions to near-field interaction problems,
 since the near field more frequently represents the
 real-life situation to which people are exposed.
 Though  models  involving near-field irradiation are
 considerably more complex than are those for the far
 field (plane wave), it should be possible to attempt the
 solutions of these based on our existing knowledge of
 plane-wave solutions.

3.2.4 Mechanisms of RF Interaction  with
Biological Systems
Biophysical models that relate molecular structure to
the interaction of RF radiation with biological systems
provide a  base  for  understanding the  potential
hazards of this type of  RF radiation and for designing
relevant experiments. The purpose of this section is to
review these interactions and, where possible, to
discuss the changes in  structure  and  biological
function that may result.

3.2.4.1 Complex Relative Permittivity
The molecular-level interactions may  be  averaged
over a macroscopic biological sample and character-
ized by an average interaction parameter called the
complex relative  permittivity (£*) of the sample. This
interaction parameter facilitates discussion of the
effects of a biological sample on thermodynamic
properties. It  is  a measure of the capacity of the
charge distribution within the sample to adjust to a
change  in an applied electric field. This measure  is
composed of two parts, the real component of the
complex permittivity ft') and the imaginary component
of the complex permittivity (f" =  a/&>£0). The real
component is a measure of the energy stored in the
charge distribution of the sample by the interaction
with the field. The imaginary component is related to
                                                                      3-19

-------
Figure 3-14.   Absorption for man block model standing on ground plane (Gandhi et al. 1979).
                                                                  Incident Power Density = 1 mW/crn*
                                                    102

                                                Frequency, MHz
the energy dissipated by the field in the sample.
Initially, the dissipated energy may be restricted to
just  a few  molecular modes within the sample.
Eventually, the dissipated energy will be thermalized,
but temperature gradients  may persist within the
sample if the rate of energy dissipation is a function of
position, due to the geometry of the sample or to field
gradients  or differences  in c" within the system.
Finally, after the RF-radiation  exposure has ceased,
the dissipated  energy will be equally distributed
throughout the various modes  of the sample and will
result in either a change in sample temperature or a
transfer of thermal energy from the sample as well as
possible chemical and structural changes.

The r* of many biological tissues has been measured
by several investigators. Some excellent reviews are
available (Schwan  1957;  Johnson and Guy  1972;
Durney et al. 1978, 1980;  Stuchly and Stuchly 1980;
and  Schwan and Foster 1980).  Complex relative
permittivity is a function of frequency over our region
of interest.

Table 3-4 shows the values of e' and  c" (abstracted
from Durney et al.  197^.,  Schwan and Foster 1980;
Stuchly and Stuchly 1980). The dielectric properties
of muscle tissue have been  used to characterize the
dielectric properties of all biological  tissues with  a
high water  content (Schwan  1957; Schwan  and
Foster 1980). This characterization also serves as an
illustration  of some of the underlying phenomena
responsible for f *. Figure 3-15 showsthec' of muscle
as a function of frequency.
It can be seen in Table 3-4 that in the region between
0.5 MHz (the lower end of the range of interest) and
~  50  MHz, the  complex permittivity of muscle
decreases by two orders  of magnitude. However,
between 100 MHz and 10 GHz it decreases by < 50
percent; and again, at  frequencies above ~ 10 GHz,
but still  within  the  range  of interest,  the  real
component of complex  permittivity decreases rapidly.

The imaginary component of the complex permittivity
(O has two components—one due to the movement
of free charges, and the other to the orientation of
either permanent dipoles or dipoles resulting from
the interaction of the field with polarizable microscopic
structures. The former component is proportional to
the inverse  of the applied RF-radiation frequency;
energy absorbed this  way acts directly to heat the
sample. The modes  of absorption  of the latter
component relate  to the specific structure of
biological molecules and molecular complexes,  and
this component  has  a  maximum where the  real
                       3-20

-------
component of the complex permittivity £' is changing
most rapidly with respect to frequency.
Frequency regions with rapidly changing dielectric
parameters indicate that a  particular component of
the system, or  a particular type of response to the
field, is reaching the limit of its capacity to respond
(i.e., at  higher frequencies the field is changing too
rapidly for a particular component to respond). It has
been suggested that the rapid change in the region
below  50 MHz results from polarization effects in
which   cellular  membranes are  charged  by the
surrounding electrolytes (Schwan and Foster 1980).
Another  response  that  is limited to this  low-
frequency range is the rotation of large  dipolar
macromolecules or molecular complexes orienting
with the  changing  field (Takash'ima  and Minikata
1975).  The dielectric response below 50 MHz may
result from more than one mechanism. The response
above ~ 5 GHz results from the rotation of water
molecules in the biological systems, and it is similar
to the  response  seen  in pure water. It has been
postulated that a small decrease in c' observed near 1
GHz results from  rotation of  water molecules
constrained by their interaction with large molecules
and molecular complexes (Schwan 1965; Grant et al.
1968) and, therefore, these water molecules cannot
respond to the field at higher frequencies.
The general features of c* of biological tissues can be
explained in terms  of  the major molecular and
structural components of the tissues. However,
biological systems are inhomogeneousand inherently
complex at the molecular level, and interaction with a
minor  molecular component may have negligible
influence  on the dielectric parameters of tissue (an
average property of a macroscopic volume) but may
interfere with the capacity of that minor component to
perform its biological function. Therefore, it is useful
to discuss the specific mechanisms for the absorption
of RF radiation  at the molecular level. Furthermore,
since the mechanisms  are similar  for the major

Table 3-4.   Dielectric Permittivities for Various Tissues*
components, this discussion also provides a more
detailed explanation of the interactions responsible
for the complex relative permittivity e*.

3.2.4.2 Mechanisms for RF-Radiation
Absorption at the Molecular Level
Table 3-5 presents information on the energy of an RF
beam relevant to molecular level interactions. The
binding energy of a single electron in a molecule is
typically of the order of tens of electron volts, whereas
excitations of electrons within molecules are of the
order of electron volts, and the energy of a hydrogen
bond is > 0.1 eV. Accordingly, direct linear mecha-
nisms for the absorption of a photon or a few photons
by biological molecules from an RF  field  cannot
involve changes in electronic or covalent molecular
structure,  nor  can  they involve the breaking of
hydrogen  bonds. That does  not  imply that single
photons cannot cause changes in the structure of
biological  molecules or molecular  complexes. They

Figure 3-16.    The real component of the complex permittivity
             of muscle (;') a* a function of frequency (f)
             from the values for muscle (Schwan 1957).
                  4567
                    Log f (CPS)
8   9  10
Frequency
(MHz)
Tissue
Type
Muscle

Brain

Liver

Bone

Fat

Dielectric
Value
£'
f"
c'
f"
f'
t"
£'
c"
c'
f"

0.1
3 x 104
106


10"
5 x 10"
8.0
0.5



1
2 x 103
10"

2.6 x 103
1.5 x 103
5x 103





10
220
1.2 x 103
240
500
240
800





50
90
400
110
180
90
195
7.2
10.8
12
17

100
75
80
80
95
78
110


10
12

103
54
27
40
16.5
46
18
5.8
1.3
6.0
2.0

104 1.7X104
40 34
21 21
35
15
36
11
4.9
0.7
4.0
0.7
'These values are abstracted from reviews by Schwan and Foster (1980) and Stuchly and Stuchly (1980). The original sources are quoted
 therein. The values have been chosen in the middle of the range for experimental values. Reported values may differ from one another by
 as much as a factor of 2.
                                                                       3-27

-------
contain sufficient energy to change  the three-
dimensional structure of biological molecules in a
manner that leaves the covalent and  hydrogen-
bonded structure  intact. The  following  paragraphs
describe some of the  excitations of biological
molecules that can be induced by an RF field and the
changes in three-dimensional molecular structure
that may result.
A  change  in the rotational energy  of  dipolar
molecules or molecular segments is a ubiquitous
mode for energy absorption  from an RF  field by
molecules  in a  biological system. The absorbing
molecular structure increases its angular momentum
by quantized units. The size of these units  and the
characteristic frequencies for absorption depend on
the distribution of mass within that structure. In
general, the more massive the rotating structure, the
lower the characteristic frequencies.

The rotational mode of interaction is relevant through
the entire range of the frequencies of interest for the
purpose of  this document. Free rotation of  proteins
and other bipolymers are modes for absorption in the
frequency range from 10 kHz to 100 MHz and beyond.
In addition,  smaller molecules and  molecular
segments absorb at higher frequencies by this mode.
The absorption of energy into free rotational modes
does not lead to structural changes in the absorbing
molecule or  segment  because it rotates rigidly.
Through collisions with other molecules in the
biological system, this initial  increase in rotational
kinetic energy is dispersed throughout the various
molecular motions of the system (i.e., the energy is
thermalized). With this mode of absorption, functional
changes, if any, result only from these local increases
in temperature. In a biological system at normal body
temperatures, an individual molecular  species
absorbs over a broad frequency range by this mode
because of its concurrent interaction  with other
molecules.

Since water, which has a characteristic frequency for
free rotation near 20 GHz, is the major component of
biological systems, a change in the rotational energy
of water molecules is the principal mechanism for
absorption of microwave radiation at frequencies > 2
GHz. (The frequency range for which this interaction
is likely to be  important is broadened by the
interaction  of absorbing molecules  with surrounding
molecules.) At the  lowest frequencies  of  interest,
changes in molecular rotational motion may also
make  a significant contribution to the overall
dielectric properties through the interaction of
microwaves with large molecules.
For many biological molecules, segments as large as
many amino acids or as small as a methyl group have
some rotational freedom, but are constrained both by
covalent bond(s) to the main structure of the molecule
and by electrostatic interactions with other nearby
Table 3-5.   Energy Units for RF Radiation

A                 f                  Energy-photon
(m)                (MHz)              (eV)
10*
1.24x10'
3x10"
10
3 x 102
3x 104
2 45 x 103
103
3x 10
1
1.2x 10 *
1   x 10's
4.1 x10"6
1.2x10'7
4.1 xlO'9
molecules. This rotational freedom provides a mode
for the absorption of energy from an  RF field.
Although this mechanism is similar to the preceding
one, it has been suggested (Illinger 1970; Rabinowitz
1973) that it can result in changing the equilibrium
position of the rotating segment relative to the main
structure of the molecule and surrounding molecules.
This structural change is probably reversible but
could have consequences for biomolecular function.
The  importance of this mechanism for  a biological
effect of RF radiation has  not been substantiated
experimentally.
More complicated intramolecular motional modes
involve rotation and vibration, bond stretching, and
twisting. These modes also absorb energy from an RF
field and have potentially similar effects on molecular
structure  and  function.  In general, these more
complicated  modes are at frequencies above the
range of interest, but there is some suggestion that
they are important near  10 GHz. Prohofsky et al.
(1979) have calculated the stretching modes for an
artificial double-stranded DNA  polymer containing
one  strand of adenine that is hydrogen  bonded to a
strand containing only thymine. They suggest  that
absorption of energy into these modes may affect
DNA helix melting and replication. The characteristic
frequencies of these calculated modes are, however,
> 40 GHz.

The amount, of energy required to excite either
rotational  or the  more complex motional modes is
small  with respect to the  amount of  energy
exchanged  in  molecular  collision at biological
temperatures.  This implies that, unless there are
special circumstances (Illinger 1970;  Rabinowitz
1973;  Prohofsky  et al. 1979;  Ginzburg 1968), the
molecular configuration  that  results from the
absorption of energy from the RF field is not unusual.
However, the  relative distribution of normally
occurring  configurations  may be affected by the
absorption of energy. The effect produced may be only
a relative change in a normally occurring process, and
may be reversible on the molecular level. However,
reversibility for the molecular-level absorber does not
imply that  the resulting—if any—macroscopic
biological effect is  reversible. If the  unusual
distribution of states does not return to the normal
distribution  as quickly as one would  expect from
strictly  thermal  considerations, then these effects
will be much more important (Fermi et al. 1965).
                       3-22

-------
 Another general mechanism for the interaction of RF
 radiation with biological molecules results from the
 field-induced migration of ions associated with
 biopolymers. The  migration  may  be of positively
 charged ions—like protons or other cations—from
 one negatively charged site to another within the
 same biopolymer (Kirkwood and Schumaker 1952) or
 from the polarization of the ionic cloud that surrounds
 the  biopolymers (Schwarz 1972). The distinction
 between these  two  mechanisms  has been made
 because in the case of the former, the migration of the
 ion(s) may directly  affect the  function  of the
 biopolymer if the migrating ion(s) or one of the sites is
 directly involved in molecular function. In both cases,
 there  are  effects on the  long-range interaction
 between biopolymers. The minimum field strength
 needed for the first case can be roughly estimated
 from the size and shape of biopolymers, and even for
 rod-like structures, a field of at least 105 V/m would
 be needed.  For the second case, the field strength
 necessary is smaller and depends on the size of the
 surrounding ionic cloud. The  frequencies at  which
 these processes are important depend on the size and
 shape of the molecule (Pollak 1965). For sonically
 fragmented DNA segments  that  have a rod-like
 structure, broad absorption regions have been found
 near 10 kHz and 10 MHz (Takashima 1963;  Pollak
 1965).

A  mechanism has been described for  the  direct
influence of an RF-electric field on the configuration
of  biopolymers (Schwarz 1967). If the interaction of
the field with  one configuration of a biomolecular
system  is much  greater than the interaction with
other configurations, then—in the presence of that
field—a shift in the relative populations of the various
configurations results. The difference in interaction
energy  may be  due to differences  in the dipole
moments or in the polarizabilities between configura-
tions. This mechanism  has been demonstrated in the
model protein poly (x-benzyl-L-glutamate), or  PBLG
(Schwarz and Seelig 1968).  In  its  random-coil
conformation, PBLG has  essentially  no dipole
moment, whereas in its helical conformation it has a
large dipole moment (~ 103 Debye). The presence of a
field of  sufficient strength can induce a transition
from the coil to the helical conformation, in PBLG this
mechanism  is responsible  for absorption of energy
near 1  MHz; for the distribution of conformations to
be affected significantly, the field must be greater
than 105 V/m. (If there were systems in which the
change  in dipole moment was larger, then the field
needed would be proportionally smaller.(This change
in  conformation is not unusual (although the relative
amounts of  the  two conformations are  altered) or
irreversible,  but, as previously stated, reversibility
on the molecular level does not necessarily imply that
any  resulting, macroscopic,  biological  effect is
reversible. For instance, the conformations may have
differing biological activities,  and the products of
these activities may remain even after the distribution
of the conformations has returned to equilibrium

3.2.4.3 Unresolved Issues
The mechanisms for trre absorption of energy from an
RF field previously discussed have been demonstrated
in solutions of biological molecules, and the response
of macroscopic biological systems to these fields is
consistent with these  molecular mechanisms. The
question that remains is: Are any of these mechanisms
likely to impose a potential hazard to human health?
Certainly, if  enough  RF energy  is absorbed and
converted to thermal energy, corresponding biological
effects ensue. Other mechanisms have  been
proposed for effects at the molecular level.  They
require changes in molecular structure and function
as the result of absorption. It is extremely likely that
some changes in biomolecular structure result when
RF  energy  is absorbed. But  are these changes
functionally  significant?  Some  mechanisms for
biological effects in these circumstances have been
proposed. They are plausible but often not quantitative,
and have  not been  demonstrated in  biological
experiments. Further  research to quantitate and
consider these mechanisms  experimentally  is
urgently needed. Since exposure to RF radiation often
leads to increased temperature, careful studies are
needed to separate effects that are not due to heating
from other concomitant changes in temperature.
Two other general mechanisms at the supramolecular
level need further discussion. A theoretical mechanism
has been proposed for coherent  action over long
distances in biological macromolecular systems that
are far from thermal equilibrium (Frbhlich 1968). This
mechanism has been developed to explain the
extraordinarily high catalytic power of enzymes; one
of its  consequences  is an  extremely long-range
interaction  between macromolecules.  The require-
ments  of this model are that the  macromolecules
possess a metastable excited state that has a  large
dipole moment (> 103 Debye) and that its polar modes
couple with elastic  modes. This  highly dipolar
metastable state will be stabilized by the migration of
ions and by the structure  of  the  water near the
macromolecular surface. One of the implications of
this model for enzyme action is that interaction with
an EM  field that is capable of supplying energy above a
critical rate results in all of the energy going into a
single homogeneous electric vibration of this system
(Frolich 1975). This means that all the other modes
of the system are at or near thermal equilibrium, and
that single  mode  is  far removed. Although  it  is
unlikely  that this  redistribution of  energy will
increase the rate at which the energy is absorbed by
the system, the biological consequences of putting all
the energy into a single mode that is related to the
functional  properties of the  system  could be
extraordinary. If that single mode relates to enzyme
function  or  recognition,  it  could  greatly increase
                                                                      3-23

-------
enzyme function  until saturation is reached. The
possibility also exists of re-emission of large amounts
of energy and action at sites remote from the initial
absorber. The specific biological consequences of this
model relative to  RF-field effects on  biological
systems cannot be well  understood until detailed
characteristics of  particular biological  systems are
included. Some data can be explained by this model
(Grundler et al. 1977); the resultant in vivo vibrational
spectral properties of biological systems also have
been discussed (Illinger 1982).

One of the interesting and relevant consequences of
Frohlich's  (1968)  model  is that the  particular
biological result  of RF-field  interaction with  a
biological system may be frequency dependent. It has
been suggested that the range of frequencies where
this process is most likely to be significant is probably
> 30 GHz (Frbhlich 1968, 1975). The actual range of
applicability  is uncertain. The effect also may be
important at lower frequencies.
Frohlich's mechanism is general. Its applicability in
real  or model biological  systems  has not been
demonstrated. For potential biological effectsfromRF
fields, it is  a possible mechanism for grouping
individual photons or phonons with energies « kT
(the average thermal energy). This process results in
the application of  energy in a significant amount (>
kT) at a  single locus.  Effects resulting from this
process could not be duplicated by addition of the
same amount of energy to the system by a different
process.

Another set of mechanisms at the supramolecular
level has been proposed recently. These mechanisms
are theoretical means for the direct interaction of RF
fields with microscopic biological processes that
depend on naturally occurring electric potentials. In
general, these mechanisms depend on a nonlinear
response to an applied field by the cell membrane.
Barnes and Hu (1977) have proposed that the ion
gradient  across a cell membrane can  be altered
significantly if PW radiation at  105 V/m peak field
strength  or  more is applied. This theoretical view
derives from consideration  of the balance  between
field-driven  and thermal currents. Changes  in the
time-averaged concentration gradient occur that
depend on the square of the field applied across the
membrane, but not on  the frequency of the applied
field. There  are  larger terms  that are frequency
dependent but when averaged over a complete cycle
are zero. This mechanism has not been demonstrated
in a biological system.


Pickard and Rosenbaum (1978) considered a similar
problem and came  to the same conclusion for the
time-averaged concentration gradient across the
membrane.  In their model,  they included unidirec-
tional ion channels through  the membrane, and they
postulated that the new concentration gradient is
achieved by the  field-induced  movement of ions
through these channels. They calculated that, for this
particular  method of  achieving  the  concentration
gradient, the frequency of the field must be < 200
MHz if the ions are protons, and < 10 MHz for less
mobile  ions. Pickard and Barsoum (1981)  have
demonstrated activity across the membrane of a plant
cell that begins at the onset of an RF field greater than
667  V/m  with a  frequency  below  10  MHz. The
immediate response to the field indicates  that the
activity is not thermally induced. The  activity
demonstrated is consistent with  an RF-induced  DC
potential across the membrane. This potential  could
result from any nonlinear response of the membrane
to the RF field, and it is not clear that any particular
mechanism for that response has been demonstrated.
More data are needed  to  demonstrate  a  realistic
microscopic model for this process. It is difficult to
proceed from these mathematical models  and this
single experimental demonstration to an assessment
of the importance of this mechanism. The induction of
DC  or extra-low-frequency potentials in biological
systems by  an RF field could provide  significant
mechanisms for the biological effects of RF radiation.
                       3-24

-------
 3.3 Experimental Methods
 Claude M. Weil
 Joseph S. AH

 3.3.1 Exposure Methods Used in Biological
 Experimentation
 In this section  the most commonly used exposure
 methods employed in RF biological  effects research
 are reviewed,  and  the  inherent advantages  and
 disadvantages of each are  discussed. Researchers
 have  used a diversity  of  exposure  methods; the
 chosen method depends on the specialized application
 involved, the existing facilities available for experi-
 mentation, and  the cost of setting up new facilities.
 This diversity  may  have  created difficulties in
 reproducing biological effects studies, which have, in
 turn, fueled controversies on several reported
 effects. Whereas an effect may have been noted by
 one group  who  employed a  particular form of
 exposure system, a second group that carefully
 reproduced the experiment in every  detail but for the
 exposure method might not  see the  reported effects.
 It is possible that basic differences  in the exposure
 environment may, in some cases, explain this lack of
 reproducibility.

 In much of the literature, effects have been traditionally
 reported as a function of incident power density, the
 independent variable. As pointed  out by  several
 prominent workers (King et al. 1970; Johnson 1975),
 power density is an  inadequate and inappropriate
 independent variable  for two reasons:
 1) Biological effects are logically related to the
   internal electric and magnetic fields which are in
   turn nonlinear and complex functions of frequency
   vs. body size, body shape, and orientation to field
   vectors, etc.  (See Sec. 3.2, RF Field Interactions
   with  Biological Systems.)
 2) Although  meaningful for certain plane-wave-
   exposure situations, such as free-field or
   transverse electromagnetic (TEM)-mode cells, the
   power density concept is no longer applicable in
   most  other  commonly used  exposure systems
   with complex irradiation fields. Consequently,
   meaningful  comparisons of  the incident  field
   levels  in different types of  exposure systems
   cannot be made on this basis.
For these reasons,  it must be concluded that the
various exposure methods are unique; experimental
results obtained in different exposure environments
cannot be compared except perhaps on a whole-
body-averaged  SAR basis, and  reproducibility of
experimental data is not guaranteed unless exactly
the same exposure methodology has been used. Even
if equivalency of whole-body-averaged SAR is
achieved, there  may still be differences in biological
outcomes for exposures performed in multipath fields
(e.g., a cavity) when compared to plane-wave fields,
which  have the same  carrier  frequency  and
modulation parameters.

Although many of the specialized requirements of a
particular experiment are unique, some requirements
are common to almost all experiments. For example,
there is a  critical  need  in  virtually all work for
temperature and humidity  regulation of the environ-
ment in which animals, biological  specimens,  etc.,
are located. Such control ensures that the ambient air
temperature and humidity are properly controlled, so
that the possibility of  unwanted variability  in  the
experimental protocol is minimized. Also, airflow is
an  important variable, particularly in animal studies
that should be regulated  and reported. Two other
important requirements are common to the majority
of experiments involving whole-body or  unrestricted
irradiation of animals: (1) an animal population large
enough to allow for statistically reliable conclusions
to be drawn, and (2) an equally large population of
control animals that must be housed during exposure
periods  under conditions  that are,  in every way,
virtually identical to those of irradiated  animals
exeept for the presence of RF energy. For certain
critical experiments, there may also be a need for a
population of passive or reference control animals
that are maintained under normal conditions in their
vivarium during the tenure of a study. Much of the
early  experimental  work  on RF biological effects
frequently  contained  no provision  for controls,
involved inadequate  numbers of  experimental
subjects, and lacked adequate provision for environ-
mental control, and thus yielded data of questionable
validity. It is only within the past decade or so that
researchers have succeeded in correcting most of
these experimental deficiencies in exposure protocol.

Existing exposure methods have  been reviewed by
Weil (1977) and Ho et al. (1976);  the  latter paper
discusses exposure techniques that have been used
at the FDA's  Bureau of Radiological Health.  The
commonly used methods are categorized  into  two
groups and considered separately: (1)  free-field, and
(2)  enclosed. In the former category, the exposure
fields  are  generated within a  reflection-free
environment by a radiating  antenna  or applicator,
whereas in the latter category, the exposure fields are
those excited within a struture  of reflective or
conducting walls.
3.3.1.1 Free Field
Free-field  exposure methods were historically the
first to be  used in such experimentation and remain
the most commonly used techniques. These methods
involve  placing the experimental subject(s) in the
near-field  or far-field region of a  radiating antenna,
                                                                      3-25

-------
such as a  horn, parabolic dish, or diathermy
applicator. For exposures in the far-field region, all
unwanted scattering of RF energy from the surround-
ing  room structure, etc., must  be eliminated.
Otherwise, the power density levels at the subject's
location are  not well established. A  microwave-
absorbing material that wholly or partially surrounds
the experimental subjects must be used to create
what is termed an anechoic chamber. (The term
 a   jnuic,"  meaning "without  echoes," has been
borrowed from  acoustics research.) The design of
anechoic chambers  is a specialized  art normally
performed by commercial manufacturers of absorbing
material (e.g., Emerson and Cuming, Canton, Mass.).
The anechoic properties of such chambers, which are
defined by the  characteristics of the "quiet zone"
where the experimental subjects are placed,  is a
function of the reflective properties of the absorbing
material with which  the chamber is lined. (Such
properties refer to the degree with which RF energy is
reflected by the  material,  rather  than absorbed.)
Reflectivity is proportional to the incident wavelength,
so that the  performance of an anechoic chamber
deteriorates as the frequency of operation is lowered.
A minimum reflectivity of 20 dB (1 percent) is usually
considered essential;  the frequency at which this
value is reached then defines the  lower frequency
limit of operation for the  chamber. Anechoic
chambers perform  well  at frequencies  in the
microwave region of the spectrum (above 500 MHz).
However, satisfactory performance at lower frequen-
cies  or longer  wavelengths  requires  using  large
blocks of absorbing  material  formed into long
pyramids or cones. Consequently, a large structure is
needed to house such a chamber, so that the cost of
such a facility rapidly  escalates  to the point of
nonfeasibility.  For all intents and purposes, it is
impractical  to  contemplate building anechoic
facilities that operate at frequencies below ~ 250
MHz.

Some  of the most extensive anechoic  chamber
facilities devoted to RF biological effects work are to
be found at the Department of Microwave Research
of the Walter Reed Army Institute of Research, Silver
Spring, Md.  Similar conventional anechoic facilities
are  found  elsewhere;  e.g., at  the  Bureau of
Radiological  Health in Rockville,  Md.  at the  U.S.
Environmental  Protection Agency in Research
Triangle Park, N.C. (Figures 3-16 and 3-17), at the
U.S. Air Force School  of Aerospace Medicine in San
Antonio, Tex., and at the Universities of Pennsylvania,
Utah, and Washington.

The  principal advantages—applicable to exposures in
only the far-field zone of an antenna—of the free-field
exposure method are:
 • The incident power density to which the subject is
   exposed can be well defined to an accuracy of ±1
   dB (±20 percent) or less. This makes replication of
   experiments performed under free-field conditions
   fairly straightforward.

   Depending on the directionality of the radiating
   antenna, the exposure area where the irradiated
   subjects are placed is large, and the irradiation is
   nearly uniform (±10 percent or less).
The  principal  disadvantages  of free-field exposure
methods are:
 • The facilities are costly and often  require
   extensive floor space.

 • Very-high-power RF sources are needed to obtain
   the necessary power density levels. Consequently,
   many free-field facilities employed in bioeffects
   research operate on either of two ISM frequency
Figure 3-16.    EPA 2450-MHz anechoic chamber facility, or
             2.45-GHz far-field exposure facility. Horn
             antenna (top),  temperature control chamber
             (center), and one air duct (bottom left) are
             visible in the shielded anechoic room (Blackman
             or a/. 1975. Elder and Ali 1975).
                       3-26

-------
 Figure 3-17. EPA 2450-MHz anechoic chamber facility: Diagram of the microwave exposure facility (Blackman eta/. 1975; Elder
           and AH 1975)

Command Signal Power Command __
i
Microwave
Microwave Power Sense Power

Data
Acquisition
System
I

Controller Control and Sense Lines


Varian
Microwave
Generator
2450 MHz
3.0 kW CW
Microwave Power Sense
y
Parameter i^
Commands^ fm,mn 	 r, Air In IX
Control ""j^
Package Air Out 1
Environmental '
—• Parameter ^_____^__
Sense '

I 	 '
~w,,\
~ ' /
^ Animal /
^Exposure ^
Chamber^

Physiological Data

1
l^\
\
1
1
1
_l
                                                                               Anechoic
                                                                               Chamber
                                                                               3 m Long
                                                                               3 m Wide
                                                                              4.5 m High
   assignments used for microwave heating, 2450
   MHz or 915 MHz, where high-power sources are
   commercially available at relatively low cost. It is
   for this reason, as well as the explosive growth in
   the use of microwave sources at ISM frequencies
   (e.g., microwave ovens and hyperthermia treat-
   ment devices) and  the resulting potential  for
   exposure of significant portions of the population,
   that 2450  MHz has  become an important
   frequency for bioeffects research.

   Free-field systems are an excellent simulation of
   the idealized exposure concept visualized in many
   protection standards, but these systems do not
   represent the real-life exposure situation en-
   countered by humans.

   The whole-body-averaged SAP cannot be mea-
   sured directly in a free-field situation. Reasonably
   accurate  estimates of the SAP can be made by
   calorimetric measurements on animal carcasses.
   (See Sec. 3.4, Dosimetric Methods.)
 Georgia  Institute of Technology developed several
 such techniques. One employs the absorber-lined
 horn antenna, which enables a subject to be placed
 much closer to the horn aperture than is the case with
 a conventional horn and yet to remain within the far-
 field zone of the radiator, so that a source of lower
 power can be used to create the same power density
 at the irradiated subject. One of these absorber-lined
 horns has been used for several years at the National
 Institute of Environmental Health  Sciences in
 Research Triangle Park, N.C. (Figure 3-18). Another
 technique involves the compact range (Figure 3-19),
 which uses  a feed horn to irradiate a large parabolic
 reflector. The test subject is exposed in the collimated
 beam region of the reflector.  The advantage of this
 system is that it creates a much larger test volume of
 nearly uniform field intensity with associated
 decrease  in the anechoic facility size,  as well as
 reduced  requirements for absorbing materials. The
 compact-range technique is used at  the Naval
 Aerospace Medical Research facility in  Pensacola,
 Fla.
Various  attempts  have  been  made to modify
conventional free-field exposure methods to amelio-
rate the cost and space problems associated with
them. A study (Bassett et al. 1971) performed at the
As discussed earlier, a critical need in whole-body
exposure of animals is to expose large numbers of
animals.  Consequently,  in  many experiments,
unrestrained animals have frequently been placed
                                                                      3-27

-------
Figure 3-18.  Diagram of absorber-lined horn (Bassett et al. 1971).
                                         RF Absorbing
                                         Material
                  Section A-A'
                                                                       -•- A
                                                                     26"
                                                                                        26"
                                                                      L,
inside  an anechoic chamber  in a  closely spaced
matrix for simultaneous exposure. This practice tends
to obviate the principal  advantage  of  free-field
exposure methods,  because the power density to
which  the animal  is exposed can no longer be
accurately defined due to RF-energy scatter from one
animal to another. The degree of scatter depends on
much the same factors as absorption (see Sec. 3.2.1)
and varies randomly as the animals constantly alter
their relative posture and  orientation. Interanimal
scatter is  most serious when the frequency of
exposure  lies in the resonant and supraresonant
region. Several solutions to this problem have been
devised:
(a) sequential exposures of single animals in the
   same free-field facility

(b) miniature anechoic chambers for exposing
   animals individually

(c) separating the restrained animals so that they are
   spaced one or more wavelengths apart
Figure 3-19.  Diagram of a point-source compact range (Bassett et al. 1971).
                                                                                       Reflector
                       3-28

-------
The  first  solution is  obviously  costly and time-
consuming, and may  not be acceptable in certain
critical biological experiments requiring all animals to
be exposed at exactly the same age or the same time
of day, etc. The second solution has been advanced by
Guy (1979) at the University of Washington in
Seattle. He has described a system of 16 miniature
anechoic chambers designed for chronic low-level (<
10 mW/cm2) exposure of rabbits and rodents at 2450
MHz; eight of these chambers are energized, and the
other eight serve as control chambers (Figure 3-20). A
similar chamber has been constructed at EPA (Figure
3-21). Such a system has, of course, the advantage of
ensuring total RF isolation between animals and,
although costly, it may be more economical in cost
and  power requirements than is the conventional
anechoic chamber.

The third solution has been tried by a few workers and
appears to be a reasonable compromise between the
conflicting requirements of adequate isolation and
low cost. Oliva and Catravas (1977) have described a
method for simultaneously exposing 10 animals in a
conventional anechoic chamber with minimal field
interaction between  animals (Figure 3-22). The
animal cages are located  on the antenna beam's
three-dimensional contour of  constant power
density, with cages separated by ~ 2.5 wavelengths
at 2450 MHz. The authors claim that  the average
variation of power density from one cage location to
another does not exceed  ±5 percent. The major
disadvantage of this technique is that its implementa-
tion may require a large exposure  volume, which, in
turn, creates costly problems of ensuring satisfactory
environmental control. D'Andrea et al. (1979, 1980)
at the University of Utah have developed an alternate
method involving a monopole radiator on a vertical
ground plane. The animals  are located next to the
ground plane on a circular locus that is centered at
the  monopole (Figure 3-23).  Each  circular array
contains 10 animals that  are separated by ~ 5.5
wavelengths at 2450MHz, and by 2.1 wavelengths at
915  MHz. This  system offers the  advantage of
reducing both space and RF-power  requirements.
Furthermore, the presence of a ground plane makes
this  system more representative of the real-life
exposure situation for  humans. On the other hand,
the exposure characteristics of this system may not
be comparable to a conventional anechoic chamber
facility, so that caution  is needed when data are
compared.
For situations involving the exposures of only one
subject at greatly increased power density values,
D'Andrea et al. (1977) and Hagmann  and Gandhi
(1979) have discussed a modification of this system in
which a corner  reflector  is placed  behind the
monopole. By placing  the subject  1.5 wavelengths
from  the corner point,  one can get an  SAR
enhancement of more than 20 times that obtained
without the reflector.
3.3.1.2 Enclosed Systems
Enclosed exposure systems are those in which the
EM fields are contained within a  conducting
structure. They fall into two basic subcategories,
which are considered separately: (a) transmission-
line systems, and (b) cavity systems. The principal
advantages of enclosed systems are as follows:
 • By suitable monitoring of the incident, reflected,
   and transmitted power  in the exposure system,
   more accurate SAR estimates can be made for the
   exposed  subject than  is possible  in free-field
   facilities. The accuracy in  estimating SAR
   depends in part on how much base-line power loss
   there  is for the unloaded system and to what
   extent it is affected by the presence of the exposed
   subject.

 • The cost and space requirements of such a system
   are usually lower than those for free-field
   methods, depending on the number of identical
   systems  needed. Also,  much smaller  and less
   expensive power sources will suffice to establish
   the same field  intensities.

 • Better control of RF absorption in the irradiated
   subject is possible.
The principal disadvantages are as follows:
 • With the exception of TEM-mode transmission
   lines (to be explained), the fields are complex and
   not plane-wave equivalent, so that comparison
   with free-field data is more difficult.

 • With the exception of multimodal systems (also to
   be explained), the exposure space available inside
   such systems is limited. Although the available
   space may be adquate for small animals such as
   rodents, it may not be possible to fit a rabbit or
   monkey in it. Thus, the variety of species that can
   be employed in an existing facility of the enclosed
   type is restricted.

 • The field distribution within the exposure system
   is rarely as uniform as that found  in a free-field
   facility. The field distribution is well defined for
   single or dominant mode operation, but undefined
   for multiple-mode operation.
Enclosed systems have been the preferred exposure
method for samples, specimens, in vitro preparations,
                                                                     3-23

-------
Figure 3-20.  Miniature anechoic chamber facility (Guy 1979).
              Tapered Radiation Chamber
                 Front View E-Plane
                         32.4
                         cm
        11.43cm square
          Muffin Fan
                                    NARDA
                                    644 Horn
 2.0 cm sq Horn
 Mount Centering ^26.6 cm"
 Strip (Wood)
1375cm.
Distance from Horn
to SPY-12 Tips

           Door
.Styrofoam  Fastener
Floor
(Top View)
  60.9
  cm
  Typ.
FXR
S601B
Adapter
      0.32 cm
          Tapered Radiation Chamber
              Side View H-Plane
                    33.6  j       r—t-4cm
                        H
All Aluminum
Strips and Angles
Assembled with
0.32 cm Pop Rivets
0 95cm  Hardboard
Plywood
                      Plywood Horn Mount
                      (with 2 Alignment
                      Centering Strips)    /
                      Sits on Chamber    /
                          Separation Point
                                   90.4 cm
                   016cmX
                   2.54 cm Alum Strips/
     0.32 cm (0.125 in)
     Hardboard

       Door
       75.5 cm X 60.9 cm
  92.7 cm
     2.54 cm X 7.6 cm X
     20.3 cm Styrofoam
     Brace Bonded to
     AN-77 With
     Silastic
     Adhesive
                                                                   5.08 cm square
                                                                   0.95 Plywood
                                                               40W Candelabra Lamps
  T\
26.6cm 4 each mounted 4cm
       'From Each Edge

\ 0.95 cm
  Plywood
       Plywood Parts
       Glued & Nailed
                                                              0.32cm (0.375 in)
                                                              Hardboard

                                                              AN-77 Absorber
                                                              5.08 cm (2.0 in)


                                                              1.27 cm (0.5 in)
                                                              Aluminum Angle
                                                                                                 Square Air
                                                                                                 Intake Hole
a) Construction details of tapered exposure
  chamber.
                                                       1.27 cm Plywood
                                                       Plywood
                     b) Construction details of tapered exposure
                        chamber.
                                                                  /2 Scale
                                             -R = 10.32
                                            Weaved
                                            Nylon Line

                          c) Construction details of cage for housing
                            rat in tapered exposure chamber.
                           3-30

-------
Figure 3-21.   Photograph of tapered exposure chamber at EPA facility.
                                                                                 3-31

-------
Figure 3-22.  Facility for simultaneous exposure of 10 animals with minimal inter-animal interaction (Oliva and Catravas 1977).


                                   0               4            S           6


                                                                                       r- -1.5
Meters
                              Feet
                                   0   1
                Distance from Antenna
                  .   13   14 15  16  17  18  19
                      _0_ Locus of Equal Power
                       	Density (calculated)
                       L_l Cage on Tall Pedestal
                       EZ] Cage on Medium Pedestal
                       •• Cage on Short Pedestal
                                     r
                      a) Locus of equal power density in the H plane through
                         the axis of the transmitting antenna.
                                                           -  -1.0


                                                           -  -0.5


                                                           -  0


                                                           -+0.5


                                                           -+1.0
                                                           —1-1.5
                                                           Meters
                                   0
                            Meters L_
                      4
                       I
5
                              Feet
                                            Distance from Antenna

                                   01          13   14  15 16  17  18  19
                      —•— Locus of Equal Power Density
                      I    I Cage on Antenna Axis
                      |~~j Cage Intermediate Distance
                      — from Antenna Axis

                      Hi Ca9e Furthest from Antenna Axis
                                                    +3S
                                                                                 +1 S
                                                                                 -11
                                                                                    0)
                                                          r+1-0
                                                           -+0.5
                                                                                       - 0
                                                     -3
                                                   Feet
                                                                                       L-1.0
                                                                                       Meters
                      b) Locus of equal power density in the E plane through the
                        axis of the transmitting antenna.
                                              • s-gSfca-^jfisH
                                              ,i"55SiMykjMsiiasr-
                                               • • • ?.<• 9. t'--". '.  .
                      c) Multiple-animal array for equal power density
                        microwave irradiation.
                           3-32

-------
Figure 3-23. Monopole-over-ground plane irradiation facility
           (D'Andrea et at. 1979. 1980).
Ground
Plane

Monopole
Antenna
Styrofoam
Mounts &
Rat Cage
Walk-On
Eccosorb
                                        Pyramidal
                                        Eccosorb
a) Representation of the microwave-exposure chamber.
       Left Chamber

      Mean mW/cm2
      5.42 ± .21
Right Chamber

Mean mW/cm2
4.85 ± .27
                  SAR -mW/g
                  1.23 ±.25
                    (n = 3)

b) Distribution of power densities measured at each rat's location
  within the Plexiglas holding cages for both sides of the
  microwave exposure chamber in the absence of the rats
  (2450 MHz).
etc., since these materials are small enough to fit
easily inside such systems. Many experiments have
been described,  too  numerous to survey here, in
which various types of waveguide, TEM-mode lines,
and cavities have been used to  irradiate such
preparations. In  recent years, some enclosed-type
systems have been specifically  designed  for
whole-body irradiation of animals. They are discussed
here in  more detail. Because of the  need to define
accurately the field  level to which  the animal is
exposed, dominant-mode operation of these devices
has  been preferred  over multimode operation.
However, dominant-mode  operation puts an upper
limit on the frequency of operation, depending on the
size of the system needed to expose an experimental
animal. For example,  most of the dominant-mode
enclosed systems that have been used to expose rats
operate at frequencies below 1000 MHz; for mice,
which require a smaller exposure volume, single-
mode operation  to ~  3000 MHz is possible. Since
free-field methods of exposure are largely impractical
at these lower  frequencies,  as  discussed earlier,
enclosed systems do provide experimenters with
virtually the only practical and economical method of
exposing animals in the HF,  VHP, and lower UHF
segments of the RF spectrum. A few large-scale
systems have been built to operate in the dominant
mode at HF and lower VHF frequencies; they are large
enough to accommodate a group of animals within
the  exposure  area,  and  the animal-to-animal
interaction problem is minimized because exposures
are conducted  in the subresonant  region where
interanimal scatter is  minimal. However,  at higher
frequencies such methods become impractical, and it
becomes necessary to provide an identical exposure
system for  each animal  to be  irradiated. Each
exposure device  is  fed from a single  high-power
source by a power dividing  network. One of the basic
electrical problems  associated with  any enclosed
exposure system is that,  when  loaded,  there is
considerable reflection of RF energy from the load,
which propagates via feed lines back to the RF power
source.  This  characteristic creates particular
problems for multiple systems fed  by a single power
source because, without sufficient isolation between
each exposure device, energy  is continuously
reflected from one system into the others and causes
unwanted fluctuations in the  RF energy incident to
each  system.  These problems can  be  solved by
installation of isolators in each feed line to prevent
the reflected energy from reaching the dividing
network and the RF power source. The two categories
of enclosed system are discussed separately.
                       Transmission lines—In these systems, an EM wave
                       propagates down the line from source to termination.
                       When  the  line is terminated  in  its characteristic
                       impedance,  only  the  outwardly propagating wave
                       exists with no  reflected  wave present. (The
                       characteristic impedance  is a  basic electrical
                       parameter that is defined by the line's configuration
                       and dimensions.) However, if the line is terminated in
                       a short circuit or an open circuit, all the energy is then
                       reflected at the line's termination. The two systems of
                       traveling waves together create a standing wave that
                       possesses successive maxima  of electric and
                       magnetic fields  (E  and H, respectively) spaced a
                       quarter of a wavelength  apart. (This property is a
                       useful feature that has been used by some workers to
                                                                      3-33

-------
expose specimens or samples to isolated E or H fields,
depending on  the  location along the  line.) The
characteristics of a shorted or open transmission line
are similar to those of the resonant cavity, and are
discussed below.

Transmission-line devices can be subcategorized into
TEM-mode or  waveguide. In TEM-mode  lines-
coaxial air lines, parallel-plate (strip) lines, etc.—the
dominant  mode of operation is a propagating TEM
mode, provided the line is properly terminated  in its
characteristic impedance. In TEM-mode operation the
E- and H-field  vectors are mutually orthogonal, and
they lie only in the plane transverse to the direction of
wave  propagation. The resulting fields are virtually
identical to those of a plane wave, so that it is possible
to simulate free-space conditions inside such lines,
provided  that the  exposed subject does not occupy
more than about a third of the line's cross-sectional
area. (If the one-third restriction is exceeded, ground-
plane effects are  observed.) Because of its circular
cross section, the coaxial air line has been used only
with fluid or  precut  solid samples;  the primary
application has usually been in making  dielectric
measurements. A  sophisticated coaxial air-line
system, designed to expose cell cultures to high field
strengths, has been described by Guy (1977) (Figure
3-24). Parallel-plate systems consisting  of a
conductor located over a ground plane or between
two ground planes are better  suited  to  animal
exposures since they provide a rectangular volume
where animals can be exposed. D'Andreaefa/. (1976)
have described a simple parallel-plate system used in
the range 200 to 500 MHz (Figure 3-25). This system
is equivalent to a  large-scale version of a microstrip
line with air dielectric. Capacitive-plate devices,
which are equivalent to a short length of parallel plate
that is terminated in an open circuit, have been used
for exposing in vitro preparations to isolated electric
fields. This type of simple device has been incorporated
in a  unique  HF-band (10-  to  40-MHz)  exposure
system developed by  the National  Bureau  of
Standards (Figure 3-26). Termed a near-field
synthesizer, it  is capable of separate E- and H-field
excitation, the latter generated by special loop
inductors  (Greene  1976). By altering the spatial
position of the loop with respect to the E vector and by
changing the phase relationship between the signals
exciting the plates and the loop, one can simulate the
complex field environment existing in the near field of
an HF source. Two of these systems are being used at
present:  one  is at the U.S.  Afr  Force  School of
Aerospace Medicine,  Tex., and  the other at the
National  Institute  of Occupational Safety and Health
laboratory in Cincinnati, Ohio.

Triple-plate lines, in which a conductor is mounted
between two ground planes, have been used by a few
workers.  These lines have the advantage of allowing
for  simultaneous irradiation of two animals while
maintaining good isolation between them. In a more
popular version of this system,  the open sides of the
structure are closed off with two side plates so thai the
center plate is entirely surrounded by a rectangular
ground structure.  Such  lines are  commercially
available in different sizes  (from  Instruments for
Industry, Inc.,  Farmingdale. N.Y.)  and  have, been
termed "Crawford cells" by the manufacturer, after
M.L.  Crawford  (1974) of the  National  Bureau  of
Standards, Boulder,  Colo., (also available  from Narda
Microwave  Corp.,  Plainview,  N.Y.).  This author
prefers the term "rectangular strip line."  In addition,
a few large-scale facilities of this type have been built
for  various  specialized exposure applications  at
relatively long (^ 3 m) wavelengths. The largest is at
the Defence Research  Establishment  in Ottawa,
Canada, and  is being used to measure  human RF
absorption in the HF band. A smaller facility has been
used  at the U.S. Air Force School  of  Aerospace
Medicine for exposing  infrahuman  primates and
phantoms to  10- to 50-MHz radiation (Allen et al.
1976). The U.S.  Environmental Protection Agency
has also built a Crawford-type structure (Figure 3-27)
for exposing 20 rats to 100-MHz radiation.

The second  category  of transmission-line devices,
waveguides, cannot  support the propagation of a TEM
mode so that the incident radiation is inevitably more
complex than that in a TEM-mode line. However, the
lack of a center conductor means that more exposure
space is  available in  waveguides than in the
corresponding TEM  line. Multiple systems that use
commercial waveguide  systems have  not been
applied much  for animal whole-body  irradiation
because of high cost. Hoe? a/. (1973) have  described a
single, environmentally controlled system employing
a standard S-band rectangular waveguide that has
been used in studies of mice at 2450 MHz. Ho et al.
(1976) have also described a system of six waveguides
fed from a common  2450-MHz source that provides
simultaneous irradiation  of six mice. Besides cost,
another problem associated  with using standard
waveguides for whole-body irradiation is the  high
mismatch condition created by the presence of an
animal in the waveguide. As discussed earlier, this
necessitates the use of costly isolation circuitry. This
problem has, to a certain extent, been eliminated in a
specialized waveguide system developed  by Guy and
Chou  (1976) at the University of Washington  in
Seattle. Guy's  system is specifically tailored for
chronic whole-body irradiation of  rats for extended
periods of time with known and  reproducible
dosimetry. It consists of a number of 20-cm (8-in)
diameter circular waveguides (Figure  3-28),  eco-
nomically constructed from  galvanized  wire mesh
and  short  lengths  of brass  tubing, in which a
circularly polarized TEn dominate mode is excited at
915 MHz. Figure 3-29 is a photograph of  this type of
facility. The use of circular polarization (see Sec. 3.1,
                       3-34

-------
Figure 3-24.   Coaxial air-line system for high power exposures of cell cultures (Guy 1977).
                                         First Surface
                                         Mirror     —

                                        Current Sensing
                                        Probe
Thermograph
                   Constant Temperature
                   Liquid Source & Pump
                                                                  • Culture Sample Holder
                                                                Voltage Sensing Probe
                                                                             Vector Voltmeter or
                                                                             Network Analyzer
                                                                            Digital Voltmeter
                                                                            Temperature Monitor
                                                                Directional
                                                             / Coupler
                                                                                           Feedback
                                                                                           Loop
                                                  Incident
                                                  Power
                a) Complete system for exposing cell cultures to EM fields.
                  Outlet


r-,

3
ecm |



r-,

)
twi V 1 Stainless fitael
                                                              vattti&  Brass
                                                                      Cross-Linked Styrene


               b) Cross-sectional view of assembled transmission-line cell-culture sample holder and
               heat exchanger. (A section of the outer conductor is machined to 7.70-cm ID so a
               7.71-cm diameter dielectric support can be firmly locked into place after being pressed
               in while the conductor is heated.)
                                                                                          3-55

-------
 Figure 3-25.  Parallel-plate (microstrip) exposure system (D'Andrea et al 1976).
Figure 3-26.  Block diagram of the complete RF near-field synthesizer showing all principal components including RF power
             sources.
          Balanced
            Feed
         to Parallel
           Plates
                             Lower
                             Plate
RF Power
Ampl.
No 1


Doubler
Ampl
No 1


Driver
No 1
Driver
No. 2


Doubler
No 2


RF Power
Ampl.
No 2
                                                       Common      i
                                                       Oscillator
                            3-36

-------
Figure 3-27.  EPA 100-MHz rectangular strip line or
            Crawford cell.
                              Electromagnetic Field Theory) has two major
                              advantages:
                              1) The wide variations in absorption due to constant
                                changes in the animal's orientation are partially
                                reduced so that the energy coupled to the animal
                                load  remains relatively constant. Note that.
                                whereas in a linearly polarized free-field of perfect
                                field uniformity, the average SAR can vary by — 5-
                                8 dB  (~50 percent) at or near resonance, in this
                                system,  the  average variation  of SAR does  not
                                exceed —0.4 dB  (~8 percent),  even though  the
                                incident field uniformity is far from constant
                              2) The hybrid coupler used to establish the circular
                                polarization in the waveguide provides a way to
                                isolate reflected from  incident energy. One
                                achieves the  isolation by dumping most of  the
                                unwanted  reflected energy into a load connected
                                to the hybrid, rather than allowing it to return
                                through the power dividing network to the source.
                                This technique effectively reduces the  mismatch
                                normally encountered with  loaded waveguides
                                (mean voltage standing wave ratio of less than
                                1.5:1) and  permits parallel operation of multiple
                                units  with  less isolation circuitry.
                             Guy .er a/. (1979) have described a modifiction of this
                             waveguide system  such  that it  operates in a
                             multimode field configuration at 2450 MHz. Though it
                             is difficult to define quantitatively the incident energy
                             levels in  these units because of the complexity of the
                             fields, the absorption characteristics have been well
                             defined for rats  of varying weights. An  extensive
                             system of  100 of the$e units has been installed at
                             Guy's laboratory (Guy et at.  1980a) for conducting a
                             chronic  radiation study  in which rats will  be
                             continuously irradiated over a 3-year life span. This
Figure 3-28.  Physical details of the exposure chamber of the circularly polarized 915-MHz waveguide facility (Quy and Chou
           1976).
       RG 58
       Cable
 : 40.25cm    35cm j
 Aperture for
Food Dispenser
                    63 Square
                      Mesh
                                                               Shorting Plate
                11.1cm
                 Mount
                   Bolt
 •HH4  •    . '    L  '

±JJtJ.\|
^	298cm	j  \
nq               V
                                                              1.27cm Dia.
                                                                Tubing
                                                              7.23cm Long
                                                       20.32cm -t
                                           Epoxy
                                          Capsulate * x Hybrid
                                                                                         1.27cm Dia.
                                                                                           Tubing
                                                                                         3.8cm Long
                                                2.5cm Long Section
                                                 8" Brass Tubing
                                                                       3-37

-------
Figure 3-29.  Exposure chamber with associated instrumentation for the 970-MHz circularly polarized waveguide facility at EPA.
study underlines  a major  advantage of Guy's
waveguide systems, animal subjects are able to live
continuously in their exposure system with minimal
disturbance  to  their normal living patterns and
without artifactual  perturbation. Consequently, the
system is  ideal  for  the kind of chronic, long-term
studies that are  now being emphasized.
 Cavities—The resonant cavity, which is the microwave
 analog of the lumped-element tuned circuit, consists
 of  a cylindrical,  rectangular,  or square  box  with
 conductive walls.  By suitable means,  a system of
 standing waves may be excited within the cavity. By
 adjusting the cavity dimensions, the system can be
 made to resonate at the frequency of excitation, such
 that intense field levels are created within the cavity.
 The use of the resonant cavity as an exposure device
 has both a major advantage  and disadvantage. A
cavity system represents the most efficient way to
couple RF energy  into a subject per watt  of input
power; this coupling  constitutes  its principal
advantage. The primary disadvantage is that the field
structure within a resonant cavity  is probably the
most complex to be found  in any exposure device.
Although  some researchers have argued  that the
field structure within a resonant cavity,  particularly
when operated multimodally, is a good simulation of
the complex multipath environment in which humans
are often exposed, it  is unlikely that the human
exposure environment ever approaches this level of
complexity. For the reasons already discussed, the
field complexity existing within a cavity makes  it
difficult to compare experimental data obtained in a
particular cavity with those  obtained in free-field
exposures or even another cavity. Indeed, there are
some basic differences between a free-field exposure
system and  a resonant cavity. For a resonant cavity
                        3-35

-------
that is empty (unloaded), or that contains a biological
load  small  in  comparison to the cavity's volume
(minimal loading), the fields  in the cavity are in a
standing-wave condition and do not propagate; the
energy within  is only stored.  When the loading is
greater (i.e., when the exposed subject occupies an
appreciable fraction of the cavity's volume) the fields
are hybrid, containing both stored and propagating
energy. Because the  fields are in a standing-wave
pattern,  minimally  loaded cavities are  useful for
simulating a quasi-static interaction in which object
size is small  relative to  wavelength (i.e.,  a  long-
wavelength type exposure). This characteristic  has
been used by Guy et al. (1976), who designed a large
rectangular  cavity that  could be  excited either
separately or jointly in the TMno or TEioz modes at
frequencies of ~ 147 MHz (Figure 3-30). This system
has been used to study absorption  distributions in
scaled-down phantom models of man independently
exposed to electric or  magnetic fields, depending on
the mode of excitation. Single-mode circular cavities
have also been used by Edwards and Ho (1975) for
exposing the head of a monkey at 385 MHz. Guy and
Korbel (1972) have discussed the dosimetric aspects
of a rectangular cavity used to multiply expose rats at
500 MHz and have sought to emphasize some of the
problems associated with cavity exposure systems.
Recently,  Spiegel et al. (1980b) described an
interesting variable-volume  cavity that contains
movable walls such that the  cavity can be tuned
across a wide frequency range. The system is capable
of simultaneously exciting two independent cavity
modes and can be used to synthesize complex  EM
fields.
Multimodal cavities, such as microwave ovens, have
been used frequently for experimental purposes
because  of  ready availability  and  modest  cost
(Justesen et al. 1971). Where ovens have been used,
they have been extensively modified to achieve the
experimenter's objectives. Multimodal cavities have
often been preferred over single-mode types because
of the larger exposure volume available and the more
uniform field distribution. The latter is achieved by a
mode  stirrer that helps average out the field
variations  within  the  cavity.  The degree of  field
uniformity achieved  in  these devices  seems to
depend greatly on its design. Heynick et al. (1977)
have described a multimodal cubical cavity suitable
for irradiating nonhuman primates at 2.45  GHz
(Figure 3-31). Each unit possesses its own regulated
power source. Twelve of these units have been used
at SRI International,  Menlo Park, Calif., for chronic
irradiation of a group of squirrel monkeys as part of an
EPA-sponsored project.  Each cavity housed two
animals plus offspring.


3.3.1.3 Conclusions and Unresolved Issues
Diverse exposure  methods are  used  in  biological
effects experimentation, most of which differ
considerably in their field characteristics. Since each
has its own particular advantages and disadvantages,
the choice of  which system to use has depended
on the application.  For experiments  involving whole-
body irradiation of animals, some kind of standardized
method of exposure on which most researchers can
agree is  needed. Although such ideas have been
proposed and discussed in the past, they are difficult
Figure 3-30.  VHF resonant cavity facility (Guy ef al. 1976)
                    Exposure Chamber
                  2.58m


a) Exposure of phantom scale model of man in a resonant cavity.
      b) Block diagram of resonant cavity driving system.

                    3-39

-------
Figure 3-31.  SRI multimodal cavity facility for primate irradiation (Heynick at al. 1977).
     a) Microwave cavity with door open, showing mode
     stirrer, iris, radiopaque windows, and waste-
     collection tray.

to implement because many exposure requirements
are conflicting, and,  therefore, no one  exposure
system can satisfy all of the various requirements.
However, some agreement on standardization should
make experiments more readily replicable and effects
more readily confirmed. Further efforts are needed to
this end.
During the past decade, there has been significant
improvement in the exposure facilities available for
biological experimentation.  Many of these facilities
have succeeded  in overcoming  the  deficiencies
inherent  in  many of  the older exposure devices.
Exposure systems specifically designed for chronic
irradiation studies must also be developed further,
particularly to reduce  their  cost without significant
sacrifice in performance.
Also  seriously needed  are  further studies that
compare  absorption and biological  end points  for
different  forms of exposure.  This problem can be
summarized as follows:  For the same experiment
performed in different exposure systems involving
plane-wave, complex,  or multi-path fields, are there
differences in biological outcomes even where the
same whole-body-averaged  SAR  is maintained
throughout?
                b) Dielectric cage.
3.3.2 Animal Holders
Any experimental irradiation of nonanesthetized
animals requires use of some  form of holder or
restrainer to keep the animal in a location of known
field strength or power density. For multiple animal
exposures, it is also  necessary to keep the animals
separated from each other. From a dosimetric point of
view, the optimal exposure condition is one in which
the animal  is constrained  in one orientation and
posture. This way,  both whole  body and  regional
dosimetry could  be  determined with satisfactory
precision. However, complete restraint of an animal
is considered by most biologists an excessive stress
and, therefore, unacceptable. Consequently, the
commonly accepted exposure environment represents
a reasonable compromise between these conflicting
demands: the  holder is designed to restrain the
animal, allow it to change  posture and orientation,
and also  allow for a  small  degree  of  lateral and
vertical movement.
3.3.2.1 Perturbation by Restrainers
One of the important criteria  in designing animal
holders is that they be constructed of materials that
                       3-40

-------
cause  the least perturbation  of  incident  fields
possible. What that means is that cages can be built
only from nonmetallic materials. Various types of
low-loss dielectric materials are available in rigid and
semirigid form, and have been the material of choice.
The most popular has been acrylic plastic (polymethyl
methacrylate), because of its optical transparency,
superior mechanical strength, and ease of machining.
Some experimental and analytical studies have been
performed to determine the  degree to which these
materials perturb the incident field (Weil 1974; Lin et
al. 1977). These studies showed that, under certain
conditions, these  materials can indeed cause
serious field  perturbations, so that the real power
density of the field to which the animal is exposed is
not known  precisely. Use of foamed polystyrene
instead of  acrylic is  a possible solution.  Foamed
polystyrene has a relative permittivity value close to
unity  and creates only minimal field perturbation.
However,  two disadvantages are associated with
polystyrene materials: They are optically opaque, so
that animals cannot be observed in their cages; and
they are not  strong and can be readily gnawed by
rodents. Coating the material with a substance to
which  the rodents are averse, such as quinine, has
been tried by a few workers (Catravas 1976) to reduce
gnawing. However, this practice adds an unwanted
additional insult to the experimental subjects.
Nonetheless, foamed polystyrene is useful in studies
involving short-term exposures.

A more recent  paper by Ho  (1978) questioned the
significance of the incident-field perturbations
created by  holders made of acrylic materials. In an
analytical study involving a two-dimensional  model
that contained  a cylinder  of  muscle-equivalent
material mounted concentrically inside  an acrylic
cylinder. Ho showed that  despite  a significant
perturbation of the incident field in the empty holder,
there was relatively little perturbation in the whole-
body-averaged SAR and the SAR distribution  inside
the muscle-equivalent cylinder when compared with
that which  existed when no holder was present. Ho
concluded that large perturbations of the field created
by  animal  holders  do not automatically create  a
correspondingly  large perturbation  in the animal's
absorption  rate because the  presence of a  mass of
lossy dielectric (i.e., the animal) tends greatly to damp
out the standing-wave patterns created by the acrylic
holder. Further experimental observations are
needed to confirm this conclusion.

As part of a behavioral investigation. Gage et al.
(1979) investigated the influence of two different
holders on  the whole-body SAR of rats and mice at
2450 MHz.  One holder was a  cuboidal container
made of foamed polystyrene; the other was an acrylic
cylinder. They concluded that there was a significant
alteration of SAR values for both rats and mice  due to
the presence of the acrylic cylinder when the animals
were oriented parallel to the  E field. This effect was
considerably  more  pronounced  in the  mouse,
probably because 2450 MHz is a frequency close to
resonant absorption for the E-parallel orientation of
the mouse.


3.3.2.2 Feeding and Watering During Exposure
Experiments involving chronic,  long-term RF
irradiation  of  animals  usually  require that some
arrangements  for food  and water be made during
irradiation.  Provision  of  food does not  present
difficulties if the standard form of dry food pellets fed
to laboratory animals is used,  since these pellets
absorb little RF energy.  However, providing drinking
water or liquid nutriment does present some
problems because  water  is a comparatively good
conductor  and, therefore, absorbs energy. Two
methods of supplying water have been used; each
has problems.
In the first, the water  reservoir  is placed  in the
exposure system with the animal. The  major
disadvantage here  is that, since the water supply is
located in a high-field-strength environment,  it
absorbs RF energy and perturbs surrounding fields.
These factors create  uncertainty in both dosimetric
and densitometric measurements. Furthermore, it is
conceivable that,  if the field levels  are intense
enough, the water temperture could rise until the
animal would refuse to drink.  Consequently, this
approach is not recommended.
The second method  involves placing the reservoir
outside the exposure system in a region of little or no
field strength, and supplying water to the animal via a
supply tube or pipe.This method is superior, but it too
has associated pitfalls which, without special
precautions, can introduce serious experimental
artifacts. As Guy and Chou (1976) have pointed out,
the water tube or pipe provides a conductive pathway
for microwave currents. When the animal attempts to
drink, a circuit is closed between the animal, which is
at high potential, and the water reservoir, which is at
zero or ground potential. RF currents then pass from
the animal  to  the watering system; because the
contact  point area is small and involves sensitive
tissue,  a  high current  density may exist in the
animal's tongue.  Since the animal will  find this
condition aversive, it will probably refuse to drink.
Therefore, the  RF  current flow through  the water
supply system must be interrupted. Three methods of
interruption have been devised: (a) When the animal
interrupts a beam of light using  a tongue-licking
operant, a small bolus  of  fluid is  injected into the
animal's mouth (King et al. 1970); (b) the animal
holder is equipped with a small water trough which is
continuously supplied through a water drip system;
and (c) an RF-choke assembly is placed between the
animal and the water reservoir to effectively decouple
the animal from the water supply  when  drinking
(Figure 3-32). In the choke assembly design (applying
                                                                      3-41

-------
Figure 3-32.  Water-supply system for exposure chamber
           (Guy and Chou 1976).
                                                                                            (3-13)
            Metal
            Walls
                               8 Ounce
                                Bottle
R, = 1.75cm
R, =  .875
R3=  .475
          U = 1/2 (Co E2 + ,UoH2) = UE + UH
      Exposure
      Chamber
      Wall
      (Screen)
  Contacting
   Fingers
              • Plastic
             Cage Wall
                            4 mm
only to enclosed systems), water is supplied through
copper tubing that forms the center conductor of a
shorted quarter-wave section of a triaxial air line. To
prevent the animal from coming into contact with the
copper tubing, a  short section of glass or acrylic
tubing is mounted at its tip. The  choke assembly
effectively provides a high impedance at the point of
animal contact and thus reduces the RF current in the
watering system to near zero.

3.3.3 Densitometr/c Instrumentation
Densitometry is the  measurement of RF-field
strength and is  usually expressed in units of
equivalent plane-wave  power denisty. Dosimetry is
the measurement of absorbed energy, or the rate of
energy absorption by some object in  an RF field. In
this section, the relationships between the  strength
or  intensity of an RF  field and a suitable and
measurable field variable are given. The question of
what constitutes a suitable variable  has been
addressed (Bowman  1970). For a linearly polarized
plane-wave EM field in free space—and, for practical
purposes, in air—the following relationships  exist
between the electric field, the magnetic field, the
power density, the  electric  energy density, and
the magnetic energy density:
                                                              (3-14)

                                                              (3-15)

                                                              (3-16)
                       = *„
                     H
           (3-12)
  where E = electric field, V/m
        H = magnetic field, A/m
       r7o = characteristic or  intrinsic impedance of
           free space, 377 Q
       /t/0 = permeability of free space, H/m
       c0 = permittivity of free space, F/m
       W = power density, W/m2
        U = total energy density, J/m3
       UE = electric energy density, J/m3
       UH = magnetic energy  density, J/m3

The quantities /u0 and c0 are properties of the medium
and are scalar constants for  free space. Equation 3-
1 2 therefore states that in free space the magnitudes
of E and H are related by a constant factor r)0. Equation
3-14 relates the power density W to individual field
components  E  and  H  for the basic plane wave.
Equation 3-15 is a general relationship for the total
energy  density U  of an RF field and defines the
individual electric- and magnetic-field energy densities
UE and UH.
The relationships between field variables for a plane
wave are simple, so that the measurement of any one
field parameter allows calculation of  the others.
Although power density is a variable of interest in
biological effects studies, it is particularly difficult, if
not impossible, to  measure  directly because it is a
time-averaged vector cross product that, in general, is
not expressed as simply as Equation 3-14. However,
it may be indirectly evaluated for plane waves by
measurement of the E  or H field,  as indicated by
Equation 3-14. This measurement, when coverted to
units of power  density  (W/m2) by Equation 3-14, is
referred  to  as  the "equivalent plane-wave power
density." Above 1 GHz, it is usually not necessary to
measure both the E and the H fields (Bowman 1 970),
whereas below ~ 300 MHz, the H field  becomes
increasingly  important,  and an  independent mea-
surement of  H  is necessary (Asian  1979; Lin et al.
1973).
However, power density is inappropriate in the
evaluation  of near fields of antennas or fields  near
reflecting surfaces because intense fields can exist in
these regions when the power density is low or even
zero. A plane wave normally incident on a conducting
surface can create standing waves with large E and H
values but very low power density. Since power
density is a measure of the power flow across a unit
area, the net power flow of  a plane wave normally
incident on a conducting surface and reflecting back
                       3-42

-------
is zero (or small in practical cases), and hence the
power density is zero. The E and H fields in that case
would be in a standing-wave pattern with amplitudes
twice those of the fields of the incident wave.

In the near field of sources or for complex RF fields,
the E and H fields are  not related by Equation 3-12;
however. Equation 3-15, which related the individual
E and H fields to their respective energy densities, is
always true. Therefore, measurements of both the E
and H fields are required to evaluate the total energy
density where Equation 3-12 does not  apply.

Bowman (1970), Swicord (1971), and others have
described the desirable  characteristics of  an
instrument for quantifying hazardous EM fields. The
instrument and  probe should (1) measure in terms of
energy density and respond only to the variable being
measured, (2) be small  to permit a  high degree of
spatial  resolution and thus  be useful in small
volumes, and (3) have an isotropic response and be
insensitive to field polarization. The probe should also
cause little scattering of the field and operate over a
broad frequency range. The instrument and probe
should be battery powered, lightweight, and rugged,
and should read either peak or average values and
have  a  dynamic range of at least  20 dB (one
hundredfold) without the need to change probes. The
readout instrument should be direct reading and free
from susceptibility to RF interference.
Many approaches have  been  used to  develop field
probes that meet these criteria. The most successful
and enduring designs have centered around the thin-
film thermocouple or bolometer sensor/antenna and
the crystal detector/antenna. The first type of sensor is
a  square-law-responding  device  that  senses  a
temperature change due to the  absorption of energy.
The second type uses  a semiconductor  diode that
produces a voltage or current related to the strength
of a field variable. The  diode can respond as either a
square-law detector or  a linear detector, depending
on  the  region  in which  it  is  operating. Some
commercially available electric-field probes are
shown in Figure 3-33.

3.3.3.1  Electric-Field  Probes
The  NBS EDM-2  electric  energy  density meter
(Belsher 1975) was developed for accurate measure-
ment of occupational exposure to EM fields from 10 to
500 MHz. This  probe  uses three short orthogonal
dipoles terminated by crystal rectifiers  to detect the
RF field, and  produces an isotropic response within
±1 dB. The DC signal appearing on the diodes is
conducted to the meter's circuits through RF-
transparent high-resistance lines. The meter displays
electric energy  density in units of microjoules per
cubic meter. Energy density was chosen as the field
variable to be displayed because of the ambiguity of
power density in the near field of radiating elements.
The meter has a dynamic range of 50dB in nine ranges
from 0.003  to  30 fjj/m3.  Although  the  probe's
detector diodes are used in both the square-law and
linear regions,  a square-law response is achieved
over  most of the range by a varistor network. The
meter is calibrated at specific frequencies within the
FCC ISM bands at 13.56, 27.12, and40.68 MHz. The
frequency response is flat within ±1.0 dB over the 10-
to 500-MHz operating range. One of the improvements
in the EDM-2 device over previous designs is reduced
temperature-induced  drift. The reduced  drift  is
achieved by loading of the dipole-detector diodes with
equivalent load diodes. This  technique  has reduced
the temperature response to < ±0.7 percent/°C in
the 15 to 35°C range. The field probe design provides
a  rapid response time to  permit observation of
the modulation of the field and the measurement of
peak values. The rise and fall times (10 to 90 percent
of the signal and 90 to 10 percent  of the signal,
respectively) are both < 0.6 ms on the most sensitive
range.
A commercial version of the NBS electric-field probe
is marketed by Holaday lndustries(Edina, Minn.). This
field-strength meter (Model  HI-3001) operates over
the frequency range 0.5 to 1000 MHz and is accurate
within ±1.0 dB  over this range. The meter has an
isotropic response within ±1.0 dB and has peak-hold
and read  circuitry to capture the highest  reading
observed.  The response time is 1.5 s. Measuring
range of the meter is 1 to 3160 V/m in six ranges on
an analog meter. An  audible  tone, with  a  pitch
proportional to the meter  reading, is  provided to
prevent operator overexposure in survey applications.

In 1980 the General Microwave Corporation (Farming-
dale, N.Y.) introduced the RAHAM  Model 4A
radiation-hazard  meter.  This  instrument  is  a
broadband  crystal-detector design with  a frequency
coverage of 200 kHz  to  26 GHz,  and a  display
calibrated  in units of equivalent plane-wave power
density that ranges from 0.001 to 20 mW/cm2. This
meter has an isotropic response, is battery powered,
and has a response time of 1.5 s.
For use in the near field, Rudge( 1970) constructed an
electric-field probe  using two short crossed-dipole
antennas.  This meter was designed to operate from
915  to 2450 MHz  in proximity to RF emitters. The
probe detected the electric field in two perpendicular
planes, and was calibrated in power density terms.
Pudge carefully  analyzed  sources of error in the
design of near-field probes.  His study included the
effects of field curvature and antenna backscatter,
particularly with regard to  multiple coupling and
impedance variation,  both  of which affect  the
calibration  of the field  probe. The major source of
scattering  was the probe's antenna, particularly
when the  lengths of the dipole wings were of the
order of one-fourth wavelength. The prototypal probe
achieved a dynamic range of 47 dB with a square-law
                                                                      3-43

-------
Figure 3-33. Samples of commercially available survey meters for measuring RF electric-field strength. From left to right the units
           shown are: The Narda Microwave Corporation Model 8315A meter with a Model 8321 isotropic probe: the General
           Microwave Corporation Model 481A meter with a Model 81 probe; and the Holaday Industries, Inc., Model HI-3001
           meter and probe.
response,  by addition  of fixed  values of  load
resistances at increasing power levels to keep the
detector diode operation in the square-law region.
Sensitivity was reported at 300 nW/cm2 with a probe
antenna of 8-mm overall dipole length.

Bassen et al. (1975) developed a miniature broadband
electric-field probe. Although the probe wasdesigned
for  free-space operation, the ultimate goal was to
develop an implantable isotropic probe for biological
effects research (Bassen et al.  1977a). With 3-mm
dipole  elements,  the  free-space probe  achieved  a
relatively flat frequency response over the range of
915 MHz to 10 GHz.  After some refinements were
incorporated (Bassen 1977; Bassen et al. 1977b), the
probe was tested over a frequency range of 200 MHz
to  12  GHz. The improved version had a  30-dB
dynamic range with a 20-/uW/cm2 sensitivity,
isotropic response to ±2 dB, a detection bandwidth of
2 kHz,  and a spatial resolution of ~ 3mm.
The requirement of small size was imposed  on the
implantable-probe design to ensure minimal alteration
of the field when the probe is used relatively close to
sources and dielectric boundaries. At the same time,
the constraint of small dipole size  has the further
advantage of extending the frequency response of the
probe into higher frequencies. With 3-mm dipoles,
theoretically, resonance should occur at 50 GHz. The
probe uses  dipole elements terminated with three
zero-bias Schottky diodes for a maximal square-law
response, arranged in an 'T'-beam configuration.
High-impedance leads are connected to the diodes
and  routed  along the probe's body to an optical
telemeter. A  fiber-optics cable then  routes the
digitally encoded field-strength  signal to a digital
receiver that is calibrated to display the equivalent
plane-wave  power density. A commercial version of
the miniature  field probe is marketed by Electronic
Instrumentation and Technology (Sterling, Va.).
An  electric-field sensor  (EPS) and  readout  system
capable of operating over the frequency range 10 kHz
to 200 MHz has been developed by Instruments for
Industry, Inc. (Farmingdale, N.Y.) and was described
by Ruggera (1976). The EFS-1, -2, and-3 are linearly
                       3-44

-------
polarized square-law electric field detectors with an
analog readout calibrated in volts per meter. The EFS-
1 is intended for CW fields, the EFS-2 for pulsed
fields, and the EFS-3 for both types of fields with the
addition of a peak-reading  provision. The dynamic
range is 1  to 300 V/m with antennas of different
lengths. An optical transmitter is used when remote
readout is required, which eliminates field perturba-
tions introduced by metallic  cable. An  isotropic
monitor has also been designed (RHM series) that
essentially encloses three EPS-series sensors in one
box. Readings from each sensor can be summed (by
the square root of the sum of the squares of the field
strengths) to measure the  total electric  field, or a
particular sensor can be read directly when field
polarization is a variable of interest.


Another design of an electric-field probe uses thin-
film thermocouple arrays acting as both the dipole
antenna and the detection element. The dipole
elements are  lossy; therefore, their temperature
increases proportionally with the square of  the
tangential electric field. The dipole elements can be
thought of as the series connection of a large number
of tiny thermocouple dipoles. This detection method
provides true  square-law response and  produces
signals proportional to the average energy density of
the electric field in the measured volume.
The Narda Microwave Corporation (Plainview, N.Y.)
produces a wide selection of RF-radiation monitoring
instruments, such  as the  Model 8606 broadband
isotropic radiation  monitor.  This instrument uses
three orthogonal, thin-film thermocouple sensors that
respond to the square of the electric field. Two probes
are available that have the combined dynamic range
of 0.02 to 100 mW/cm2 over the frequency range 0.3
to 26 GHz. The probes have isotropic response within
±0.5 dB except when the electric field is aligned with
the handle axis. The General Microwave RAHAM
Model 3, another electric-field-sensitive meter
with thin-film thermocouple sensors, performs
comparably to the  Narda 8606  system. The Narda
8100  series meters  were designed earlier (Asian
1970) for measuring the leakage of microwave ovens,
dryers, and medical equipment, which operate at 915
and 2450 MHz. The probe used with the 8100 series
meters is not isotropic but must be oriented with the
probe handle  perpendicular to the wavefront. Three
probes are used to  measure across a dynamic range
of 0.02 to 200 mW/cm2. The probes contain two
orthogonal, thin-film vacuum-evaporated thermo-
couple elements that function both as antennas and
detectors. When the probes are used at 915 MHz, an
adapter is attached to the end of the probes, so that
the effective length of the dipoles is increased. This
increase in  length compensates  for the reduced
efficiency of the antenna at 915 MHz.
3.3.3.2 Magnetic-Field Probes
As discussed previously,  measurement of only the
electric field for frequencies below — 300 MHz does
not fully characterize the hazard or heating potential
of an RF field, because the electric and  magnetic
fields are not related  by the free-space intrinsic
impedance at frequently encountered distances from
radiation sources.  In the mid-1970's, NIOSH
supported an NBS effort to develop a magnetic-field
probe that would operate in the 10-  to  300-MHz
band. Two probes were developed (Greene 1975a,b)
that consist of small  single-turn, balanced loop
antennas 10 ancj 3.16 cm in diameter. The dynamic
ranges of the two probes are 0.5 to 5.0 A/m and 5.0
to 50 A/m, respectively. The probes are capable of
measuring  the magnetic field within a few  centi-
meters of an RF-radiation source, in the near field.
The loop antennas are constructed with two short
gaps;  one gap is terminated with a silicon diode that
detects the  induced  RF voltage; the other gap is
terminated with a capacitor  to bypass RF currents
without short-circuiting the DC output. The DC signal
is then transmitted over NBS-developed conducting
plastic lines to an electrometer voltmeter. One of the
inherent  problems of  loop  antennas  is  that the
response is proportional to frequency, which makes
these antennas  particularly sensitive  to  errors if
harmonics  (integral multiples  of the fundamental
frequency)  are present in  the field  under  study.
Greene also  shows  that  both the  electric-dipole
response of the loop and the partial-resonance effect
increase with frequency. An experimental method is
provided to minimize the error due to electric-field
sensitivity, and plots are provided for the two probes
to indicate worst-case  measurement error vs.
frequency due to the electric-dipole response. The
self-resonant frequencies  of the  two loops  were
computed to be 280 and 760 MHz for  the 10- and
3.16-cm loops, respectively, and a correction  curve
was provided to estimate the increase in  response
due to this effect.

Asian (1976)  reported the  development  of a
magnetic-field-radiation monitor that has a frequency
response within ±1.0 dB over the frequency range of
10 to 200 MHz. Two probes, the Narda Corporation
Models 8631 and 8633, are used to provide  a full-
scale  measurement range of 0.2 to 100 mW/cm2.
(Sensitivity extends to 20/vW/cm2 on the Model 8631
probe.) These  probes have  an isotropic  response
(±0.5 dB) and accept all polarizations. Each  probe has
three  orthogonal coils; each coil has two turns. The
coils  are terminated with  thin-film thermocouple
elements. High-resistance lines are used to route the
DC  signal to a preamplifier in the handle of the
shielded probe, and the preamplifier output is con-
nected via shielded cable to an indicating  meter. To
achieve  a partial solution to  the  problems of
frequency dependence  of loop antennas,  the coils
                                                                      3-45

-------
were designed to be series resonant slightly below
the low-frequency  end of  the operating band. The
response of the coils increases at frequencies above
200  MHz, as the  electric-dipole  effect begins  to
govern the response.
Magnetic-field  probes of various designs  were
reviewed by Ruggera (1976); many were commer-
cially available and others were experimental.
Ruggera described electric-field-shielded loops
constructed with semirigid  coaxial transmission line
from 2 to 6 cm in diameter. Ruggera also described an
isotropic, electric-field-shielded, three-loop, ortho-
gonal, magnetic-field probe that has a cross-polariza-
tion rejectivity of ~ 20 dB  from 40 kHz to 50 MHz.
Signals from the probe are recovered with terminated,
coaxial cables. In addition, the author described a
three-axis, magnetic-field-measuring instrument
developed by Southwest  Research Institute. The
original  design was  plagued  by cable-pickup
problems, which limited the usable frequency to 50
MHz. Refinements by the FDA Bureau of Radiological
Health  (BRH) that incorporated the optical telemetry
systems (see Sec. 3.3.3.1) used on the BRH miniature
electric-field probe led to a flatter frequency response
over the extended frequency range of 150 kHz to 150
MHz.
In 1981 Holaday Industries  introduced a new
magnetic-field probe for their 3000 series  of field-
strength meters. The  Model STH-01  probe has a
frequency response of 5 to 300 MHz within ±2 dB.
The uniformity of response improves to ±1  dB over
the range of 10 to 200 MHz. When used with the Hl-
3001 meter (Sec. 3.3.3.1), full-scale sensitivity of the
probe is 0.1 A2/m2 in the low range and 1.0 A/m2 in
the high  range. The  sensor consists of three
orthogonal  loops terminated with detector diodes
that rectify the RF voltages  induced in the loops. The
sensors can  withstand an 800-percent overload
above the full-scale field strength without damage.
                       3-46

-------
3.4 Dosimetric Methods

James B. Kinn

Dosimetric methods are used to determine where
and  how much energy is absorbed by a biological
target.  Two areas  are  considered:  (1) whole-body
dosimetry,  dealing  with the  integrated or spatially
averaged SAR for the entire animal;  and (2) regional
dosimetry,  dealing  with the  SARs or internal field
strengths in a specific site of the biological target.

3.4.1 Whole-Body Dosimetry
Whole-body dosimetry  is based  on  either a power
difference  or a  calorimetric method. The  power-
difference  method  is  limited to closed exposure
systems  such as waveguides, TEM-mode trans-
mission cells, and coaxial air lines.  In this method,
power meters are attached to the exposure systems
by directional couplers to  measure the incident,
reflected, and transmitted power through a section of
the closed  system containing the biological target.
The power absorbed by the biological target is then
measured by subtracting the values  of the reflected
and transmitted power  from the  incident power. In
the case of exposure systems using live animals, the
values  of  those  three  measurements vary as the
animals move, because absorption  is a function of
the animal's orientation relative to field polarization.
What is needed is a data-collection system that can
integrate  the measured power difference and then
determine an average over a period  of time, usually
the exposure duration (Christman et at. 1974; Ho and
Edwards  1977b). The advantage of  this method is
that it provides on-line instantaneous measurements
of absorbed power. A disadvantage is that  its
optimum  application is  in single-animal  exposure;
simultaneous exposures of multiple animals in the
same closed  system would  result in average  dose
rate  values  for the group,  rather than  for  an
individual  animal. A second disadvantage  is  that,
under exposure  conditions  where  the  RF energy
absorption  rate is very low, it may be impossible to
measure  the power  absorption,  because the
measurement error in the power meter exceeds the
computed power difference.
Calorimetric methods use the thermalization of the
absorbed power as a measure of SAR. Under the tacit
assumption that all the RF energy  absorbed  by a
biological target is converted to heat,  the calorimetric
method measures the heat added to the target. The
approach is to irradiate an  animal  carcass with a
short exposure at high power to reduce heat loss
errors due to conduction and convection during the
exposure. The thermalized  energy added to the
carcass is measured with the calorimeter. Calories
per unit time per unit  mass  are then converted to
W/kg, the SAR for the high incident power density.
The SAR for any  other power density is then scaled
from this measured value in direct proportion to the
ratio of power densities. Three calorimetric methods
that have been used in whole body dosimetry are
gradient-layer calorimetry (Gandhi et al.  1979),
Dewar-flask calorimetry (Blackman and Black 1977),
and twin-well calorimetry (Hunt and Phillips 1972;
Kinn 1977). All three methods use a freshly killed
animal carcass that has been briefly exposed to high-
power RF fields and thus raised to a higher tempera-
ture.  It has been assumed that the dielectric
properties of the carcass do not differ significantly
from those of a living animal.
The gradient-layer calorimeter measures the rate at
which heat passes from the heated animal through
the walls  of the calorimeter to the room air. It is
accurate,  but it takes several hours to  make the
measurement.

The Dewar-flask  calorimeter method  uses the
difference between the average temperature of the
animal carcass before and after rapid heating with RF
radiation to determine the heating rate and thus the
dose rate.  To determine the average temperature of
the animal carcass, it is placed (with a coupling fluid,
usually water) into a Dewar flask, and the equilibrium
temperature of the mixture is measured. From the
theory of mixtures and a knowledge of the average
specific heat capacity of the  animal, the average
carcass temperature is  determined. The average
temperature of the heated carcass  is determined
similarly, and the SAR is computed. Although the
method is  simple by design, a disadvantage is that a
long time is required  to measure equilibrium
temperature. Furthermore, the average specific heat
capacity of the carcass must be known, and, because
the body's composition  is complex, this capacity
cannot be  precisely determined.
The twin-well calorimeter method  uses pairs  of
carcasses  of equal  mass. One carcass of a pair  is
heated with  RF radiation; the heated and unheated
carcasses are each placed  in  a well  of the
calorimeter, and the calorimeter is allowed to return
to equilibrium. The  twin-well calorimeter measures
the increment of heat introduced by the exposure.
The SAR  is  then  calculated  from the measured
caloric increase, carcass mass, and exposure time.
This method does  not  require  the specific heat
capacity of the carcass but requires 4 to 16 h to make
a  measurement. A  recent improvement of the
method (Kinn et al. 1984) has reduced the measure-
ment time to 1 5 to 30 min and increased  measure-
ment accuracy by control of the calorimeter with a
microprocessor system.  This  arrangement  is the
same  as  the standard  setup except  that the
unexposed carcass  is in contact with  a heating
element (Figure 3-34). The  microprocessor program
controls the rate at which heat  is applied to the
unexposed carcass. Using the twin-well calorimeter
as a "thermal" balance, the program controls the
                                                                    3-47

-------
heating rate so that equilibrium is established in as
short a time as possible. All of the power applied to
the heating element, and thus to the animal carcass,
is integrated and divided by the mass of the carcass to
determine theSAR. A 2-percent error due to heat lost
to  the calorimeter walls has been measured using
electrically heated water bottles as animal substitutes.


3.4.2 Regional Dosimetry
Regional dosimetry is performed to determine either
the energy absorption rate or the field strength at a
specific location within an animal or tissue preparation.
The average field  strength within the animal can be
measured directly or inferred from the SAR. Direct
measurement  of  field strength  is  made  with  an
implantable probe containing one or more miniature
diodes mounted on a dielectric substrate and attached
with high-resistance leads to a voltage-reading device
(Bassen et al. 1977b). The voltage across the diode is a
measure of the field strength at the diode, provided the
animal tissue material in which the probe is implanted
does not differ significantly in dieletric properties
from the material used in the calibration procedure
(Bassen  1977). The useful frequency  range over
which  such field strength measurements can be
made is 100 to 12,000 MHz. The reduced sensitivity
at  lower frequencies and the greatly reduced field
strength in biologicaI targets at the higher f requencies
delineate the useful  frequency range.  Indirect
measurement  of  internal field strength is derived
from the SAR value. The heating rate is proportional
to  the square of the electric-field strength.
Measurement  of  regional SAR can also be accom-
plished  by temperature probes or by thermographic
imaging.  Temperature probes are small (1-mm
diameter) and are  designed to be implanted in the site
of  interest without altering the EM field in that area
(Rozzell et al.  1974;  Bowman 1976; Christensen
1977; Wickersheim and Alves 1982). An animal or
preparation is  briefly exposed to RF fields of high
intensity, and the  temperature rise is measured. The
rate of temperature rise, together with the amount of
heat loss, is then used to compute the SAR at a
specific site in an absorber, such as a rat brain.

The technique described above is a special case of
heating and cooling curve analysis,  i.e., estimating
the initial rate when the rate of heating  is much
higher than the rate of cooling. The  analysis of the
complete heating  and/or cooling curve is especially
useful in cases where the RF energy absorption is of
the same order of magnitude as the cooling rate of the
animal  or preparation. In this procedure,  the
temperature  is  recorded continuously  during
exposure and/or after exposure until a steady state or
equilibrium  temperature is reached. Analysis of
either curve yields the cooling rate, which is directly
related to the SAR. The heating curve also  contains
Figure 3-34.  Microprocessor-controlled twin-well
           calorimeter.
the cooling rate  that  can be  determined by the
appropriate analysis; therefore, the  SAR can be
determined from either a heating or a cooling curve
(Allisera/.  1977).

Thermographic imaging also requires very high
power exposures. This method uses a thermographic
camera containing a  mirror scanning system, a
liquid-nitrogen-cooled detector, and a data-collection
system  usually interfaced to a  minicomputer (Guy
1971; Guy et al. 1968,1977; Kantor and Cetas 1977).
The investigator prepares specimens  by cutting a
frozen carcass into halves and supporting each half in
an RF-transparent medium such as urethane foam.
The surface of the cut plane is covered with either a
plastic film polyester "silk" screen material to hold
each half of the specimen in its foam support. The
"silk" screen material allows  electrical  contact
between the two halves when they are readjoined
during exposure. Unlike plastic film, which prevents
induced current from flowing across its surface, the
"silk" screen material allows current flow, and the
model may be placed in a complex field of unknown
polarization. The assembled specimen  is allowed to
equilibrate  to  room  temperature (25°C).  Thermo-
graphic pictures are taken of the animal specimens at
                       3-48

-------
the plane of the cut before and after exposure, and the
SAP is computed from the point-to-point temperature
change during exposure. The SAR  values are then
used to produce a contour plot of iso-SAR lines or
averaged values over a specific area.
Scaled-down or full size  physical models made with
materials having dielectric properties similar to those
of biological tissues are  also used to determine the
SAR distribution in a biological specimen,  such as
man and small animals. Animal models have been
used for the  investigator's convenience. Scaled
models of man (Guy et al. 1976; Gandhi et at. 1977)
have been used to simulate human exposures under
varying conditions. This procedure involves frequency
scaling, in which a combination of higher-frequency
radiation with a smaller-than-life-size model is used
to simulate  the  absorption characteristics of man
when exposed to  lower-frequency radiation. The
technique also requires suitable  scaling for the
dielectric parameters  of the tissue-equivalent
materials used in  the  model. The validity of the
models  used to date is open to  some question,
because the physical "equivalent" model is filled with
a homogeneous material whose complex dielectric
constant is the estimated average for the whole body;
a more realistic model would attempt to simulate the
inhomogeneous tissue structure  of the body. The
dielectric material normally used in human modeling
is gelled water whose conductivity has been adjusted
with sodium or  potassium chloride. The  model is
made in two halves, and the thermographic method
described above for animal specimens is used.
3.4.3 Unresolved Issues
Refinements are lacking for models of  the in-
homogeneous  tissue  structure of man and animals,
and the equivalency of these models to the actual
biological target has not  been validated. In simulta-
neous exposure of multiple animals, to what extent do
animal movements create uncertainty  in  the  SAR
estimates? How much target separation is required
before this uncertainty falls within acceptable limits?
As research  into RF-induced biological effects grows
more sophisticated and as the sites for many of the
reported effects—particularly those associated with
the nervous system—become better isolated and
defined, there will be an ever increasing need for
localized dosimetry  of  greatly improved spatial
resolution. Finally, although it is well understood that
RF energy is converted to heat in a biological target,
the question remains: Are there  transient "field-
specific"  effects not explicable by a  temperature
change?
                                                                      3-49

-------

-------
                                           Section 4
                 Effect of RF-Radiation Exposure on Body Temperature
4.1 Thermal Physiology

Christopher J. Gordon
4.1.1 Temperature Regulation
Almost all animal species, including vertebrates and
invertebrates, are capable of sensing and responding
to changes in environmental temperature. The ability
to maintain a constant body temperature independent
of ambient temperature, termed homeothermy,  is
restricted to humans and other mammals, as well as
most birds. Reptiles, amphibians, fish,  and inverte-
brates generally have a body temperature similar to
ambient temperature, and are thus termed poikilo-
therms (i.e.,  having changeable temperature). Many
poikilotherms use behavioral mechanisms (e.g., seek
shade or sunlight) to regulate their body temperature
against fluctuations in ambient temperature.

Homeotherms employ  an array of  physiological
mechanisms to control  body temperature. Thermal
stimulation of the  skin  or sites within the body,
especially the central nervous system (CNS), leads to
the activation of heat-dissipating or heat-producing/
conserving effectors such as evaporation, vasomotor
tone, metabolism, and behavior.

Since the absorption of  RF radiation can lead to an
increase in tissue temperature, the bioeffects from
RF-radiation exposure at these thermal levels may be
a principal  manifestation of  a  homeothermic
response to rising tissue temperature rather than a
direct RF-radiation effect (i.e.,  athermal effect) in  a
biological system. Thus, it is  important to  have  a
detailed  assessment of the  characteristics of
thermoregulation in animals and man. This section is
divided into two major parts. First, the structure and
physiology of the thermoregulatory system of
humans and commonly used experimental mammals
are discussed in enough detail to a How the reader to
interpret  the second division  of the  section,  the
specific effects of RF radiation  on the activity of the
thermoregulatory system. The major conclusions of
this section are:
 • Thermoregulatory  effectors such as vasomotor
   tone, metabolism, evaporative  water  loss,  and
   behavior  are  activated during  exposure  to RF
   radiation.
 • Many effectors are activated in the absence of any
   measurable  change in deep-body temperature
   during RF-radiation exposure.
 • The  SAR required to increase activity of a
   thermoregulatory effector or raise body tempera-
   ture  generally decreases  with increasing body
   mass of the exposed species.
 • Thermoregulatory effectors employed by home-
   otherms during RF-radiation exposure at levels
   that  produce  heat stress  are  similar to the
   response during exposure to high ambient
   temperature.

4.1.1.1  Effect of Temperature on Biological
Systems
Below the point of protein denaturation at about 42 to
45°C, temperature has a direct effect on the rate of
biochemical reactions and thus affects  the rate of
physiologic processes. The effect of temperature on
biological reactions is described by the Qioparameter,
a dimensionless number equal to the change in the
reaction rate for a 10°C change in temperature:
              Qio=-
                      10/(T2-T,)
                   Ri
or log Qio = (10/T2 - Ti) (log R2 - log Ri)
where Ri = reaction  rate at Ti
      R2 = reaction rate at T2
       T = temperature, °C

Thus, a Qio of 2 means that the reaction rate doubles
with a 10°C increase  in  temperature, a  Qio of 3
means a tripling of the  rate,  and so forth. For
biological reactions, the  Qio generally ranges
between 2 and 3. In many  biological systems the Qio
is temperature dependent; thus, Ti and T2 must be
specified  in the calculation of
The  activity of poikilotherms is dictated by the
prevailing  environmental temperature.  At low
temperatures enzymatic activity is  decreased,
causing reduced muscle. activity, active transport, etc.
Hence, poikilotherms generally have a narrow range
of ambient temperature (~ 20 to 40°C) where normal
life functions can take place. On the other hand, many
homeotherms,  by  maintaining a relatively constant
body temperature, can function normally at ambient
                                              4-1

-------
temperatures far below 20°C and occasionally above
40°C.  Of  course,  to maintain  a constant body
temperature  at  extremely high  and low ambient
temperatures requires additional  metabolic energy.
These metabolic demands can be attenuated through
activation of behavioral  mechanisms (e.g., for
humans, by the addition or removal of clothing at low
and high temperatures, respectively).
4.1.1.2 Heat Balance
For body temperature to  remain constant, heat loss
must equal heat gain. If heat gain  exceeds heat loss,
body temperature increases, and if heat loss exceeds
heat gain, body temperature decreases.
Heat exchange between  the body and the environ-
ment takes place by the principal mechanisms of
conduction (including convection), radiation, and
evaporation. Normally, each of these represents heat
loss  from  the  body.  However, if air  temperature
exceeds the surface temperature  of the body, there
is a net exchange of heat into the body by conduction.
Infrared radiation from external sources may also
result in a net transfer of heat from the environment
to the body.
The variables for heat loss can be incorporated into a
simple equation, patterned after the first  law of
thermodynamics,  which relates metabolic  heat
production  to  whole-body heat loss (Bligh and
Johnson 1973):
            M=±K±C±R±E±S
where M = metabolic heat  production (always positive)
       K = conductive heat transfer (+ for loss)
       C=convective heat transfer (+ for loss)
       R = radiative heat transfer (+ for loss)
       E = evaporative heat transfer (+ for loss; gene-
          rally always positive)
       S = heat storage by body (+ for increase)
M may include  heat  produced in the body during
work, as well  as heat produced by absorption of RF
radiation.  Dimensions of each term  in  the  heat
balance equation are usually normalized with respect
to heat transfer per unit surface area per unit time or
heat transfer per unit body mass per unit time (e.g.,
W/m2 or W/kg,  respectively).
The three avenues of heat  exchange depend on a
number of external factors, the most important of
which is temperature. As ambient temperature (Ta)
increases, heat  loss by  radiation, convection,  and
conduction decreases, leaving evaporative heat loss
as the primary  avenue for  heat  loss. When  body
temperature is constant, S =  0, and M=K+C + R + E.

4.1.1.3 Autonomic and Behavioral Mechanisms
of Temperature Regulation
The neural regulation of body temperature is depicted
in a simple scheme in Figure 4-1. Thermal receptors
located on  the  periphery and deep  in  the  body
transduce temperature into nerve impulses which
are relayed to the anterior hypothalamus and preoptic
area (POAH),  the  primary site in  the CNS for the
integration of  temperature information. A relatively
high percentage of the POAH neurons are sensitive to
changes in temperature  (e.g., POAH and/or  skin
temperature).  Activity from the peripheral and deep-
Figure 4-1.  Simple neural model of thermoregulation in a mammal for predicting the motor responses to short-term (i.e., hourly)
          changes in ambient and/or body temperature. Activation of the warm-sensitive pathway leads to an increase in heat
          dissipatory mechanisms, whereas activation of the cold-sensitive pathway leads to an increase in heat-generating/
          retaining mechanisms. This information is taken from several models proposed by Hammel (1968) and Hensel (1973).
              Trigeminal Nucleus
                 Spinal Cord       Thalamus
  . Warm *•
     Peripheral
 Cutaneous Receptors
  Deep Body
Warm and Cold
  Receptors y

          Sensory
          Inputs
                                               Preoptic/Anterior Hypothalamus
                                              ', CNS Cold-Sensitive Neuron  ,'
                                Motor
                               Outputs

                                  I
                         Heat Dissipation
                         • Peripheral Vasodilation
                         • Panting/Sweating
                         • Behavioral Selection of a
                          Cooler Environment
                        Heat Generation/Retention
                        • Peripheral Vasoconstriction
                        • Shivering Thermogenesis
                        • Nonshivering Thermogenesis
                        • Behavioral Selection of a
                          Warmer Environment
                        4-2

-------
 body thermal receptors is integrated within the POAH
 and other  CNS  sites.  Electrophysiological studies
 have shown that neurons in the POAH are approxi-
 mately 10 times more sensitive to changes in POAH
 temperature than to changes in Ta (Reaves 1977).
 Specific areas in  the  CNS generate appropriate
 signals to drive thermoregulatory effectors that either
 raise or lower heat loss or increase heat production to
 maintain core and skin temperatures at their
 regulated (i.e., set point) levels.
 There is a complex interaction of systems in the body
 that are, in some part, responsible for the regulation
 of body temperature. Werner  (1980) developed a
 schematic of the major components of the thermore-
 gulatory system that  is useful in explaining, in
 general  terms, the  physiological and  behavioral
 mechanisms of temperature regulation (Figure 4-2A).
 The complex systems portrayed in Figure4-2A can be
 grouped into four principal parts: sensory system,
 control system, effector system, and passive system
 (Figure 4-2B). Simply stated, heat exchange between
 the  environment and the  passive system causes
 changes in temperature that are detected by  the
 sensory system and are then integrated in  the
 controller system for generating appropriate motor
 signals to the effector system, which then counteracts
 the  influence of environmental  heat exchange.
 However, each system is relatively complex; Figure 4-
 2A shows  that the temperature regulatory system
 operates through  the utilization of a variety of
 physiological systems to achieve a regulated body
 temperature.
 The sensory and control systems are very important
 but  are  only briefly described here (see above),
 because few studies on the bioeffects of RF radiation
 have been directed  toward these components of
 thermoregulation. For review of sensory studies in
 humans, see Sec. 5.6.5. The main thrust of work  has
 leaned toward effects on the effector and passive
 systems; thus, the basic characteristics of effector
 control  and regulation of the passive  system, body
 and skin temperatures, will receive primary attention.

 4.1.1.4 Metabolism
 The  effects of changing  Ta on tissue blood  flow,
 cardiac  output, and  evaporation  impart additional
 metabolic demands  on  a  homeothermic  species,
 thereby  increasing  its basal metabolic rate (BMR).
 Because measurement of BMR  in experimental
 animals is generally not feasible, resting metabolic
 rate is  a parameter commonly  used  to assess
 thermoregulatory function of homeotherms.
The resting metabolic rate of a homeotherm normally
 exhibits three major phases as  a function  of Ta  (cf.
 Figures 4-6, 4-7, and 4-10). There is typically a range
of Ta's where metabolism is at a minimal, stable level
(close or equal to the BMR). This range of T.'s is
 identified as the thermoneutral zone (TNZ). In the TNZ
the body temperature of a homeotherm at rest can be
kept constant by controlling the amount of passive
heat loss through adjustments in thermal conduc-
tance. When Ta increases above the TNZ, the rates of
metabolic heat production and evaporative heat loss
increase because excess body heat must be dissi-
pated actively rather than passively (e.g., sweating,
increase in ventilatory  rate). Metabolism also  in-
creases because of the "Qio effect" of rising temper-
ature on metabolism (Sec. 4.1.1.1). The  Ta where
metabolism begins to  increase  with rising T. is
defined as the upper critical temperature (UCT). As Ta
decreases below the TNZ, metabolism increases as a
result of  activation of thermogenic physiological
processes to maintain a  constant body temperature
(e.g., shivering). The Ta at the low end  of the TNZ
where metabolism increases is defined as the lower
critical temperature (LCT).

Ideally,  at  all Ta's equal to and below  the  LCT, a
homeotherm has reduced heat loss to the minimum
through vasoconstriction of blood vessels in the skin.
Therefore, whole-body thermal conductance, which
relates to the ease at which heat is lost from the body,
is as low as physically possible. AsTa increases above
the LCT, more blood is  shunted to  the peripheral
vessels, and whole-body  thermal conductance
increases.
Minimal whole-body thermal conductance is rather
simple to calculate and has been  reported for a
multitude of avian and mammalian  species with a
body  mass less than 10 kg (Aschoff 1981). In  an
idealistic situation, decreasing Ta below the LCT is
associated with a linear  increase in  metabolic rate.
Under these conditions,  metabolic rate can  be
calculated with a linear  equation derived from the
principles of Newton's law of cooling which relates
the heat loss of a  warm  object placed in  a cold
environment (Kleiber 1972):
                MR = C(Tb - Ta)
where MR = metabolic rate, W/kg
        C=whole-body thermal conductance,
           W/kg/°C
       Tb = body temperature, °C
       Ta = ambient temperature,  °C
Solving  for C when  Ta is below the  homeotherm's
LCT,
                   r=
                      Tb-Ta

This  equation  is commonly  used  in thermal
physiological studies of relatively small homeotherms.
Above the LCT, thermal conductance begins  to
increase as the animal's metabolic rate stabilizes and
Ta approaches body temperature. Above the UCT,
thermal conductance takes a sharp increase because
of the combined increase in metabolism, along with a
                                                                      4-3

-------
Figure 4-2A.  Current view of the principal systems employed in the regulation of body temperature (from Werner 1980).
               Receptors
                                                   Controller
                                          Effectors
Behavior
(Nutrition, Action,
Clothing, Change
of Environment
/
ferent i
\
                                                                Environmental
                                                                Temperature
                             I         Velocity                                i
               Receptors
Passive System
                                                                                         Effectors
                          4-4

-------
Figure 4-2B. Overall relationship of the sensory, controller, effector, and passive system components of the thermoregulatory
           system. Note that the overall structure of Figure 4-2A is retained (modified from Werner 1980).
             Receptors
             Peripheral
            and Deep-Body
             Warm and
            Cold Sensors
      Controllers
Integration of Temperature
   Information in CMS
    Passive System
Skin and Core Temperature
    Effectors
• Behavior,
• Vasomotor Tone,
• Metabolism,
• Evaporative
  Heat Loss
                                          Environmental
                                            Heat Flow
further  reduction of the ambient-body-temperature
gradient.

There is an  inverse correlation between minimal
thermal conductance and body mass (Figure 4-3). In
general,  the  thermal conductance  of  a  10,000-g
homeotherm  is 10 times less than that of a 100-g
species (Scholander et al. 1950). Simply stated, for a
given decrease in Ta, to maintain a normal body
temperature a small homeotherm must elevate heat
production (W/kg) much higher compared to a larger
homeotherm. The relationship also implies that at a
given Ta the rate of passive heat loss (W/kg) increases
with decreasing body mass.

In studies on  the thermal physiology of humans and
other primates,  thermal  conductance is often
calculated as  follows:

                   K=MR-E
                      Tre - TSk

  where K = tissue thermal conductance, W/mV°C
      MR = metabolic rate, W/m2
        E = respiratory evaporative heat loss, W/m2
      _Tr=rectal temperature, °C
      T»k =mean skin temperature,  °C
This value of conductance is commonly called tissue
thermal  conductance. The  numerator includes all
heat produced from metabolism minus the heat lost
through  respiration. The lower the gradient between
rectal and mean skin temperature, the  greater the
peripheral blood flow and the  higher the value  of
tissue thermal conductance. For example,  total skin
blood flow in humans ranges from 150to 200ml/min
in a cool environment, to as high as 2000ml/min in a
hot environment (Bullard 1971).
             4.1.1.5 Cardiovascular and Vasomotor
             Responses
             The thermoregulatory system  is a unique control
             system  in that it has  no specific  organ or tissue
             devoted solely to thermoregulation. (Sweat glands
             may be an exception.) The cardiovascular system is a
             primary example of a thermoregulatory effector
             which  is  perhaps more  critical to other  bodily
             functions (e.g., tissue perfusion). Heat transfer from
             the deep body  to  the skin  and thence to the
             environment would be nil if  it  were  not for the
             enhanced convective transfer of heat via blood flow
             by the cardiovascular system.

             Blood flow to  the skin is  a crucial mechanism of
             thermoregulation. In humans the rate of blood flow to
             the fingers ranges from 0.5 to 1.0 ml/min/100 ml of
             tissue during maximal vasoconstriction,  to 80 to  90
             ml/min/100 ml of tissue during maximal vasodilation
             (Burton  1939). This  tremendous adjustment  in
             peripheral blood  flow  allows for a five- or sixfold
             change in tissue thermal conductance (Hardy et  al.
             1941). Blood flow from the deep body  to the skin
             allows for dissipation of internal heat loads accrued
             during exercise. In hot environments, raising the skin
             blood flow increases skin temperature, which
             reduces  the temperature gradient between the air
             and skin and so lessens overall heat gain from the
             environment. In extremely cold conditions, blood flow
             to  the skin prevents cold-induced  tissue  damage
             (such as frostbite).

             Decreasing Ta  below the thermoneutral zone (see
             Sec. 4.1.1.6) also results in an elevation in heart rate
             and cardiac output, which is a response to the need
             for more oxygen in skeletal muscle and other tissues
             that are metabolically more active at lower Ta's. When
                                                                       4-5

-------
Figure 4-3.  Effect of body mass on whole-body thermal conductance of various mammals (data modified from Aschoff 1981).
       1.0
  u
  o
   0>
   CD
   S
   O
   u
   fa


   I
       0.1
      0.01
          «™    I
i inn
         I   i   i  i 11 ill      r  i -i  mm   -
                                                                                            c_
                                                                    LOG, = LOG 0.18-0.52 LOG,
         0.01
     0.1
  1.0

Body Mass, kg
10
100
Ta increases above  the thermoneutral zone the
increase in metabolism, along with the shunting of
more blood to the skin, increases the requirement for
a greater cardiac output.
The rat is often used in a variety of thermophysiologi-
cal studies. The tail of the rat is the principal site for
the control of nonevaporative heat exchange. At Ta's
between 27 to 30°C  there is an abrupt increase in
tail blood flow (Figure 4-4A). Below these Ta's blood
flow to the tail is minimal. For comparison, the blood
flow versus skin temperature in the hand, foot, and
forearm of a human is shown in  Figure 4-4B. The
overall response pattern is similar for the two species,
but the flow rates differ substantially, with the rat
having a  much higher rate of blood flow at given
temperatures above 27°C.

4.1.1.6 Evaporation
Evaporation of water from the skin and respiratory
tract is a major avenue of heat loss in homeotherms at
high TB'S. Approximately 2426 joules (580 calories) of
heat are lost from the body in the evaporation of 1 g of
water. Insensible water  loss, the water  lost by
diffusion  through the skin  and expired  during
ventilation but excluding sweating,  occurs at all
times,  provided the relative humidity is below 100
percent. In the rat, a nonpanting  and nonsweating
homeotherm, up to 50 percent of evaporative water
loss occurs through insensible water loss through the
skin (Tennent 1946).
                          Typically, as Ta increases above the upper critical
                          temperature (approximately 30 to 32°C for many
                          species), homeotherms undergo an active increase in
                          evaporative heat loss (EHL). Below this T. EHL is
                          relatively stable and amounts to roughly 25 percent of
                          the total heat loss. As Tfl increases the temperature
                          gradient between the skin and air decreases, which
                          reduces the rate of the heat loss by conductance,
                          convection, and  radiation. Thus,  EHL is a crucial
                          avenue of heat dissipation in homeotherms at
                          relatively high Ta's above the thermoneutral zone.
                          The major avenues of active evaporative water loss
                          are sweating and panting. Humans and many other
                          primates sweat to dissipate heat by evaporation (Sec.
                          4.1.1.9). The dog,  cat, and rabbit  have  few sweat
                          glands and dissipate heat evaporatively by panting.
                          Rodents neither pant nor sweat when heat stressed
                          but increase their ventilatory rate, and subsequently,
                          increase pulmocutaneous water loss. Rodents also
                          apply saliva to their fur to enhance evaporative heat
                          loss. Denervating saliva glands in rats reduces their
                          tolerance to heat exposure (Strieker and Hainsworth
                          1971).

                          4.1.1.7 Thermoregulatory Behavior

                          Heretofore, the principal autonomic thermoregulatory
                          responses, metabolism (heat generation), blood flow,
                          and evaporation, have been discussed. Thermoregul-
                          atory behavior is another important effector which
                          animals and  humans use  to regulate the  body
                       4-6

-------
 Figure 4-4. Blood flow vs. skin temperature. (A) Tail blood flow vs. tail skin temperature in three different rats. Notice the abrupt
           rise in blood flow at temperatures of 27 to 30°C (data from Rand ttal. 1965). (B) Blood flow in a human hand. foot, and
           forearm vs. temperature of the water in which the limb is immersed (skin temperature equals water temperature) (data
           from Thauer 1963).
    50
   40
E
8
I  30
   20
   10
                                                    30
                                                    25
                                                    20
                                                    15
                                                    10
                                  •  Hand
                                  •  Foot
                                  «  Forearm
     15
               20
25
                                    30
                                              35

                                             Temperature,
                                 10
20
                                                    30
                   40
50
 temperature in response to changes inTa. Changes in
 behavior require less metabolic energy compared to
 changes in  activity of autonomic effectors; thus,
 behavioral responses normally take precedence over
 autonomic responses (Adair 1976; Gordon 1983a). For
 example, when  placed in a temperature  gradient
 (Ogilvie and Stinson  1966) adult mice will select a
 temperature of approximately 31 °C (Figure 4-5). This
 is a thermoneutral Ta where the animals maintain
 normal body temperature without  increasing
 metabolic rate or evaporative heat loss. Very young
 mice  select  notably  warmer  temperatures  in a
 gradient because their metabolic capacity is not fully
 developed compared to the adult. A wide variety of
 species can be trained to bar-press for heat or cold
 reinforcements and thereby select a comfortable Ta
 (Satinoff and Hendersen 1977).

 All  organisms display subtle thermoregulatory
 behaviors  in the absence of complex experimental
 apparatus. For example, in the cold, groups of rodents
 huddle together  to minimize heat loss. Mice build
 dense, thick nests in the cold and thin-walled nests in
                          the heat.  Solitary  animals adjust their posture to
                          control  heat  loss. For  example, in  a very  hot
                          environment  the  rabbit  extends its  limbs in a
                          sprawled position to maximize the surface-area-to-
                          volume  ratio,  whereas in a cold environment it
                          withdraws its  limbs under the body and assumes a
                          ball shape to minimize the rate of passive heat loss.
                          Humans vary clothing in accordance with Ta.

                          4.1.1.8 Autonomic Thermoregulatory Responses
                          vs. Ambient Temperature

                          Because the absorption of RF radiation may activate
                          thermoregulatory mechanisms similar to those that
                          occur during an increase in Ta, it is of importance to
                          show in this section the effect of Ta on the activity of
                          autonomic effectors. In Figures 4-6 and 4-7 the rectal
                          temperature, evaporative heat loss,  tissue thermal
                          conductance,  and metabolic rate are plotted as
                          functions of Ta for the dog and rhesus monkey. (For
                          data  on humans, see Sec.  4.1.1.9). These thermo-
                          regulatory profiles were selected from the literature
                          because they  demonstrate not  only  basic thermo-
                                                                        4-7

-------
Figure 4-5.  Effect of age on the preferred ambient temperature
           of mice (data from Ogilvie and Stinson 1966).
   38.0
o
°.  36.0
   34.0
a

   32.0
M
   30.0
   28.0
      10    20    30   40   50   60   70    80   90

                         Age, days
Figure 4-6.  Effect of T. on evaporation (Etoi), conductance (K)
           rate  (M),  and rectal temperature (!„) of dogs
           acclimatized to summer and winter (data from
           Sugano 1981.
\

5
 E
 x
 E
 x
 o
 o
 S
      60


      50


      40


      30

       6
       5
       4

       3

       2
       1

       0

      30
      20
      10

       0

      39


      38
                          Dog
     37
                                       • Summer
                                       O Winter
          18   20   22   24   26    28

                  Ambient Temperature, °C

                         4-8
                                         30   32
                                                     Figure 4-7.  Effect of T. on evaporation (E10i). conductance (K)
                                                               metabolic rate (M). and rectal temperature (T,.) of
                                                               the rhesus monkey  under  unanesthetized  and
                                                               ketamine-anesthetized conditions (data from
                                                               Hunter «f«/. 1981).
                                                     E
                                                     x
                                                     o
                                                     o

                                                     £
                                                     X
                                                     X

                                                      I
                                                         80
                                                                         Rhesus Monkey
    70-

    60-

    50-

    40-

    30

    50-

    40-

    30-

    20-

    10-

     0-

    50-

    40-

    30

ui   20-

    10-

     0-

    40


y   39-
O

^   38


    37-
                                                         36
                                                                                     • Unanesthetized
                                                                                     O Ketamine
      14 16 18  20  22 24 26  28  30 32 34 36  38  40

                  Ambient Temperature, °C

regulatory effector activity as a function of T,, but also
the effect of anesthesia and seasonal acclimatization
on thermal responsiveness.
In the dog (Figure 4-6), seasonal acclimatization has a
tremendous effect on the thermoregulatory profile
(Sugano 1981). For example, summer-acclimatized
dogs (mean Ta = 25.3°C) have a UCTof 29°C, whereas
winter-acclimatized dogs (mean Ta = 3.1°C) have a
UCT of 22°C. At warm T»'s the winter-acclimatized
dogs  have  a  higher  rate  of metabolism  and
evaporation. These differences in thermoregulatory
profiles in the dog may  prove to be helpful  in
understanding the metabolic and  evaporative
response of other mammals to RF-radiation exposure.

-------
The effect of ketamine anesthesia on the thermoreg-
ulatory response  in  rhesus  monkeys  is shown in
Figure 4-7. These data are presented because many
RF-bioeffects studies are done with  anesthetized
animals. The anesthetized monkeys have a reduced
thermal conductance and depression in heat loss at
the high Ta (38°C), which contributes to an elevated
rectal temperature compared with unanesthetized
animals. Note the linear relation between Ta and Tre in
the anesthetized monkey. This and other studies on
the effects of anesthesia  on thermoregulation
indicate a general depression  in the ability of the
animal to  defend  its body temperature against heat
stress or cold stress.  Anesthetized animals exposed
to RF radiation are likely to also be more susceptible to
RF heating depending on the prevailing Ta(Sec. 4.1 .3).

4.1 .1 .9 Thermal Physiology of Humans
In view of the paucity of studies on human physiology
during RF-radiation  exposure, this section  deals
primarily with the  responses of laboratory mammals.
One must  be cautious in relating measured biological
effects in experimental mammals to man because the
thermal physiology of humans  and other mammals
differs substantially in certain aspects.
The human body has an inner core with a temperature
of approximately  37°C and an outer  shell with a
variable  temperature (Figure 4-8).  The normal
strategy of the human thermoregulatory system is to
have a well-controlled inner-core temperature at the
expense of a fluctuating shell temperature. Circadian
rhythms cause core  temperature to oscillate daily
with an overall amplitude of approximately ±  1°C
(Hardy and Bard 1 974).
Generally,  about  two-thirds of the body mass is
assumed to be at the core temperature and one-third
is at the skin temperature. This assumption may be in
error because of shifts in the relative size of the core
and peripheral shell. The average body temperature
(Tb) can be calculated as:
              Tb  = 0.33Tsk + 0.67 T,e

where TSk  is the average skin temperature collected
over a variety of sites  on the surface of the body, and
Tre is rectal temperature.
The average body temperature is a useful variable
because it allows  one to calculate the rate of heat
storage (S) in watts:
                                                  Figure 4-8.  Temperature distributions in the human body at
                                                            T.'s of 20 and 35°C (data from Aschoff and Wever
                                                            1958).
                S =
                  = c
                             m
where Tb
         andTb2 = initial  and final average-body
                 temperatures, respectively, during
                 the specified time (t, seconds)
              c- specific heat of the body tissues
                 (-3.48 kJ/kg/°C)
              m = mass (kg)
                                                           20°C
                                                                                   35°C
The calculation of rate of heat storage from average
body temperature is useful because the dimensions
for the rate of heat storage are the same as that used
in RF-radiation dosimetry (W or W/kg).  For example,
humans placed in a very hot environment (Ta = 55°C,
vapor pressure =15 torr) for 2 h undergo a more than
1 °C rise in rectal temperature and an increase in heat
storage from 2.1 to 6.5kJ/kg, or a heat storage rate of
0.6 W/kg. At a slightly cooler but  more humid
environment (Ta = 48°C, vapor pressure = 34 torr) the
change in heat storage over 80  min was 2.0 to 6.5
kJ/kg, or a heat storage rate ,of 0.94 W/kg. These
heat  storage  rates  were  associated with large
increases in plasma cortisol and  the feeling of
discomfort (Follenius et a/. 1982).
A major difference between humans  and most
experimental mammals is the ability to secrete sweat
on the skin. Under resting, thermoneutral conditions
there is a certain amount of  insensible EWL through
the skin, which  accounts for about 8 W  of heat.
Approximately 25 percent of the total metabolic heat
at rest is lost through evaporation. Raising Ta causes
a tremendous increase in the amount of heat lost by
                                                                       4-9

-------
evaporation. The partitioning of heat loss at various
Ta's in humans is listed in Table 4-1.

Sweating in man and certain other homeotherms may
be evoked by heating the skin or thermally sensitive
sites  in  the nervous system (Figures 4-9A.B). It is
interesting to compare the relatively high threshold of
sweating at low vs. high skin temperatures (Figure 4-
9B). This comparison may be relevant to some deep

Figure 4-9.  Effect of skin temperature on sweating: (A) Effect
           of increasing skin temperature on sweating from
           the forearms  of a  human (data from Elizondo
           1973); (B) effect of skin temperature on the
           threshold internal cranial temperature for activa-
           tion of sweating in humans (data from Benzinger
           1969).
Table 4-1.   Partitioning of Heat Loss in Humans as a Function
           of Ambient Temperature under Still Air Conditions
 c
 E

 E
 u
 o>
ID


W
      0.15-
    400
    300-
    200 H
    100-
          Skin Temperatures. °C
         39, 38, 37, 36, 35. 34, 33
                     33  32  31 30  29
                    —• Skin Temperatures, °C
Ambient
Temperature
(°C)
25
30
35
Radiation. IR
(%)
67
41
4
Convection
(%)
10
33
6
Evaporation
(%)
23
26
90
Data derived from Hardy era/. (1941 land compiled by Folk (1974)

penetrating frequencies  of  RF radiation  that  heat
predominantly the inner tissues and organs without
first substantially warming the skin. Such responses
may affect the organism's ability to respond to RF
heating.

The  effect of Ta on the steady-state rate of  heat
production, heat loss, evaporative water loss, thermal
conductance,  and  rectal and skin temperature of
nude adult men and women is shown in Figure 4-10.
Sweating is activated at Ta's of 30 to 32°C. For both
sexes, thermal conductance increases at a Ta of 28°C,
which is indicative of an increase in skin blood flow.
The thermoneutral zone  is quite narrow:  the upper
and lower critical temperatures are approximately 32
and 30°C, respectively. At warm Ta's women have a
    *
Figure 4-10.  Effect of calorimeter temperature (x-axit) on rectal
           and skin temperature, heat loss, heat production,
           evaporation, and thermal conductance in men
           and women (data from Hardy at al. 1941).
                                                       W/m2
                                                      g/m2 • h
                                                      W/m2 • °C
          36.4      36.8     37.2     37.6
            Cranial Internal Temperature. °C
                24   26    28   30    32    34

                    Calorimeter Temperature, °C
                         4-10

-------
 steeper thermal conductance response, which
 perhaps indicates increased ability to shunt blood to
 the skin compared to males.

 It is important to understand the thermoregulatory
 responses of humans during exercise and fever, since
 these represent two cases where internal temperature
 can increase  without  rising  ambient  or skin
 temperature and, in this respect, may appear similar
 to that of RF heating at certain frequencies. However,
 as will be demonstrated,  thermoregulation during
 exercise and fever may differ  substantially from
 thermoregulation during RF exposure.

 Humans are well adapted to dissipate excess heat
 loads accrued from exercise. Examples of thermoreg-
 ulatory responses during two levels of exercise on a
 bicycle ergometer are shown in Figure 4-11. A work
 load of 40 W causes an increase in metabolism from
 1.5 to 4.2 W/kg and  a concomitant 0.34°C rise in
 tympanic temperature. The temperature of  the
 tympanic membrane in the ear is considered to be a
 good indicator of the  internal temperature of the
 head. A 90-W work load increased metabolism to 7.0
 W/kg  and raised tympanic temperature by 0.96°C.

 The  rise in metabolism from pedaling the ergometer
 and  the heat load normalized to time and mass from
 RF-radiation exposure  have the same  dimension
 (W/kg), albeit the two forms of heat loading may not
 always evoke the same physiologic effects. However,
 measuring the rise in body temperature, evaporation,
 and other physiological parameters during an
 exercise-induced heat load may aid our understanding
Figure 4-11.  Effect of exercise on heat loss, metabolic rate,
           and tympanic temperature of humans (data from
           Chappuis et al. 1976).
     Ambient Temperature = 30°C
     Rest T Work at 40 WJWork at 90 W~T Recovery
        0   20   40   60   80  100  120  140

                   Time, min
 of the human response to RF radiation. It is clear from
 the data in Figure 4-11 that healthy  humans can
 tolerate exercise-induced metabolic heat loads of 7.0
 W/kg without significantly stressing the thermoreg-
 ulatory  system.  However,  it remains to be shown
 whether humans can tolerate RF-induced heat loads
 of the  same magnitude  under  similar ambient
 conditions.
 Fever can be distinguished from forced hyperthermia
 in that the former represents a regulated increase in
 body temperature whereas forced hyperthermia can
 be defined  as an increased thermal  load  on the
 organism (Stitt 1979; Gordon 1983d). The process of
 pyrogenesis can be described in three major steps, as
 shown in Figure 4-12: (i) a hypothermia state where
 the reference or set-point temperature (in the POAH)
 has been raised above body temperature as a result of
 the introduction of a  pyrogenic agent, so that the
 animal generates and conserves heat to raise  body
 temperature to the new reference temperature; (ii) a
 febrile state where body temperature is regulated at
 the new reference temperature; and  (iii) a hyper-
 thermic state where the influence of the pyrogen is
 removed,  so that body temperature is now higher
 than the reference temperature and heat-dissipating
 mechanisms are employed to reduce body tempera-
 ture to the original normothermic level (Bligh 1973).
 The thermoregulatory state of a homeotherm
 exposed to RF radiation is similar to the hyperthermic
 stage of fever only when body temperature is above
 the normal set-point temperature of 37°C. Otherwise,
 the thermoregulatory  responses of fever and  RF-
 radiation  exposure are radically different. Fever
 entails a  regulated change in  body temperature,
 whereas during RF-radiation exposure the organism
 activates heat-dissipating effectors in an attempt to
 maintain normal body temperature.

 4.1.1.10 Mechanisms of Heat Gain During RF-
 Radiation Exposure
 The coupling of RF energy into irradiated biological
 subjects exposed to RF radiation was discussed in
 detail  in Sec. 3.2,  RF-Field  Interactions with
 Biological Systems.  To introduce  the  following
 material, some of the important  features of  that
 discussion are reiterated and summarized here.
The penetration of RF energy into biological tissues is
 dependent on the wavelength of the incident energy.
 Longer wavelengths can penetrate deeply into living
tissues,  but  the  shorter wavelengths found in the
 microwave region of the spectrum cannot penetrate
deeply. For example, at 2450 MHz (wavelength of
 12.5 cm) the RF energy is largely absorbed within
approximately  the first 2 to 3 cm of muscle tissue,
assuming  that the RF radiation is incident on an
object that is large in comparison to this wavelength
(such as a human). Similarly, RF energy of still higher
frequencies (i.e., in the millimeter wave region of the
                                                                      4-11

-------
Figure 4-12. Relationship between body temperature (TB) and set point (!„,) during pyrogenesis; (A) before the onset of fever; (B)
           during the rising phase of the fever; (C) during a maintained fever (heat loss and heat gain are balanced); (D) during the
           subsiding phase, and (E) after the return to normal body temperature (data from Bligh 1973).
          Normothermic
           at Normal Tb
Hypothermia
Normothermic
 at Raised TD
Hyperthermic
Normothermic
 at Normal Tb
    T..,
RF spectrum) does not penetrate more than 1 to 2 mm
and will be absorbed primarily in the skin.

This difference in penetration poses a major problem
for comparing the effects of RF radiation of various
wavelengths on thermoregulatory function in various
species. For example, when a 20-g mouse or a 70-kg
man is exposed to RF radiation of millimeter-size
wavelengths or to infrared (IR) radiation, the energy is
deposited in the  first 1 to 2 mm of  the skin in a
basically similar fashion. However,  at 2450 MHz a
mouse is compa rable in size to the wavelength so that
a resonant absorption condition exists and results in
efficient energy coupling with deep penetration and
nonuniform internal energy deposition. On the other
hand, when a human is exposed to the same 2450-
MHz  radiation,  the  energy  will  be deposited
peripherally within a few centimeters of the body's
surface in the area facing the radiation source.
                    interaction takes place at this frequency (Sec. 3.2).
                    The coupling of RF energy into the animal is less than
                    optimal (i.e.,  the coupling is of  intermediate
                    efficiency), and most of the energy is peripherally
                    deposited,  although less so for the rat than for the
                    man. If the same interspecies comparison were to be
                    performed under conditions  in  which the RF-
                    radiation coupling is optimal (i.e.,  the resonant
                    absorption case), then the results would probably be
                    very different (i.e., the power density values needed
                    for a  1 °C elevation in  rectal temperature may well be
                    significantly reduced). Telf and Harlen (1979) have
                     Figure 4-13.  Power densities at 2450 MHz necessary to raise
                                the rectal temperature by 1 °C in 60 min for the
                                rat. squirrel monkey, and  rhesus monkey (de
                                Lorge 1979). T. ranged from 22.5 to 24°C.
Notwithstanding these problems,  some  attempts
have been made to compare thermoregulation in
different species when subjected to RF radiation of
the same frequency. For example, de Lorge (1979)
compared the  power  density  at  2450  MHz that
induced a 1 °C elevation in the rectal temperature of
rats, squirrel monkeys, and rhesus monkeys. Using a
semi-log plot of power density vs. body mass (Figure
4-13), de Lorge extrapolated the data on rats and
sub-human primates to a 70-kg man. He predicted that
a value of 92 mW/cm2 is needed to raise the rectal
temperature of  man by  1°C at 2450 MHz. This
comparison  is probably acceptable because, for the
four species considered, a supraresonant kind of RF
                     E
                     i
        120

        100

         80

         60

         40

         20

          0
                               i   i i  nun    IT
                            _  Y = 24.88 Log X + 45.82
                               r = 0.975
                           0.1
                Rhesus
                Monkey
                                                    11111
                        1.0           10

                         Body Weight, kg
                                100
                        4-12

-------
indicated that if a human were exposed to 2450 MHz
at a power density of approximately 90 mW/cm2,
irreversible local peripheral tissue damage may occur
without an appreciable rise in body temperature.
When there are supraresonant-type interactions, the
absorption coefficient is roughly constant and  is
~ 0.5. Under these conditions, the RF-energy absorption
depends only on the geometric cross-sectional area
of the exposed animal. Assuming a spherical shape of
radius r for the animal, it can be shown that the cross-
sectional area-to-volume ratio is proportional to r23
(Sec. 4.1.4.1). This means that the  area-to-volume
ratio, as well as the area-to-mass ratio, decreases as
animal  size  increases.  Consequently,  at a given
power density, the absorbed energy  per unit of body
mass (i.e., whole-body-averaged specific absorption
rate,  SAR) decreases with  increasing size. This
general rule of thumb is valid  only for interaction
conditions of the supraresonant type and for
resonant- and subresonant-type conditions.

4.1.2. Effect of RF-Radiation on
Thermophysiological Effectors

4.1.2.1 Vasomotor Control

Interest during the early 1940's in using short-wave
diathermy (13 to 43  MHz) as a therapeutic agent
prompted research into the effect of RF radiation  in
producing localized changes in peripheral blood flow.
Near-field diathermic application with  capacitance
plates (SAR not determined) was shown to produce a
more  than twofold increase in blood flow of an
exposed limb of a dog (Wakim et al. 1948) and of a
human  (Abramson et al.  1957). Using  the  133Xe
clearance technique,  McNiven  and Wyner (1976)
found that localized exposure to 2450 MHz caused
nearly a fourfold increase in blood flow of the vastus
lateralis muscle (femoral)  in humans. Using local
diathermy application  to the human thigh (915 MHz)
Lehmann et al.  (1978) found that at a muscle
temperature  of 43 to 45°C, blood flow to the muscle
increased, which caused a  reduction  in muscle
temperature. Although it was not possible to
determine the SAR,  these studies are  important
because they contain some of  the few  data
concerning RF effects on blood flow in humans.
RF-radiation-induced  increases  in peripheral blood
flow can be attributed to a direct effect of heat on the
caliber  of arterioles, or to  an  indirect, neurally
induced vasodilation via the activation of peripheral
and deep-body thermal receptors .(Sec. 4; 1.1.5). The
increase in femoral blood flow during RF-radiation
exposure (2450 MHz and 27.3 MHz) is similar in dogs
with intact and denervated limbs (Siems et al. 1948).
Hence, RF radiation  can directly affect peripheral
vascular resistance;  however, the fact that this
response occurs in the denervated  limbs does not
preclude a neural response to  RF radiation under
normal circumstances.
Gordon (1983b) recorded the tail skin temperature of
mice exposed to 2450 MHz in a waveguide. As in rats,
the  tail  of the  mouse  is a  principal  site  of
nonevaporative heat exchange (Sec. 4.1.1.5). Heat
loss from the  tail is proportional to the difference
between  tail skin temperature and Ta. At a Ta of 25°C
an  SAR  of  11.5 W/kg was sufficient to promote
vasodilation in the tail. It was shown that the
vasodilation response was sensitive to the rate  of
heating  (i.e.,  SAR).  The integrated vasomotor
response (°C • s) was normalized to the absorbed heat
load (J/g) to yield a  skin temperature index (°C • s •
g/J). A doubling of  the SAR leads  to more than a
twofold  increase in the skin  temperature index
(Figure 4-14).  These data indicate that the mouse
responds not only to  the total amount of energy (heat)
absorbed from RF radiation but also the rate at which
the energy is absorbed.
Phillips ef al. (1975b) measured tail skin temperature
in the restrained rat immediately after exposure for
30 min to 2450 MHz at SARs of 0,4.5, 6.5, and 11.1
W/kg. At a Taof 24°C an SAR of 4.5 W/kg resulted in
a 1.5°C increase in  tail skin temperature. This was
associated with a 0.5°C  increase in colonic
temperature. Higher  SARs caused larger increases in
colonic temperature with minor additional increments
in skin temperature.
Adair and Adams (1980a) found that RF-radiation
exposure in the squirrel monkey at 2450 MHz and 8 to
10  mW/cm2 (SAR  = 1.5  W/kg) at a Ta of  26°C
promoted vasodilation in the tail without any change
in rectal  temperature. Furthermore, it was shown
that an equivalent power density from an infrared
heat source, which has a  much shorter wavelength
and is absorbed on the skin surface, was ineffective in
promoting vasodilation.
Exposing the  squirrel  monkey to RF radiation
promotes vasomotor  responses similar to direct
heating of the preoptic area/anterior hypothalamus,
an  extremely thermally sensitive  area of the
brainstem considered to be an integrative center for
the control of body  temperature (Sec. 4.1.1.3). The
conclusion was that RF-induced vasodilation was
caused by the activation of warmth-sensitive neural
sites in and/or outside the CMS, which in turn affect
the central neural control of heat-dissipating motor
outputs,  including  the dilation of  the  peripheral
vasculature (Adair and Adams 1980a).
4.1.2.2 Evaporative Heat Loss
Gordon  (1982a) measured whole-body evaporative
water loss (EWL) in mice exposed in a waveguide to
2450 MHz over a 90-min exposure period at a Taof
20°C (Figure 4-15A). EWL remained stable up to an
SAR of  29 W/kg. Above  29 W/kg  EWL increased
abruptly. In the same animal EWL was measured at a
Ta of 20,  25, 30, 33, and  35°C (Figure  4-15B).
                                                                     4-13

-------
Figure 4-14.  Examples of change in tail skin temperature (AT*) in restrained mice exposed to 2450 MHz at 25°C and specific
           absorption rate (SAR) of 10.6 W/kg (A) and 20.0 W/fcg (B). Note that the absorbed heat load (J/g) is similar in both
           cases (data from Gordon 1983b).
<  °
c/j    0
               10
   20

Time, min
                                     30

-
1 1
6.4 J-g"
i i
i i i i i
i i i i i

-
1
6.0 J-s
1
1 1 1 1 1
. -
1 I 1 1 1
40     0
10          20

       Time, min
30
Between 20 and 30°C EWL remained at the same
level as observed at SARs  of 0 to 29 W/kg. EWL
increased significantly at Ta's of 33 and 35°C. The
mouse increased EWL at an SAR of 29 W/kg and at a
T» of 20°C and at TB's of 30 to 33°C at an SAR of 0
W/kg. These data allow one to relate the effects of RF
exposure and Ta on the activation of a thermoregula-
tory effector (EWL).  Heat-stressed rodents can
actively increase EWL by raising their ventilatory
frequency (e.g.. Hart 1971) and by applying saliva to
the fur (Strieker and Hainsworth  1971). These two
mechanisms are no doubt responsible for the rise in
EWL which occurred in mice exposed to SARs above
29 W/kg or above a Ta of 30°C.
Michaelson et al. (1961) demonstrated the importance
of evaporative  cooling in  the thermoregulatory
response of dogs to high-intensity RF exposure (100
or 165  mW/cm2 at 2790 MHz). During exposure dogs
will pant to increase EWL. If the animal is dehydrated
its rectal temperture reaches a critical level  much
                           faster than in hydrated dogs. With exposures lasting 4
                           h (165 mW/cm2, SAR unknown) the dogs lose body
                           mass while hematocrit increases, which indicates a
                           depletion of body water.
                           The rate of heating, or SAR, is also a critical factor
                           in the control of EWL. Gordon (1982c) exposed mice
                           to 2450-MHz  radiation  in  a  waveguide for brief
                           periods of time while continuously monitoring EWL at
                           a T. of 30°C. EWL was converted to evaporative heat
                           loss (EHL) by assuming that 1.0 g of evaporated water
                           -was  equal to a heat loss of 2426 joules. After  an
                           episode of exposing the mice, total EHL normalized to
                           body mass in dimensions of J/g was calculated. The
                           total heat load absorbed from RF exposure was
                           calculated by integrating SAR over time, which yields
                           the dimension  of J/g. The open-loop  gain (OLG) of
                           EHL  was calculated by dividing the integrated EHL
                           response  by  the  RF heat load. The  OLG EHL, a
                           dimensionless  number, describes the sensitivity of
                           the control of evaporative water loss (Gordon 1982b).
                       4-14

-------
Figure 4-15. Evaporative water loss (EWL) of mice exposed to
           2460 MHz for 90 min: (A) Relationship between
           EWL and SAR of mice exposed at a T. of 20°C; (B)
           effect of T. on EWL of mice (data from Gordon
           1982a).
          1	\	1	1	1	1	1
      0      10       20      30      40

           Specific Absorption Rate, W kg'1
                  25           30
                 Ambient Temperature, °C
35
At a Ta of 30°C the OL.GEHL increased nearly three-
fold by doubling SAR. Thus, the mice responded not
only to the absorbed heat load but also to the rate the
heat load was absorbed. This finding  is important
because it indicates that biological sensitivity can be
defined not only in terms of a dose perse (i.e., J/g) but
also in terms of the dose rate (i.e., J/(g-s)). A similar
pattern was observed when skin temperature of mice
was measured during  2450-MHz exposure  (Sec.
4.1.2.1).
Using the same methods as described above, Gordon
and White (1982) exposed mice to whole-body heat
loads of 12 to 13 J/g by exposure to 2450 MHz at
SARs of  19,  68, or 194 W/kg while recording  the
OL.GEHL. The  OL.GEHL at a Ta  of 32.5°C displayed a
saturation at an SAR of approximately  160 W/kg. The
large change in OL.GEHL over an order of magnitude
change in SAR demonstrated a high  degree of rate
sensitivity at a low range of SARs.
Adair (1981) measured the rate of sweating from the
foot of squirrel monkeys exposed to 2450 MHz at a T.
       of 33°C (just below the monkey's upper  critical
       temperature). Sweating was activated consistently
       with 10-min exposures at power densities of 6 to 8
       mW/cm2 (SAR - 1.1 W/kg) (Figure 4-16). Lowering
       Ta led to an increase in the threshold SAR for activating
       sweating and a decrease in the sensitivity of sweating
       (i.e. A sweating/A power  density).
       4.1.2.3 Metabolism
       When a homeotherm is exposed to a Ta below its lower
       critical  temperature  it must increase metabolism
       above the basal level to maintain a normal deep-body
       temperature (Sec. 4.1.1.4). If a homeotherm is
       exposed to RF radiation at a Ta below its thermoneutral
       zone one would expect a decrease in metabolic rate
       as the animal substitutes RF heat for metabolic heat.
       On the  other hand, RF exposure  at a TB above the
       upper critical temperature of the thermoneutral zone
       should lead to an increase in metabolic rate as the RF
       radiation causes additional heat stress to the animal.

       Ho and Edwards (1977b) exposed mice to 2450 MHz
       at a  Ta  of  24°C  in a waveguide while recording
       metabolic rate by indirect calorimetry (i.e., oxygen
       consumption). They found  a  decrease  in metabolic
                                                   Figure 4-16. Change in the rate of sweating from the foot of a
                                                              squirrel monkey exposed for 10 min to 2460
                                                              MHz. The parameter is ambient  temperature.
                                                              Data are from Adair (1981).
       c
       E

       .1
       £
       to
       a:
       o>
       o>
       o>
       o
5.5


5.0


4.5


4.0


3.5


3.0


2.5


2.0


1.5


1.0


0.5
2450-MHZ CW Microwaves
10-min Exposures
             02   4   6   8   10  12  14  16  18  20

                        Power Density (mW/cm2)

                          4-15

-------
rate when SAR for a 30-min exposure exceeded 10to
23 W/kg. In another study (Ho and Edwards 1979)
mice were exposed to 2450 MHz at Ta's of 20, 24,30,
and 35°C. Exposure at SARs greater than 10 W/kg at
T.'s of 20 and 24°C caused depressions in metabolic
rate. However, the decrease in metabolic rate never
reached the level  of the sham irradiated group
maintained  at a  Ta of 30°C (i.e.,  thermoneutral
temperature for the mouse). At a TBof 35°C metabolic
rate increased during microwave exposure.
Phillips et a/. (1975b) measured oxygen consumption
and carbon dioxide production in rats immediately
after being exposed to 2450 MHz for 30 min at SARs
of  0  to 11.1 W/kg at a Ta  of 24°C. Oxygen
consumption was not affected at an SAR of 4.5 W/kg.
At SARs of 6.5 and 11.1 W/kg oxygen consumption
decreased and did  not recover for at least 300 min
following microwave exposure (Figure 4-17). Carbon
dioxide production followed a pattern similar to that of
oxygen consumption.

Adair and Adams (1982a) recorded metabolic rate at
TB'S of 15, 20, or 25°C in restrained squirrel monkeys
exposed  to  2450  MHz  in an anechoic chamber.
Reductions in metabolic rate were achieved with 10-
min exposures at power densities of 4 to 6 mW/cm2
(SAR = 0.6 to 0.9 W/kg). In a 10-min exposure period
the metabolic response was vigorous with a nearly
2.5-W/kg decrease in metabolic rate for a 1.0-W/kg
increase in SAR. However, with prolonged exposure
(90 min) metabolic rate adapted to a level where there
was an  approximate 1:1 substitiution of microwave
energy for metabolic energy.

To  summarize briefly, the metabolic rate of three
species  has  been  measured at a  Ta below the
Figure 4-17.  Oxygen consumption and carbon dioxide produc-
           tion of rats immediately after a 30-min exposure
           to 2450 M Hz at a T. of 24°C (data from Phillips ft
           •/. 1975b).
  8
c


I6
                          Oz Consumption
                   468

                      SAR, W/kg
10
12
thermoneutral zone during exposure to 2450 MHz.
For the 0.032-kg mouse at a Ta of 24°C, metabolic
rate is reduced reliably when SAR exceeds 10 to 12
W/kg; for the 0.43-kg rat at a Ta of 24°C, metabolic
rate is reduced at an SAR of 6.1 W/kg; and for the 1.0-
kg squirrel monkey at a Ta of 20°C, metabolic rate is
reduced at an SAR of 0.9 W/kg.

4.1.2.4 Thermoregulatory Behavior

Artificially raising the temperature of the hypothalamus
(Cabanac and Dib 1983; Gale eta/. 1970), spinal cord
(Carlisle and Ingram 1973), and rectum (Adair 1971)
will activate behavioral thermoregulatory responses
leading to  a  reduction in the preferred Ta.  These
changes in behavior can be initiated before there is a
change in skin temperature. Hence, it is not surprising
to find that  deeply  penetrating RF  radiation can
similarly affect behavioral thermoregulation as does
local, artificial  warming  of  thermosensitive areas
described above.

Gordon (1983c) measured thermoregulatory behavior
of unrestrained CBA/J mice in a combined  wave-
guide-temperature gradient system during exposure to
2450 MHz. Without microwave exposure, the mice
selected a  preferred ambient temperature (PTA)  of
31.5°C. The mice did not select a lower PTA until SAR
exceeded 5.3 W/kg. After a 1 -h exposure at 18 W/kg,
the mice selected a PTA that was 9.5°C cooler than
during the sham treatment. Then, 30 min after the
termination of RF exposure, the mice returned to the
warm end (~ 30°C) of the temperature gradient.

Using  a waveguide-temperature gradient system
similar  to  that described above, Gordon (1983a)
measured PTA and ventilatory frequency of BALB/c
mice during exposure for  60 min to 2450 MHz. The
mice did not select a cooler PTA until SAR exceeded
7.0 W/kg. At an SAR of 25 W/kg the mice selected
the coolest part of the temperature gradient (19°C).
There were slight increases in ventilatory frequency
at 20.5 W/kg when the mice could behaviorally select
a cooler PTA. However, if the mice were forced (while
unrestrained) to remain at their normal PTA of 31 °C,
ventilatory frequency increased significantly at SARs
of 9.6 W/kg and above (Figure 4-18). Thus, mice will
preferentially activate a behavioral thermoregulatory
response (selecting a cool PTA) rather than activate
an autonomic effector (ventilatory frequency) during
RF exposure.

Stern et a/. (1979) trained  rats to bar press for
infrared heat  in the cold and then exposed them to
2450-MHz  RF radiation for  15-min periods. They
found a decrease in the rate of bar pressing at a power
density of 5 mW/cm2 (SAR = 1.0 W/kg) compared to
bar-pressing  activity at  0 mW/cm2. This power
density, as well as exposure to 10 and 20 mW/cm2,
did  not produce any increase in rectal temperture
under similar environmental conditions (Ta = 5°C, fur
of rat clipped from body). Thus, behavioral thermoreg-
                       4-16

-------
Figure 4-18. Ventilatory frequency and preferred T. of mice
           during exposure to 2450 MHz inside a waveguide-
           temperature gradient system. Responses without
           temperature  gradient were measured in mice
           exposed to RF radiation at their preferred T. of
           31 °C. Data are from Gordon (1983a).
Ventilatory Frequency, breaths/min
I 8 8 8

i i i i •
t,i
/'-*- With Temperature Gradient
^X -••- Without Temperature Gradient ~
• /' ^^~
i i i i

u 38
o
a
3
d Temperat
i
0
I <
I 1 1 1

"\^
l i i i
D 10 20 30 40
              Specific Absorption Rate, W/kg
ulation was offset by RF-radiation exposure in the
absence of a change in rectal temperature.
Adair and Adams (1980b) trained squirrel monkeys to
control their Ta  while  being exposed for  10 min-
periods to 2450 MHz in an  anechoic chamber.
Without microwave exposure the monkeys selected
a Ta of approximately 35 to  36°C. Power densities
below 6 to 8 mW/cm2 (SAR -1.1 W/kg) had no effect
on the controlled Ta. Increasing power density above
6 to 8 mW/cm2 led to a decrease in the controlled Ta
(Figure 4-19). For example, at an exposure level of 22
mW/cm2  (SAR  ~  3.2 W/kg), preferred Ta was
reduced by 5°C while rectal  temperature  remained
constant. Exposure to infrared radiation at the same
power densities as RF exposure elicited no change in
thermoregulatory behavior.
Using a similar experimental apparatus, Adair and
Adams (1982b) exposed squirrel monkeys to 2450
MHz for 5 to 150 min while observing changes in
thermoregulatory behavior. Exposure to 4 mW/cm2
{SAR ~ 0.6 W/kg) had no effect on controlled Ta no
matter how long the exposure lasted. Exposure to 10
and 20 mW/cm2 (SAR ~ 1.5 and 3.0 W/kg) resulted
in a lowering of  the controlled Ta by 1.5 and 3.0°C,
respectively. With few exceptions, the duration of RF-
radiation exposure had no significant effect on the
controlled TB.
 Postural adjustments in an RF field can be used to
 modify energy absorption, as well as the rate of heat
 loss,  resulting from  an RF-radiation heat load.
 Although rectal temperature was not measured, this
 behavior may still be viewed as a form of behavioral
 thermoregulation  (Sec. 4.1.1.7).  Monahan and Ho
 (1977) exposed mice to 2450 MHz inside a waveguide
 while measuring SAR. They found that at sufficiently
 high  intensities the  mice would reduce their
 absorption  of  RF power,  presumably by postural
 adjustments. For example, at a Ta of 30°C the percent
 power absorbed did not decrease until SAR equalled
 or exceeded 25.8 W/kg. At a Ta of 20°C, changes in
 RF absorption were  not seen until SAR equalled or
 exceeded 43.6  W/kg.
 Gage et al. (1979) monitored the orientation of mice
 and rats exposed to 2450 MHz for 1 h in the far field at
 a power density of 15 mW/cm2. Depending on the
 type of animal  cage, by reorienting their position in
 the RF field, mice could change their SAR by a ratio of
 1:1.2 to 1:2. However, because of their large  size
 relative to the RF wavelength, the rats could not
 change SAR by reorientation. AtTa's of 22 and 28°C,
 exposure for 1  h  elicited no significant change in
 orientation for the mouse or rat (maximum SAR ~ 3.6
 W/kg for rat and 11.1  W/kg for mouse). The lack of
 effect in  mice may be due to the relatively  low
 intensity of exposure. Monahan and Ho (1977) did not
 observe a change in absorption until SAR equalled
 25.8 W/kg at a Ta of 30°C. The lack of effect in the rat
 is probably due to the inability of this species to
Figure 4-19.  Mean T. selected by one squirrel monkey exposed
           to 2450-MHz RF radiation (data from Adair and
           Adams 1980b).
                 2540-MHz CW Microwaves
O
0
1 37~
-H 36-
TJ
CD
8 35-
0)
£ 34-
3
I «-
E
,2 32-
c



Threshold
\\ i
J T
f T
1 I
t
i




| 02468

•

A



I
I
}'



10 12 14
c Incident Power Density,

S 0 0.5 1.0
1 1
1.5 2.0

Mean of 5 Runs
(Series 1) ±1 SEM
Mean of 5 Runs
(Series II) ± SEM







I

16 18 20 22
mW/cm2














24


2.5 3.0 3.5
Specific Absorption Rate, W/kg
                                                                      4-17

-------
reduce SAR regardless of orientation in the far field at
2450 MHz.


4.1.2.5 Endocrine Systems
The  thyroid  and adrenal glands are two principal
endocrine glands that are highly sensitive to thermal
stimulation from RF radiation, as well as to other
sources of thermal stimuli. For example, reducing Ta
or direct cooling of the preoptic/anterior hypothalamic
area (POAH) leads to an increase in the serum level of
thyroid-stimulating hormone (TSH) and thyroxine
(Gale 1973). Prolonged exposure to heat or cold leads
to hypertrophy of  the  adrenal  cortex, along with
increased secretion of glucocorticoids. Local heating
of the POAH leads to a rise or fall (species dependent)
in glucocorticoids  and a rise in serum  levels  of
antidiuretic  hormone. Hence, thermally sensitive
neurons in the POAH and perhaps other sites  in the
CNS mediate some control  over  the thyroid gland,
pituitary gland, and the cortex of the adrenal  gland
(Gale 1973).
Parker (1973) exposed  rats to 2450 MHz at 0 and 15
mW/cm2 for du rations of 4,16, or 60 h at a Ta of 22°C.
Plasma-bound iodine (PBI) and thyroxine in the serum
were significantly reduced at 15 mW/cm2 for the 60-
h exposure (SAR ~ 5 W/kg, assuming a conversion
factor of 0.3 W/kg per mW/cm2; Durney et al. 1978).
These parameters were not affected at 10, 20, or 25
mW/cm2 for the 16-h exposures. Lu et al. (1981)
found a decrease in serum TSH in the rat following
exposure to 2450 MHz for 1 h at 10mW/cm2 at a Taof
24°C (SAR  = 2.5  W/kg).  The changes  in  serum
thyrotropin were significantly correlated with RF-
radiation effects on body temperature. Lu era/. (1977)
found  decreases in serum  thyroxine in  rats after
exposure to 2450 MHz at 20 mW/cm2 for 4 or 8 h at a
Ta of 24°C (SAR = 5 W/kg). There was a transitory
stimulation of thyroid function in rats exposed for 1 h
to 1  mW/cm2 (SAR = 0.25 W/kg).
Lu  et  al. (1977)  found an increase  in serum
corticosteroid levels in rats exposed to 2450 MHz at
20 mW/cm2 for 8 h (SAR = 5 W/kg).  Lotz and
Michaelson  (1978) found  a positive correlation
between serum corticosterone and cotonic tempera-
ture in rats exposed to 2450 MHz for 30,60,  or 120 m in
at a Ta of 24°C. This  relationship has  also been
reported in humans exposed to extreme ambient heat
stress  (Follenius et al.  1982; see Sec. 4.1.10). No
significant increases in plasma corticosterone for the
30-  and 60-min exposures were observed below
power  densities of  50  mW/cm2 (SAR ~ 8.0 W/kg).
However, the 120-min exposure elicited a significant
increase in plasma corticosterone at a power density
of 20 mW/cm2 (SAR ~ 3.2  W/kg).  Lu et al. (1981)
found a significant increase in plasma corticosterone
in rats exposed to 2450 MHz for 1 h at 50 mW/cm2
(SAR -10.5 W/kg) and for 4 h at40mW/cm2(SAR ~
8.4 W/kg).
Many of the above data indicate an accumulative
thermal effect of RF-radiation exposure on the activity
of the endocrine systems. The general trend is that
longer exposure times are needed to elicit significant
changes if) hormone levels as SAR is reduced. In
addition,  the RF-radiation-induced  activation of
thyroid  and adrenal  cortical systems  is intimately
related to RF-radiation effects on body temperature.
In general, hormone  levels in the blood increase or
decrease in unison with significant changes in core
temperature.  RF-radiation effects  on the endocrine
system are addressed in more detail in Sec. 5.7.1.


4.1.3 Body Temperature Regulation During RF-
Radiation Exposure
Referring to the heat balance  equation in  Sec.
4.1.1.2, it can be seen that when heat gain (from
metabolism or RF-radiation exposure) exceeds heat
loss there will be  positive  heat storage  and,
consequently, an increase in average body tempera-
ture.  Note the  use  of the term "average  body
temperature," meaning the average temperature of
all tissues in the body (Sec. 4.1.10). Changes in the
average body temperature represent imbalances
between heat gain and heat loss that cannot always
be detected by measuring rectal temperature alone.
For example, rectal temperature may remain  fixed
while subcutaneous temperature increases and
results in an increase in the average body tempera-
ture.

In a steady-state  condition  during RF-radiation
exposure, if the thermoregulatory effectors described
above (blood flow, evaporation, and behavior) cannot
dissipate the RF heat load, then the  average  body
temperature  will rise.  If we  assume that during
temperature regulation deep-body temperature (e.g.,
rectal, colonic, or core) should be regulated within a
restricted range (see Glossary), in a first approximation,
it can then be sard that when the core temperature
rises above the normal mean temperature by at least
one standard deviation (defined as hyperthermia by
Bligh and Johnson 1973) the regulatory system is no
longer capable of maintaining the regulated variable,
core temperature, within normal limits. However, this
conservative view of  defining the  characteristics of
the regulatory system may need to be compromised
because of peculiarities of temperature regulation
during RF-radiation exposure (see  below).
Principal factors that influence the ability of a species
to thermoregulate (i.e., maintain core temperature
within one standard deviation  of the  normal mean
level) during RF exposure are (i) species characteristics,
(ii) degree of restraint, (iii) wake-sleep state (including
anesthesia), (iv) ambient temperature, (v) relative
humidity,  and (vi) air velocity. Certain species are
better able to maintain a constant body temperature
in warm environments than others. For example, in
hot desert conditions a rodent weighing 100 g would
                       4-18

-------
 have to dissipate 15 percent of its body weight in
 water per hour to thermoregulate, whereas a 70-kg
 man need only  evaporate 1  to 2 percent per hour.
 Since a water loss of 10 to 20 percent per hour is
 lethal, small rodents in the desert burrow in the earth
 to avoid the heat during the day, whereas relatively
 large homeotherms, such as the human, can survive
 direct exposure from  the  desert heat (Schmidt-
 Nielsen  1964). Based on these examples one would
 predict tremendous species differences in thermo-
 regulatory capacity during  RF-radiation exposure.
 However, there are problems in relating ambient heat
 stress to RF-radiation heat stress (Sec. 4.1.4.3).
 Restraint  has  a  large impact on the ability to
 thermoregulate.  Restrained animals have a reduced
 thermoregulatory capacity (e.g., Frankel 1959). The
 effect of anesthesia  on thermoregulatory capacity
 at different ambient temperatures  has already been
 discussed (Sec. 4.1.1.8).  Generally,  anesthetics
 reduce the ability of  homeotherms to defend body
 temperature in the heat and cold.
 Ambient temperature (TJ, relative humidity (RH), and
 air velocity (V) affect the capacity of homeotherms to
 thermoregulate.  As Ta increases,  the temperature
 gradient between the skin and air is reduced, thereby
 causing  a reduction in passive, nonevaporative heat
 loss. Thus,  with increasing T«,  homeotherms rely
 more on evaporative heat loss to balance heat gain
 and  heat loss.  The  efficiency to evaporate is
 dependent on the partial pressure of water in the air
 (which  is, of course,  related to RH). The ability to
 evaporate water from a surface is reduced as the
 partial pressure  of water of the surrounding air
 increases.

 4.1.3.1  Mouse

 CD-1 mice exposed 100minto2450MHz(farfield)at
 28mW/cm2(SAR~22.2W/kg)atTa = 20°CandRH =
 50 percent  underwent a 0.8°C  increase in rectal
 temperature, when compared to the post-treatment
 shams (Berman et al. 1978). AJ mice restrained in a
 waveguide and exposed for several  minutes to 2450
 MHz at a Ta of 32.5°C (dry air) underwent an abrupt
 increase in colonic temperature at SARs of 20 to 205
 W/kg (Gordon and White 1982). The rate of body
 heating was 0.021 °C/min per W/kg increase in SAR.
 Gordon (1982d) exposed hypothermic (body temper-
 ture  equalled 17 to 30°C) CD-1 and AJ mice to very
 intense 2450-MHz radiation inside a waveguide at
 SARs of  200 to 1800 W/kg. The resultant warming
 rate  of colonic temperature of the restrained mice
 ranged from 0.05 to 0.65°C/s. Provided that body
temperature did not exceed the lethal limit (~43°C)
the  mice could  survive any warming rate  up to
0.65°C/s. Thus,  the rate of warming does not affect
the tissue per se. but it does have a large impact on
the degree of activation of thermoregulatory effectors
(Sees. 4.1.2.1 and 4.1.2.2).
 Smialowicz et al. (1981b)  used a unique form of
 hypothermia as an assay for detecting the thermal
 effects of low level RF-radiation exposure. Below the
 thermoneutral zone, mice injected intraperitoneally
 with 5-hydroxytryptamine (5-HT) became hypothermic,
 the decrease in body temperature being inversely
 related to Ta (Figure 4-20).  The  5-HT-induced
 hypothermia  was attenuated  in  a dose-related
 response by exposure to 2450-MHz radiation at 1 to
 10 mW/cm2. The  linear  regression of colonic
 temperature after 5-HT vs. Ta calculates to a slope of
 0.30°C colonic temperature/°C Ta. Plotting the 5-HT
 colonic temperature against power density yields a
 slope  of 0.15°C colonic temperature/1.0 mW/cm2.
 Dividing the  Ta response by the RF-radiation
 response, we obtain
Figure 4-20.  Effects of 6-HT injection* on mice (data from
           Smialowicz »t al. 1981 b). (a) Linear regression of
           colonic temperature after an intraperitoneal injec-
           tion of 5-HT at various T.'s. The drop in body
           temperature increases with decreasing T..  (b)
           Linear regression of 5-HT-induced hypothermia
           for various power densities at 2450 MHz. Similar
           to ambient temperature, the magnitude of hypo-
           thermia is less with an increasing power density.
     38
     36
     34
     32
  o
  o

  0)
  3
  <5
  c
  o
  o
  o
30
       I    I    I
I	I
I
   16      20      24       28      32      34

            Ambient Temperature, °C
          CBA/J Mice

           I    I   I    I   I
     32  -
     30
              2      4       6       8      10

             RF Radiation Power Density, mW/cm2

                    4-19

-------
    030   °C colonic temp
         1.0°C ambient temp  _  ,.,,   , ,„_
    	 = 2 mW/cmV°C
    0 15   °C colonic temp
            1.0 mW/cm2
That is, for the mouse a 2-mW/cm2 increase in the
power  density (2450 MHz) is equivalent to  a  1°C
increase in TB when tested at a T, of 22°C. This study
is one  of the very first attempts to relate ambient
temperature and RF-radiation exposure.

4.1.3.2 Hamster
Berman  et al. (1982) exposed  Syrian (Golden)
hamsters to 2450-MHz (CW, far field) at a 22°C Ta and
50-percent RH. The change  in rectal temperature
before  and after 100 min of exposure was -0.8°C ±
0.4 at 0 mW/cm2, -0.4°C ± 0.5 at 20 mW/cm2 (SAR
~ 6 W/kg), and + 0.8°C ± 0.7 at 30 mW/cm2 (SAR
~ 9 W/kg). Considering just the rectal temperatures at
the end of RF-radiation exposure, an SAR of 9 W/kg
induced a temperature of 39.8°C, compared with
38°C for the controls.

4.1.3.3 Rat
A majority  of the data on temperature regulation
during  RF exposure  has been collected from rats.
Phillips et al. (1975b) measured colonic temperature
in  restrained  rats immediately after they were
exposed to 2450 MHz at a Ta of 24°C for 30 min.
Colonic temperature  was  elevated by 0.6°C and
0.7°C  immediately after exposure  to  4.5 and 6.5
W/kg,  respectively. Immediately after exposure to
11.1 W/kg, core temperature was  1.7°C above the
control group. Within 30  min after  RF-radiation
exposure the colonic temperature of all  exposed
groups had  returned to  the unexposed  level.
However, during the course of recovery (over 5 h) the
colonic temperature  of  the exposed animals
undershot the control group. For example, the colonic .
temperature of the 11.1 -W/kg group was nearly 1 °C
below the controls 2 h after the termination of RF-
radiation exposure.
Several investigators have measured rectal or colonic
temperature of  unrestrained rats following  RF-
radiation exposure. Berman era/. (1981) exposed rats
to 2450 MHz for 100 min (Ta = 22°C, RH =50 percent)
at a power density of 28 mW/cm2 (SAR —4.2 W/kg)
and found colonic temperature to increase from 38.2
to  40.3°C.  Berman  and  Carter  (1984) exposed
unrestrained rats to 2450-MHz RF radiation for 100
min (Ta = 22°C, RH = 50 percent) to 40 mW/cm2 (SAR
~  6.0  W/kg) and found  an increase in colonic
temperature from 37.6 to 39.6°C. Lotz and Michaelson
(1978) found a significant increase  in colonic
temperature (0.27°C) of unrestrained rats following a
30-min exposure to 2450 MHz at a power density of
13 mW/cm2 (SAR -2.1 W/kg) at a Taof 24°CandRH
of  40  to 60 percent. Increasing the duration  and
power density led to a further increase in colonic
temperature.  From the same laboratory, Lu et al.
(1977) found a 0.7°C increase in colonic temperature
after 120 min of exposure to 2450 MHz (Ta=24°C. RH
= 40 to 60 percent) at a power density of 20 mW/cm2
(SAR — 5.0 W/kg). Under similar environmental
conditions, Lu et al. (1981) found a significant (0.5°C)
increase  in  colonic temperature after 60 min of
exposure to 2450 MHz at a power density of 20
mW/cm2 (SAR - 4.2 W/kg).

D'Andrea et al. (1977) found a significant rise in
colonic temperature  (~ 0.04°C/min) in unrestrained
rats (Ta = 21 °C, RH =  27 percent) exposed to 600 MHz
for 55 min at a power density of 10 mW/cm2 but not
at 7.5  mW/cm2 (SAR  ~  7.5 and  5.7 W/kg,
respectively). In the same study, rats were exposed to
RF frequencies of 400, 500, 600, and 700 MHz at a
power density of 20 mW/cm2 while the rate of
increase of colonic  temperature was measured. A
frequency of 600 MHz was found to have the greatest
thermal effect. Changing frequency in 100-MHz
increments above or below  this  point resulted in
lower rates of warming. Thus, 600 MHz appears to be
the resonant frequency for the rat (body weight — 350
g). The Radiofrequency Radiation Dosimetry Handbook
(Durney et  al.  1978) also predicts 600 MHz to be
resonant for the rat, as based on a prolate  spheroid
model.

Phillips et al. (1973) assessed the effects of repeated
RF-radiation exposure on metabolic rate and on skin
(tail) and colonic temperature immediately after a 30-
min exposure to 2450 MHz at an SAR of 11.1 W/kg.
One group of rats was exposed for 10 days to 30
min/day of 2450 MHz (11.1 W/kg). The control group
was subjected to 9 sham exposures followed by a 30-
min exposure to 2450 MHz  on the 10th  day. The
group that was exposed every day showed acclimation
to RF-radiation exposure. For example, the group
exposed only  once  to RF radiation  had  a 3.9°C
elevation in colonic temperature, whereas the group
exposed to RF radiation on 10 consecutive trials had a
temperature elevation of 3.1 °C. Skin temperature
and colonic temperature followed similar  patterns.
Both rat groups had similar depressions in metabolic
rate following exposure. This study shows that rats
can apparently acclimate to  repeated exposures of
2450 MHz.  Acclimation to  RF-radiation exposure
appears to enhance heat-dissipatory mechanisms,
since the  acclimated group  had  a lower  body
temperature following exposure.

The variability in temperature regulation of  rats
reported in the above studies is of interest. At 2450
MHz Berman et al.  (1981) found a 2.1°C rise in
colonic temperature  after 100 min of exposure at 4.2
W/kg,  whereas Lu et al. (1977), using similar
environmental conditions,  found that a  120-min
exposure to 5.0 W/kg led to only a 0.7°C increase in
colonic temperature. D'Andrea et al. (1977) did not
                       4-20

-------
observe significant colonic warming at 600 MHz until
SAR equalled 7.5 W/kg. The variation in methodology
(e.g.,  airflow, degree  of  restraint,  animal  training)
used at different laboratories appears to have a large
impact on the results  of  temperature  regulation
during RF-radiation exposure.

4.1.3.4 Rabbit and Dog
Ely and Goldman  (1956) exposed restrained rats,
rabbits, and dogs  to  2884-MHz (PW) radiation  A
rectal thermistor probe was connected in a closed-
loop feedback to the  RF generator such  that the
average power for maintaining rectal temperature at
a given level could be constantly  delivered to the
animal. They found for each species that  a power
density of ~ 25 mW/cm2 was required to maintain
rectal temperature 1 °C above normal. At this power
density the estimated SAR was 5.0 W/kg for the 0.2-
kg rat, 1.6 W/kg for the 4.0-kg rabbit, and 0.9 W/kg
for the 10-kg dog. This SAR value for a 1°C risemthe
temperature of the rat is similar to the data obtained
at 2450 MHz that have been reported more recently
by other investigators.  (See citations above.) It should
be noted that the  metallic temperature  probe may
have affected the measurement of rectal temperature
in the rat (Ely and Goldman 1956).
Michaelson et al. (1961) exposed confined dogs  to
2790 MHz (PW) while recording rectal temperature
at a Ta of 23 to 41 °C. Exposure for 1 to 2  h at power
densities of 100 or 165 mW/cm2 (SAR ~ 3.7 to 6.1
W/kg) caused elevations in rectal temperature  of
approximately 1.5°C at a Ta of 22°C. The change  in
rectal temperature with time had three principal
phases (Figure 4-21):  (I) during  the first 25 min  of
exposure the rectal temperature rose rapidly, with the
dogs increasing their ventilatory frequency (panting);
(II) for the next ~ 40 min  rectal -temperature was
relatively constant at the hyperthermic level; and (III)
after  approximately 1  h  of  exposure there was a
breakdown in the capacity of the dogs to dissipate
heat, and rectal temperature  increased rapidly,
approaching lethal limits. Anesthetizing the  dogs
with pentobarbital sodium or chlorpromazine greatly
reduced the capacity to maintain body temperature
during RF-radiation exposure. (See also Ely et  al.
1964).

The response of RF-exposed dogs to have  a relatively
constant but hyperthermic body temperature (Phase
II; see Figure 4-21) might suggest that the animals are
thermoregulating normally since rectal temperature
is  maintained at a steady level below lethal limits.
However,  as discussed  at  the  beginning  of this
section, when core temperature exceeds the normal
mean by more than 1 standard deviation,  the animal
is classified as hyperthermic. Conservatively, one can
view any increase  in body temperature that occurs
during Phase II as a failure to maintain normal body
temperature; however, the rise in body temperature
may also be viewed as a response of the dog to better
Figure 4-21. Example of the triphatic rectal temperature
           response of a dog exposed to 2790-MHz RF
           radiation at a power density of 165 mW/cm2
           (data from Michaelson ft al. 1961).
    • Initial Heating
    • Increased
      Respiratory Rate
    • Decreased Depth
    • Panting
             • Thermal Equilibrium
• Breakdown
  in Thermal
  Equilibrium
• Collapse
  43


  42 -I
165 mW/cm2 2790MHz

     I                II
           15    25    35    45    55    65   75  85

                       Time, min
survive the thermal effect of RF-radiation exposure.
For exa mple, if the average body temperature of a 10-
kg  dog is allowed  to rise  by 2°C, a considerable
amount of water is saved that would normally be used
to dissipate the additional heat. The water savings
would be calculated as:

Water saved (g) =(ATb) 
-------
of power density at 2450 MHz to achieve a 1 °C rise in
rectal temperature of the rat, squirrel monkey, and
rhesus monkey and  the logarithm of body mass (cf.
Figure 4-13). The approximate SAR was 5.8 W/kgfor
the rat, 2.5 to 4.5 W/kg for the squirrel monkey, and
4.7 W/kg for the rhesus monkey (Ta = 22.5 to 24°C).
Adair and Adams (1982b) found very slight increases
in rectal temperature of squirrel monkeys exposed to
2450 MHz at SARs from 1.5 to 3.0 W/kg (initial Ta -
35°C). However,  in these experiments the monkeys
could control  their  Ta.  The  animals appeared to
counter the rise  in core temperature by selecting a
cooler Ta.
Lotz and Podgorski (1982) recorded rectal temperature
in restrained rhesus monkeys exposed for 8 h to 1290
MHz at power densities of 0, 20,28, and 38 mW/cm2
(SAR ~ 0 to 4.1 W/kg) at a Ta of 24°C and RH of 55
percent. Rectal temperature increased an average
0.5°C at 2.1 W/kg, 0.7°C at 3.0 W/kg, and 1.7°C at
4.1  W/kg. It is of interest to compare these results to
that of the de Lorge (1979) experiment with monkeys
working for food at a frequency of 2450MHz. At 2450
MHz and a Ta of 24°C, a 1 °C rise in recta I temperature
of the rhesus monkey was observed at an SAR of ~
4.7 W/kg (de  Lorge), compared with a 1.7°C rise
when exposed to 1290 MHz at 4.1 W/kg (Lotz and
Podgorski).
Lotz (1982) compared the effects of 225-MHz (near
resonance) and  1290-MHz (supraresonance) RF-
radiation exposure  on the rectal temperature of
rhesus monkeys at  a Ta of 24°C. Monkeys  were
generally exposed for 4 h at either frequency. At 1290
MHz a 0.5 to 0.6°C rise in rectai temperature was
achieved at a power density  of 28 mW/cm2 (3.0
W/kg). During exposure to 225 MHz, a similar rise in
rectal temperature was achieved at a power density
of only 2.5  mW/cm2  (1.2 W/kg). Thus,  as the
frequency approached resonance the efficacy to raise
rectal temperature of the rhesus monkey improved.
Overall, for the rhesus monkey at 2450 MHz, rectal
temperature rose 0.21 °C per W/kg increase in SAR; at
1290 MHz the conversion factor was 0.24 to 0.41 °C
per W/kg; and at 225 MHz the conversion factor was
0.45 to 0.78°C per  W/kg increase in SAR. It was
estimated that  the rhesus  mo'nkey  could not
thermoregulate within reasonable limits (42°C rectal
temperature) for  longer than 1 h at an SAR of 2.4
W/kg at a frequency of 225 MHz.
4.1.3.6 Lethality
Heretofore this section has discussed the following
points:  (i) relatively  low levels  of  RF-radiation
exposure lead to the activation of thermoregulatory
effectors such as peripheral vasodilation, evaporation,
metabolism (decrease), and behavior (selection of a
cool Ta), which together increase heat loss from the
body to the environment; and (ii) as the level of RF-
radiation exposure  exceeds  the  capability of the
species  to dissipate the RF heat load, heat gain
exceeds heat loss, and there is a rise in average body
temperature.  If the exposure reaches the  level
whereby body temperature cannot be controlled, then
thermal death will occur. Therefore, it is appropriate
to review briefly the levels of RF-radiation exposure
that cause thermal death.
In general, the lethal core temperature of homeo-
therms is approximately  6°C above the average
normal core temperature (Table 4-2). These data are
derived from work done at high Ta's, not RF-radiation
exposure.  The information  is provided  only to
familiarize the reader with the general upper limits of
body temperature of homeotherms.
Table 4-2.    Approximate Normal and Lethal Core Tempera-
           ture* of Some Bird* and Mammals*
Animal
Marsupials
Euthenan mammals
Man
Birds, nonpasserine
Birds, passerine
Normal Core
Temperature
<°C)
35-36
36-38
37
39-40
40-41
Lethal Core
Temperature
(°C)
40-41
42-44
43
46
47
'Compiled by Schmidt-Nielsen (1979).
Rugh et al. (1974) measured lethality in mice exposed
to 2450 MHz in a waveguide for various durations up to
5 min. The forward or incident power into the
waveguide was 7.37 W. Assuming an average
absorption efficiency of 42 percent (Ho and Edwards
1977),  the estimated  SAR was  approximately 99
W/kg. The lethal dose, calculated in dimensions of
J/g  body  mass,  was measured  at various air
temperatures (15 to 40°C) and relative humidities(25
to 70 percent). The temperature-humidity index(THI)
was  used to assess the influence of the interaction
between temperature and  relative  humidity on
lethality (THI = 1.41T + 0.1 RH + 30.6). The lethal dose
at 2450  MHz varied from 20 to 77 J/g and was
inversely related to THI (Figure 4-22).
In a similar experimental  protocol Rugh (1976b)
measured the lethal dose of RF radiation and the final
rectal temperature at the time  of death in mice
exposed to 2450 MHz. The average lethal dose for 1 -
month-old  weanlings, 2-month-old animals, and
adults more than 12 months of age was approximately
40 to 45 J/g. The mean rectal temperatures at the
time of depth for male and female mice of all ages was
46.6°C; however, in isolated cases, body temperature
rose as high as 50°C before death occurred. The
average lethal temperature is substantially  higher
compared to previous reports (Table 4.2).
4.1.4 Effect of Body Size on Thermoregulatory
Sensitivity to RF Radiation
Few data exist on the whole-body thermoregulatory
effects of RF-radiation exposure in humans. Detailed
                       4-22

-------
Figure 4-22.  Effect of an increasing THI on the lethal dose of
           RF radiation (2450 MHz) in mice (data from Rugh
           «tal. 1974).
  100
             60      70       80       90    100
              Temperature-Humidity Index (THI)


information on the dose response of humans to whole-
body RF-radiation exposure may never be collected.
Thus, we must rely on data collected on laboratory
mammals.

Extrapolation of  RF-radiation effects  in laboratory
mammals to humans has been attempted (Michaelson
and Schwan 1973; de Lorge 1979; Ely et at. 1964). A
degree of  uncertainty exists in predicting the
response of humans to RF-radiation exposure by
extrapolation from  the known bioeffects data
collected  in laboratory mammals. An  obvious
disadvantage to  such extrapolations is the  large
difference in body  size between  experimental
mammals and adult humans. For example, the mass
of a laboratory rodent or a primate is 1.0 to 3.4 orders of
magnitude smaller than  the  mass  of a  human
(assuming an average adult male human body weight
of 70 kg).
In many respects, the thermoregulatory physiologic
processes of laboratory mammals and humans differ
markedly.  A prime  example  is the physiologic
mechanism of heat dissipation during thermal stress.
At high Ta's, primates rely principally on sweating to
dissipate excess body heat,  whereas rodents are
unable to sweat but can dissipate  heat  through
insensible evaporative water loss from the skin and
respiratory tract (Sec. 4.1.1.6).  Primates commonly
used for RF-radiation experiments, such as squirrel
monkeys (which sweat only on  feet and hands) and
rhesus monkeys sweat when subjected  to heat
stress,  but these mammals are considerably smaller
than humans (~ 1  and  5 kg, respectively). Hence,
physiologic data collected on  laboratory primates
during  RF-radiation exposure may  aid in predicting
human responses, but the dose-response differences
between humans and primates  may be very large.
 Computer models of the human response  to RF-
 radiation  exposure are useful for  estimating the
 probable magnitude of the threshold SAR capable of
 activating physiologic responses (Sec. 4.2). Modelers
 attempt to compensate for the problems mentioned
 above by using appropriate physiologic responses of
 adult humans to heat stress, for which abundant data
 exist, to calculate human responses to RF-radiation
 exposure. The physiologic  responses  incorporated
 into the computer models  are  those of humans
 exposed to high Ta's and/or exercise, but not to RF-
 radiation.  Thus,  these models can  be  used to
 calculate a dose-response relationship to RF-
 radiation exposure if the physiologic responses are
 the same as those of humans subjected to natural
 forms of heat stress (e.g., exercise and/or radiant or
 convective heat). However, the similarity of response
 between RF-radiation exposure and heat stress from
 exercise or exposure to high Ta's is currently under
 debate because of the uniqueness of  RF-energy
 absorption.
 The purpose of this section is to demonstrate that a
 comparative analysis of the bioeffects of RF-radiation
 exposure on laboratory mammals may be useful in
 predicting SAR thresholds for exposed humans, if the
 effe'ct is assumed to be due only to heating of tissue.
 Heretofore, data on the physiologic and behavioral
 effects of RF-radiation exposure have been presented
 for various laboratory mammals with body masses
 ranging from 0.02 to 7.0 kg, the largest body mass
 being 350 times  greater  than the smallest.  If a
 relationship exists between an animal's mass and its
 sensitivity to thermalizing levels of RF-radiation, then
 an analysis of the known physiologic effects and their
 corresponding  threshold SARs over such a wide span
 in body mass should allow a reliable extrapolation to a
 body mass of 70 kg. This would be an extrapolation of
 only a tenfold  increase in mass and could give an
 estimation of SAR thresholds for humans.

 4.1.4.1 Effect of Body Mass on Thermal
 Physiology of  Mammals
 An animal's weight, mass, and volume are propor-
 tional to the cube of its radial dimension, whereas its
 total surface area is proportional to the square of its
 radial dimension  (for  discussion,  see Schmidt-
 Nielsen 1975). Thus, an animal's total surface  area
 increases  by approximately  the  0.67  power  with
increasing body mass. Also, the surface area.mass
 ratio decreases logarithmically with  increasing body
mass. A regression line based on mammalian body
 mass vs. surface area/body mass has a slope of -0.33
(Figure 4-23).

The importance of the surface area:mass ratio in the
physiology  of homeothermic animals  can be
illustrated by  comparing the metabolism of two
species with large differences in body mass, such as a
70-kg human and a 0.03-kg mouse. In  Figure 4-23,
                                                                      4-23

-------
Figure 4-23.  Relationship between the surface area:body mass ratio and the body mass of various mammals (data from Altman and
           Dittmer 1972). The solid line represents the same relationship for a sphere of density 1.0.
     1.0-
           Mice
   
-------
Figure 4-24.  Relation between metabolic rate and body mass of mammals (data from Schmidt-Nielsen 1975). All measurements
           were made on resting animals under thermoneutral conditions.
    IOC-
(D
tr
.2
"o

     10-
      1 -
                           Mice
                              \
                                   Squirrels
                                               Rats
                                                                          Dogs •
                                                                                   Sheep
                                                                                         Humans
      0.001
                        I
                       0.01
 I
0.1
 I
1.0
                                                                            10.0
  I
100.0
                                               Body Mass, kg
larger mammal. The 72-percent increase in a human
represents slightly less than a doubling of metabolic
rate.

The  relation  between  body mass and percent
increase  in  heat  loss  necessary to maintain
normothermic  temperature during  exposure to RF
radiation at SARs of 0.1, 0.4, and 1-.OW/kg is shown
in Figure 4-25. An SAR of 0.4 W/kg represents the
recently suggested exposure guideline recommended
by ANSI  (1982). Because  of  the logarithmic
relationship between surface  area and  body mass,
the percent heat loss increases exponentially with
body mass.
Data from exercise physiology  research provide
valuable insight into the influences of body mass on
the threshold  percent  heat  loss  necessary for
activating  a physiologic response. In a comparative
analysis of these data,  Taylor (1977) calculated
the maximum rate of passive heat, or nonevaporative
heat, which mammals can dissipate as a function of
body mass. An inverse relationship was shown to
exist between body mass and  magnitude of
nonevaporative heat loss (Figure 4-26). For example,
a 0.02-kg  animal can dissipate 9 times its metabolic
heat production with nonevaporative heat loss,
whereas a 100-kg animal can dissipate only 4.5 times
its metabolic  heat  production (assuming a 20°C
gradient between body and ambient temperature). If
an increase in evaporative heat  loss is used as a
                                                 biological end point, the total heat loss of 0.02-kg
                                                 animal would have to exceed 90 W/kg before it was
                                                 observed, whereas that of a 100-kg species would
                                                 have to exceed only 6 W/kg. Of course, the thermal
                                                 load from exercise and RF-radiation exposure cannot
                                                 always be  equated physically or physiologically.
                                                 However, Taylor's analysis is presented here to show
                                                 that because of surface area:body mass relationships,
                                                 a small mammal can passively dissipate heat at a far
                                                 greater rate than a larger mammal.
                                                 Ambient  temperature is critical  in  determining  a
                                                 homeotherm's sensitivity to a heat load. For example,
                                                 at a Ta of 20°C, humans exercising at a work load of
                                                 40  W experience a 2.5-W/kg  increase in heat
                                                 production (and loss) and a  slight elevation in
                                                 evaporative heat loss. However, when the same work
                                                 is performed at 30°C, the increase in evaporative heat
                                                 loss is 5  times the magnitude of  heat loss at 20°C
                                                 (Sec. 4.1.1.6).

                                                 It would not be surprising to find, given the above
                                                 analysis,  that  relatively high SARs are required to
                                                 promote  a thermoregulatory  response  in  small
                                                 experimental  animals. Figure 4-27 is a plot  of the
                                                 SARs necessary to either activate a thermoregulatory
                                                 effector or  raise body temperature  in the mouse,
                                                 hamster,  rat, squirrel monkey, and rhesus monkey.
                                                 The data points essentially represent all the data
                                                 discussed in Sees. 4.1.1.1, 4.1.2, and 4.1.3. The data
                                                 include SARs required to raise skin temperature.
                                                                      4-25

-------
Figure 4-25.
Relation between mass and percent increase in
whole-body heat loss necessary to maintain
normothermia in mammals exposed to RF radi-
ation at SARs of 0.1. 0.4. and 1.0 W/kg. The
calculations are based on the assumption of
resting metabolic rate at a thermoneutral T., as in
Figure 4-24. Percent heat loss was calculated as
[SAR/fSAR + metabolic rate)] X 100.
   80-
 10
 CO
 =  40-
     0.01
Figure 4-26.
    0.1         1.0         10

           Body Mass, kg
                                                100
Effect of body mass on the maximum rate of
nonevaporative heat loss relative to resting metab-
olism in mammals during exercise at a T. of 20°C
(data from Taylor 1977). The calculations are
based on the assumption that thermal conduc-
tance is a constant 1 cal (cm2 • h • °C)~'.
    10
   2 5
      0.1
            1.0      10

            Body Mass, kg

          4-26
                                         100
1000
       Figure 4-27.  Relationship between SAR and T. on the activa-
                   tion of various thermoregulatory effectors and
                   elevation  in body temperature  in the mouse,
                   hamster, rat, squirrel monkey, and rhesus monkey
                   exposed to RF radiation. Data points represent
                   those discussed in Sees. 4.1.2 and 4.1.3. Expla-
                   nation  of code: B—threshold  for change in
                   preferred T., BT—elevated colonic temperature
                   (0.3 to 2°C), C—elevated serum corticoids,
                   EWL—threshold for increasing evaporative water
                   loss, MR—threshold for lowering metabolic rate,
                   P—threshold for altering posture, PVMT—thres-
                   hold for altering peripheral vasomotor tone (i.e.,
                   vasodilation), and VF—threshold for elevating
                   ventilatory frequency. Note logarithmic scale of
                   abscissa. RF—radiation frequency was 2450 MHz
                   for all species except rhesus monkey (1290 and
                   225 MHz).
o
o
eg
i
0>
h-
c
.2
1
30
20
40
30
20
10
40
30
20
10
40
30
20
10
40
30
20
10-
Mouse B B pVMT P
VF •• .\'EWLD
MR BTV -P
MR
Hamster
BT
BT T^MR
B BT
Squirrel Monkey B
MR 	
Rhesus Monkey
BT BT BT.BT
C
0.1 1.0 10 10C
                                                                                SAR, W/kg
elevate evaporative water loss, lower metabolic rate,
alter  behavioral  temperature regulation,  and raise
deep-body temperature by 0.3 to 2.0°C. The data are
also plotted as a  function of the Ta at which the
thermoregulatory parameters were mesured. A trend
is evident in Figure 4-27, despite the variations in Ta
and RF frequency; that is, the SAR required to activate
a thermoregulatory effector or raise body temperature
decreases with increasing mass of the species.
The lowest SAR shown in Figure 4-27 at which body
temperature of rhesus monkeys was elevated (Lotz
1982) is important because exposure occurred at a
near-resonant frequency of 225 MHz.  Resonant
exposure, in which the energy is deposited very deep
in the body, may be a worst-case situation. Similar
SARs  at a supraresonant frequency of 1290  MHz
caused less severe rises in body temperature of the
rhesus monkey. The mouse body temperature data in

-------
Figure 4-27 were also collected at a near-resonant
frequency (2450 MHz). For a 0.03-kg mouse exposed
to resonant RF radiation at a Ta of 22°C, an SAR of 22
W/kg was required to raise the body temperature 0.5
to 1 °C. For a comparable temperature rise in a 4-kg
rhesus monkey at a Ta of 24°C, an SAR of only 1.2
W/kg was required. Assuming that  large species
follow a pattern similar to that of small species, an
animal having  17 times the mass  of the rhesus
monkey (e.g., 70 kg, the mass of an adult human)
exposed to a resonant RF radiation would be expected
to undergo a similar rise in  temperature at an SAR
less than 1.0 W/kg.
4.1.4.3 Relating Heat Stress from RF-Radiation
and Ambient Temperature Exposure

To some, there is a paradox in the hypothesis of the
foregoing discussion concerning an inverse relation-
ship that exists between body mass and the threshold
SAR  for activation of a thermoregulatory response.
As mentioned earlier (Sec. 4.1.1.6), humans are well
adapted to warm  environments and  are able  to
survive  high Ta's much longer than  smaller
homeotherms such  as rodents, rabbits, and infra-
human primates. There is an abundance of research
on the functioning  of humans in hot environments
(Dill et a/. 1964; Hardy and Bard 1974). However, one
must  be very careful in  relating the adaptability  of
humans to survive high Ta's to that of thermoregulating
during RF-radiation exposure.

The problem with such a comparison is that the data
on human thermoregulation in hot environments are
reported intermsofTa(°C), relative humidity (percent),
wind velocity (m/s), etc. In the study of RF radiation
the whole-body dose rate is commonly measured  in
W/kg. The  dimensions °C  and  W/kg have no
relationship to each other. Simply stated, an RF-
radiation field represents  a  heat  source (i.e.,  M
increases in the heat balance equation; see  Sec.
4.1.1). On the other hand, increasing  Ta does not
represent a heat source (provided Ta is less than Tre)
but rather impedes the dissipation of metabolic  heat
(i.e., raising Ta lowers K, C, and R in the heat balance
equation). One can be misled by a prediction of an
animal's response  to  a W/kg dose  rate  if the
prediction is based only on the animal's response  to
ambient conditions.  For example, in Sec. 4.1.1.9 it
was  shown that a human exposed to a  very hot
environment of 55°C Ta underwent a positive change
in the rate of heat storage of 0.6 W/kg. A small rodent
could not possibly survive this Ta exposure for  very
long, mainly because its surface area/mass ratio is so
much larger than the human that the rodent will  heat
at a tremendously faster rate.  On the other hand, the
small rodent, because  of  the  surface area/mass
relation, can tolerate a much  larger dose rate of RF
energy (W/kg). A mouse with a body mass of 30 g has
a metabolic rate of 10 W/kg, which  is more than 16
times greater than the increase in the rate of heat
storage of the human exposed to a Ta of 55°C.
Another way of viewing this problem  is to look at the
change in heat production during  exercise. The
maximal increment in metabolism of mammals with a
mass of 0.018 to 25 kg is approximately 7 times
greater than the  basal metabolic rate (Hart 1971).
Well-trained athletes can endure 10 times the basal
metabolic rate. This implies that the 0.03-kg mouse
can endure an overall increase of approximately 50
W/kg, which is 3 times greater than the maximal
increase of a well-trained athlete (~ 15 W/kg). Thus,
it is apparent that  heat tolerance measured in °C (Ta)
is not a valid indication of tolerance  in terms of the
rate of tissue energy absorption by the whole body
(W/kg). The relationship in Figure 4-27 only affirms
this deduction. Small homeotherms  generally have
much  higher threshold SARs  than  larger species
during exposure to RF radiation.

4.1.5 Unresolved Issues
At least three major issues in thermal physiology
remain unresolved: (i) the effect of various thermal
environments (exogenous or endogenous) encount-
ered by most human beings that might affect  their
sensitivity to RF-radiation exposure, (ii) the validity of
extrapolating the known bioeffects data in laboratory
animals to humans, and (iii) the effect  of RF-radiation
frequency on thermoregulation. These issues are
discussed below.
(i)  Elevating Ta above  thermoneutrality places a
greater strain on the capacity of the thermoregulatory
system to dissipate body heat. In such a situation an
organism's normal thermoregulatory response to RF
radiation is compromised. Most of the data concerning
RF-radiation effects on thermoregulation have been
collected in experiments with animals at a Ta below
their  thermoneutral zone. At  a Ta below the
thermoneutral zone the thermoregulatory system has
a substantial capacity to dissipate the RF heat load. At
Ta's above the thermoneutral zone homeotherms rely
mostly on evaporation  to dissipate heat. One can
visualize the potentially stressful situation  of a
human exposed to 0.4 W/kg, or ~ 25 percent of the
resting metabolic rate, at a Ta of 35°C(96°F). At this Ta,
at least 90 percent of the heat  lost is dissipated by
evaporation (Sec.  4.1.1.9). Assuming that a human
must  evaporate  1  g  of  water for  each 2426 J
dissipated, in an environment of a Taof 35°C in which
the human is absorbing 0.4 W/kg RF radiation, the
amount of evaporative water loss would increase by
— 25 g/h -m2 over the base-line level of 50 g/h-m2. If
the exposure took place at a Ta below the thermoneu-
tral zone, most  RF heat could be  dissipated by
radiation and convection or by lowering  metabolic
rate.

Either hot and dry  or warm and humid environments
are commonly encountered  by much of the human
                                                                     4-27

-------
population during the summer. In addition to these
ambient heat loads, one must consider altered thermal
states created by exercise and fever. These conditions
raise body temperature, which makes the effects of RF
radiation on thermoregulation more apparent than in
the nonworking or afebrile subject. Overall, combina-
tions of abiotic factors (ambient temperature, relative
humidity, solar radiation, "wind chill," and insulation)
and biotic factors (fever, exercise, and the presence of
medication or pathological conditions that impair
thermoregulation) clearly make  predicting  the
thermoregulatory  capacity of  a human population
a very difficult task.
(ii) Can  the  thermoregulatory data collected in
laboratory animals be used to predict how  human
beings might respond to RF-radiation exposure? In
Sec.  4.1.4 a discussion is presented of how body
mass influences the  sensitivity of homeotherms to
thermalizing levels of RF radiation. On this basis  it
may be possible to extrapolate the known bioeffects
data of laboratory  mammals to humans. However,  it
should be noted that this concept of extrapolation is
not universally accepted in the scientific community.
For example, Adair et a/. (1983) were strongly critical of
the attempt by Gordon (1982a) to extrapolate  the
known bioeffects data from laboratory mammals to
humans based on body mass. It was stated by Adair et
al. (1983) "Since we understand the thermoregulatory
responses of  man far better than the responses of
any other  species,  and can  quantify them more
accurately, it is clear that no relation as that depicted in
Fig. 5 (i.e., the graph  showing  an inverse relation as
between body mass and threshold SAR) of Gordon
(1982a)can possibly exist." Accordng to Adair, in the
area of thermoregulation it is impossible at this time
to extrapolate from animals to humans (E.R. Adair,
personal communication).

However, Gordon (1983e) countered these criticisms
by stating that  our  understanding  of  human
thermoregulatory physiology is based on research of
exposure to conventional heat sources (e.g., high Ta,
infrared heat, and exercise), not to whole-body  RF-
radiation exposure. Thus, in the study of the effects of
RF radiation on thermoregulation we understand the
thermoregulatory responses of laboratory animals far
better than those of human beings (Gordon 1983e).
Because nearly  all research in the study  of RF
radiation has been performed on inf rahuman species,
it is essential to have a means of extrapolating the
animal data to humans. The use of body mass as an
independent parameter for scaling the bioeffects of
RF radiation may be useful.

(iii) Because of the nonuniformity of RF-energy
deposition, the frequency of RF radiation (at a given
SAR) has a large impact on the thermoregulatory
system. Resonant frequencies appear to be a worst-
case situation (e.g., see discussion on  primates in
Sec. 4.1.2). Average whole-body SAR perse may not
adequately define the potential thermal effect of RF-
radiation exposure because the frequency relative to
the size of the  animal  (that is,  infraresonant,
supraresonant, or resonant)  must be considered in
evaluating the thermoregulatory effects of RF
radiation.
                       4-28

-------
4.2 Numerical Modeling of
Thermoregulatory Systems in Man

Ronald J. Spiegel
This section reviews the models for calculating the
thermal response of  human beings exposed to RF
fields. From this review the following generalizations
can be drawn:
 • The models predict that an exposure to a plane-
   wave field at 80 MHz (the whole-body resonant
   frequency for the adult human) will produce the
   highest temperatures in the legs.

 • An exposure to a  plane-wave field at 200 MHz
   (partial-body resonance in the arms) produces the
   highest temperatures in the arms.

 • Elevated temperatures tend to occur where the
   localized SARs are the largest.
 • Higher localized temperatures are produced at
   whole-body resonance because the body absorbs
   more energy from  the incident field.

The limitations  of  the  numerical  models  are
discussed in Sec. 4.2.4, Unresolved Issues.

4.2.1 Heat-Transfer Models
It  is important  to develop  accurate mathematical
models to predict the  absorption of electromagnetic
energy and the  thermal  response  of biological
subjects.  Mathematical modeling  can provide a
method for  gaining  qualitative  and quantitative
information that may  be valuable in explaining and
understanding many  of the effects  measured in
experimental animals. More important, experimenta-
tion on human subjects is often unethical. Thus, to
extrapolate RF-field results in experimental animals
to equivalent effects in   man, one must know
thoroughly the dosimetry for  both  species. This
knowledge can  be achieved by the application of a
sophisticated and realistic numerical simulation of the
energy absorption and subsequent thermal response
of subjects exposed to RF fields.
Several investigators  have developed mathematical
models to calculate the thermal response of the
human body when it is subjected to different environ-
mental conditions or levels of exercise. The same basic
approach can be taken to develop a model to simulate
the thermal response  to RF radiation. However, this
model must take into  account that RF  fields deposit
energy nonuniforrnly in the human. The simple one-
dimensional heat transfer models used in the past do
not accurately simulate this condition. However, a
short review of these previous attempts to model man
in one dimension provides a good background for
developing a more general model.

One of the earliest attempts at modeling the human
body was made by Pennes (1948), who developed a
cylindrical model of a  human limb. This model was
first used to simulate the human forearm but was
later applied  to  the general  case of limbs. The
following factors were  included in this model:  (1)
radial conduction, (2) metabolic heat generation, (3)
convection to the blood,  and (4) environmental
exchange by convection, radiation,  and evaporation.

Machle and Hatch (1947) introduced the concept of a
core-and-shell model by comparing measured values
of rectal and skin temperatures representing the core
and shell temperatures used in the  model. Empirical
correlations for radiation, convection, and evaporation
were experimentally developed for  inclusion in this
model. A  modification by Kerslake and  Waddell
(1958) extended  the model to  include  the case of
complete skin wetness due to sweating.

Wyndham  and Atkins  (1960) further extended the
core-and-shell model by introducing several concen-
tric cylinders  representing the various  body layers.
This model used a finite-difference technique to solve
a set of resulting  first-order differential equations by
an analog computer.

A more  rigorous  analytical approach was  taken by
Hardy (1949), who applied the laws of thermodyna-
mics and heat transfer to the human system. This
analysis included  radiant exchange with  the
environment, thermal conduction through concentric
cylinders,  natural  and forced convection, and
evaporation from the skin and lungs. The analysis
was  accompanied by  experimental verification of
several of the calculated responses.

Wissler (1961, 1964) modified the models of Pennes
and Wyndham and  Atkins and combined them to
obtain a model of the entire human  body. This model
sub-divided the body into six elements: head, torso,
two arms, and two legs. Each of these elements was
assumed to  have the  following:  (1) a uniformly
distributed metabolic heat generation, (2) a uniformly
distributed blood  supply, (3) a  composition of
homogeneous materials,  and (4) a geometry of
isotropic cylinders. The  effects of heat loss through
the respiratory system and countercurrent heat
exchange between the arteries and veins were also
included.

In contrast to the  models based on concentric
cylinders, Crosbie era/.  (1963) used an  infinite-slab
model to represent an element of  the  body. These
investigators argued that the physics and physiology
of the human system could be better understood if
based on this configuration and that many uncertain-
ties in  previous  analyses  could be clarified. This
model was programmed on an  analog computer and
verified by experimental observation.
Smith and James (1964) developed another analog
model to study thermal stress in man. This model had
the following characteristics: (1)  metabolic heat
production in the working muscles, (2)  muscles
insulated by a layer of subcutaneous tissue, (3) blood
                                                                    4-23

-------
flow from the muscles to the skin, and (4) blood flow
between various elements of the body and heart. This
model considered radial conduction  through  three
concentric cylinders and countercurrent  exchange
between the arteries and veins, and  it was verified
experimentally.
The next major effort at modeling the entire human
body was made by Stolwijk and associates at the John
B. Pierce Foundation Laboratory. The initial effort by
Stolwijk and Hardy (1966) was a model composed of
three cylindrical segments, one each for  the  head,
trunk, and extremities.  The trunk was divided into
three concentric layers: skin, muscle, and core. The
head  and  extremities were divided  into two
concentric layers: skin  and core. In  this  work, the
concept of the body being composed  of a  controlled
system and a controlling  system  was suggested.
These investigators also did a rigorous  review to
determine  accurate thermal properties of the
constituents of the human body. This model was then
programmed for analysis by an analog computer and
compared to experimentally developed parameters.

The  1966  Stolwijk-Hardy  model was expanded
(Stolwijk and Cunningham  1968; Stolwijk 1969,
1971;  Stolwijk and Hardy  1977)  to  include six
segments: head, trunk,  arms, hands, legs, and feet.
All these segments were composed  of four layers:
skin, fat, muscle, and core. The geometry of each was
cylindrical except for the head, which was spherical.
This model was programmed for analysis by a digital
computer and included high metabolic heat production,
sweating, blood flow to all layers, and  convective and
radiant exchange  with  the environment.  Stolwijk's
model is one of the most comprehensive programs to
date, and has been used to investigate a variety of
heat stress  situations. For  example, for  better
understanding of  the effect of local hyperthermia,
heat has been deposited in local areas such as the
brain (Stolwijk 1980). In addition, several investigators
have adapted it to special cases; e.g., Montgomery
(1972;  1974a,b; 1975) used the model to study the
effects of man immersed in water.

Gordon et al. (1976) extended and improved the basic
ideas formulated by Stolwijk (1971)  and Wissler
(1964) to model the human temperature  regulatory
response after exposure in a cold environment. This
model characterized the human body as 14 cylindrical
and spherical segments with a cold exposure control
system. The control of metabolism, skin blood flow,
and muscle blood flow was achieved  by feedback
controller  signals consisting of the head core
temperature, mean skin temperature, and  mean skin
heat flux.  The heat flux  control  signal  was not
included in previous models, but the results of this
study indicate that it is important when modeling the
human thermoregulatory response in cold environ-
ments.
4.2.2 RF-Radiation/ Heat-Transfer Models

As discussed earlier, there exist reasonably realistic
RF-energy deposition models (Sec. 3.2.3, Analytical
and  Numerical  RF Electromagnetic Interaction
Models) and  heat transfer models (see above
discussion)  for the human body. However, few
attempts have been made to combine the two models
to predict the body's thermal  response under
exposure to RF fields. Inone study, Emery era/. (1976)
determined the  thermal  effects of  a uniform
deposition of RF energy for a one-dimensional model
of heat conduction. Although this model may yield
realistic values for whole-body temperatures,  the
heating pattern produced by nonuniform deposition of
energy may deviate substantially from that produced
by uniform absorption; this deviation may occur if
various  parts of the  body (head,  arms,  legs, etc.)
selectively absorb  energy from the incident field
because of whole- and  partial-body resonance. In
another study, Guy et al. (1978) used thermographic
determinations of the distribution of  RF energy in
phantom models of man, which were subsequently
used to  provide input  for Emery's one-dimensional
thermal  model. This method certainly accounted for
the nonuniformity in the RF-energy deposition, but it
apparently did not  account for heat flow along the
major axis of the  body.  This factor  is  important,
because the primary nonuniformity in the RF-energy
deposition will occur along  the body's major axis,
since the body is much longer than it is thick. Thus,
this model will tend to overestimate the temperature
profile in the body because the model allows heat
flow to occur only from the core to the skin. In reality,
as a result of localized RF-energy deposition, heat
flow must also occur along the major length of the
body.

The most general attempt to date has been the work
of Spiegel et al. (1 979, 1 980a), who have used block
models to calculate the RF-energy deposition in the
body with a two-dimensional extension of Stolwijk's
model to determine the resulting thermal response of
the body. This model allows heat flow from the core to
the skin as well as along the major axis of the body.
The authors used a transient heat conduction model
with internal heat  generation and heat dissipation.
The internal heat generation is caused by metabolism
and absorption of RF energy. The internal dissipation
is caused by convective exchange with the cardiovas-
cular system and a combined convective and radiant
exchange with the surrounding environment at the
surface  of  the  skin.  The equation simulating  the
response is, therefore,
     pc
        9t
    where p= tissue density
         c= tissue specific heat
         T = local tissue temperature
                       4-30

-------
          t=time
         V=tissue volume
          k=tissue thermal conductivity
      OEM = electromagnetic energy deposition
        QM = metabolic heat generation
        QE = evaporative heat dissipation in the skin
        QR = respiratory heat loss in the lungs
To solve Equation 4-1, the body is divided into several
finite elements and this relationship is applied to each
element or node.  The body is  represented  by 15
segments,  with  each segment subdivided  into 4
concentric layers — core, muscle, fat, and skin. The
head is  modeled by a sphere; the neck,  hands, and
feet are  approximated as single cylindrical segments;
the  arms  and legs  are  each divided  into  four
cylindrical segments, and  the trunk is divided into
three cylindricl segments. The radius and length of
each of these cylindrical  segments are based on
dimensions for a standard man (Diffrient et al.  1974).
Calculations of heat capacitance, thermal conductance,
and  density are based on the type of  tissue, the
surface  area, and the volume for each segment and
layer.
In this model,  the time and spatial derivatives are
represented by finite difference approximations, and
the resulting system  of equations is solved by an
iterative procedure in which the initial temperatures
are used to compute the temperatures a short time
later. These new  temperatures are then used to
compute the temperatures at the new time and so on
until  thermal steady-state conditions are reached.
Figure 4-28 shows the relationship of the various
thermal  conductances for  a typical segment that
               includes four nodes in the finite element analysis.
               Each lumped conductance represents heat exchange
               with the  other  nodes. The  quantities  KC and  KA
               represent the  radial and axial  conductances,
               respectively,  the values of which are determined by
               layer and segment geometry and by tissue thermal
               conductivity (Carslaw and Jaeger 1959).  Heat
               exchange between the skin and surrounding
               environment by convection  and radiation is repre-
               sented by the quantity H, which is determined in  the
               standard engineering manner as described by
               Stolwijk (1971).  The BF terms designate the amount
               of heat exchanged by each node and the central blood
               pool, and are a function of blood flow. Blood flow rates
               to  all segments, except the skin, are  set to basal
               values (Stolwijk 1971). The skin-blood  flow rate is
               controlled by vasodilatation, which is a function of the
               temperature  difference  between  the local skin
               temperature  and the skin  set-point temperature, as
               well as the difference in temperature  between  the
               hypothalamic temperature and  its set-point tempera-
               ture. It is believed that all tissues  respond to local
               temperatures in excess of 40°C by an increase in
               blood flow. However, the literature does not appear to
               provide enough  information to  include this response
               for all 100 compartments of the model.

               The other heat exchange modes shown in Figure 4-
               28 include metabolic heat production QM; heat loss
               by sweat evaporation on the skin QE; respiratory heat
               loss in the lungs QR; and electromagnetic heat input
               OEM. The respiratory loss is determined  according to
               Stolwijk, and the model assumes that the expired air
               has come to thermal equilibrium with the upper trunk
Figure 4-28.  Block diagram for one segment of the thermal model.
               OEM
            QR
                  QM
                        KA
 OEM


    QM
                                                      OEM
                OEM
                CORE
           BF
M
                                             KA
                                    MUSCLE
                        KA
                                 BF
n
                                             HS/NA/-
                                                                        QE
QM
                                                                KA
                                                         FAT
                                KC

                              •'WV-
                                                     QM
                                                                                   KA
                  SKIN
                                             KA
        Central
         Blood
         Pool
                                                     BF
                                                                KA
                                                                                   KA
                                                                     4-31

-------
core temperature and is saturated. This term enters
the equation for thermal balance (Equation 4-1) only
at the node that represents the core of the upper trunk
segment. Several mathematical models have been
proposed to  calculate  the heat lost  by  sweating
(Emery et al.  1976). All are empirical and attempt a
best fit to experimental data for various conditions,
such as for a sedentary subject or for various levels of
exercise. This model is based on the detailed work of
Stolwijk (Stolwijk 1971, Stolwijk and Hardy 1977).
Evaporation  basically is controlled by the  sweating
rate. As with the rate of blood flow,  the term that
simulates  sweating  is  a function  both of the
temperature difference  between  the local  skin
temperature and the skin set-point temperature and of
the difference between the hypothalamic temperature
and its set-point temperature. The body develops its
own source  of heat as a  result of metabolic  heat
production QM. For a sedentary subject, this value
can be at the basal level or can  be accelerated  by
shivering. Shivering is initiated when the body is
subjected to a low ambient temperature (< 28°C for a
nude subject) and experiences  a chill. Since the air
temperature for this study is always in excess of 28°C,
the effects of shivering and vasoconstriction are not
included  in the analysis. The RF energy input term
OEM is calculated by a block model comprised of 180
cubical cells of various sizes (Hagmann et al. 1979a).

4.2.3 Numerical Results

To illustrate the model used by Spiegel et al. (1980a),
Table 4-3 shows the resulting steady-state tempera-
ture distribution  in a resting, nude, 70-kg, 170-cm-
high man in a thermally  neutral environment (air
temperature = 30°C; relative humidity = 30 percent)
exposed  to a plane wave  at 80 and 200 MHz. The
electric field vector is oriented parallel to the major
axis of the  body (i.e.,  E  polarization). Whole-body
resonance occurs for the 80-MHz field, and partial-
body resonance occurs in the arms for the  200-MHz
field. For the 80-MHz case, an incident power density
of 10 mW/cm2 is used (SAR = 2.25 W/kg). For the
200-MHz field, incident power densities are 10 and
32.5 mW/cm2  (SARs = 0.58 and 1.9 W/kg,
respectively). As mentioned above, the RF-energy
deposition term OEM is inserted into Equation 4-1 in
a manner prescribed by the 180-cell model.

The distribution of temperatures is quite different in the
two cases. For the  80-MHz field, a thermal hot spot
(a temperature of around 41.6°C) is generated in the
lower thigh; for the 200-MHz field the hot spot tends
to occur in the arms.  In addition, for the  200-MHz
case there is an elevated temperature (40.6°C) in the
neck. Another difference is that the hot spot occurs
for an incident power density of 10 mW/cm2 for the
80-MHz field, whereas 32.5 mW/cm2  was required
to produce a  hot spot for the 200-MHz case. This
difference occurs because the whole-body SAR is at
least four times greater for the 80-MHz field.
Table 4-3.   Steady-State Temperatures (°C) ina Human Body
           After Exposure to 80- and 200-MHz RF Fields
                80 MHz           200 MHz
10mW/cm*
Segment
Head
Neck
Upper
trunk
Middle
trunk
Lower
trunk
Upper
humerus
Lower
Humerus
Upper
forearm
Lower
forearm
Hand
Upper
thigh
Lower
thigh
Upper
calf
Lower
calf
Foot
No.
1
2

3

4

5

6

7

8

9
10

11

12

13

14
15
Muscle
36.3
38.3

37.7

38.8

38.9

34.5

34.6

34.4

34.0
36.4

39.5

40.3

39.9

38.4
36.4
Core
37.4
38.9

37.3

37.5

37.8

35.9

36.0

35.7

35.0
36.6

40.3

41.6

41.2

40.0
36.8
10mW/cm2
Muscle
36.1
37.7

36.9

37.2

37.1

35.3

36.9

36.3

35.1
36.0

35.7

35.5

35.7

35.5
35.6
Core
37.0
38.2

36.9

37.0

37.0

36.5

38.0

37.5

36.1
36.2

37.0

36.7

36.9

36.4
35.8
32.5 mW/cm?
Muscle
37.0
40.6

37.6

39.3

38.4

37.0

41.2

39.8

36.9
36.4

35.8

35.7

36.7

36.7
35.9
Core
37.2
40.6

37.1

37.2

37.2

38.4

42.9

41.4

384
36.6

37.1

37.0

37.9

37.9
36.1
As another useful method to present the data. Figure
4-29 graphically shows the time it takes to obtain a
hot spot for a given level of incident  power density
level for the two frequency conditions. The production
of a hot spot can be viewed in terms of a strength-
duration model; for example, Figure 4-29 shows that
the time required to create a hot spot in the lower
thigh for an 80-MHz field takes only a few minutes at
50 mW/cm2, but at 10 mW/cm2 it will take an hour or
more of exposure.  In addition, different minimum
levels of power density are required to produce hot
spots at various  body locations.  For instance,  the
rectal temperature profile has an asymptote at a level
just above 35 mW/cm2, but the lower-thigh  core
temperature asymptote occurs around 10 mW/cm2.
For  200-MHz radiation, the lower humerus  core
temperature was considered,  and the intensity-
duration curve for this case is shown by the dashed
curve.

Based on these results the most critical temperature
elevations  evidently occur at whole-body  resonant
frequencies, as might be expected. In fact, the rise in
temperature  is sufficient  to produce a  hot  spot
(41.6°C) in the legs for an incident power density of
10 mW/cm2. The  increased blood-flow  response
(based on the calculated results) for tissue tempera-
tures in excess of 40°C was not included in the model.
Therefore,  the model  most likely overestimates the
magnitude of the temperature rise in the thigh.
                       4-32

-------
Figure 4-29. Incident power density vs. exposure duration to
           obtain a hot spot (41.6°C).
  100

   90

   80


   70


   60

   50

   40

   30

   20

   10

    0
——— Exposure at 80 MHz
— ~— Exposure at 200 MHz
      Lower Thigh Core Temperature
               50        100       150

                 Exposure Duration, min
                                             200
4.2.4 Unresolved Issues
Although the combined RF-heat-transfer model
reported  here is realistic enough to predict gross
effects and trends, the model could be further refined.
The most obvious refinement is the extension of the
two-dimensional heat transfer  model to a three-
dimension! simulation. Although this extension is
numerically straightforward, problems might arise in
computer storage requirements and in the computa-
tional time needed to generate the heating patterns.
In addition, altered tissue blood flowfor temperatures
in excess of approximately 40°C  has not been
implemented in the model. This modification would
be simple, but its  implementation  would require
experimentation to quantify the altered blood flow
rates.
Another  problem is that little  is known about the
effects of RF energy on  the  predominant heat-
dissipating mechanisms of sweating and vasodilata-
tion. Although the empirical models yield results that
compare well with experimental measurements from
a human subject placed in a room at a  controlled
temperature and relative humidity, it is certainly open
to question  whether good agreement  between
measured and calculated results would likewise be
obtained  when the thermal load is caused by an RF
field. This problem could be resolved by experimental
verification of the model  with animal models.
                                                                      4-33

-------

-------
                                             Section 5
                              Biological Effects of RF Radiation
 5.1 Cellular and Subcellular Effects
 John W. Allis
 The investigation of RF-radiation effects on cellular
 and subcelluar systems is principally an attempt to
 elucidate  specific biochemical mechanisms for
 the  interaction  of the radiation with macroscopic
 biological systems. This approach takes advantage of
 the relative simplicity of in vitro systems, the ability to
 control variables, the rapid and economical way with
 which results may be obtained, and the perceived
 ease of mechanistic interpretation of the results. In
 practice, however, these advantages are not always
 realized.
 The divisions between molecular, subcellular, and
 cellular systems are somewhat arbitrary, but for
 convenience, results presented in this section will be
 discussed  in  this format,  illustrated in Table 5-1.
 When investigating  molecular and subcellular
 systems, one normally looks at a single key structure
 or function where, one hopes, all other variables may
 be held constant.  In these cases, thorough under-
 standing of the end point under study allows the
 investigator to interpret results more easily in terms
 of a detailed biochemical mechanism. For instance, if
 a  change  in  the kinetics  of  an enzyme-catalyzed
 reaction is found,  the type  of inhibition  can be
 determined, the particular molecules involved can be
 identified,  and a hypothesis for the  effect can be
 constructed.
                  In some experiments on intact cells, these conditions
                  can also be approached. An example of this is sodium
                  (Na+) and potassium (K+) transport across the red-
                  blood-cell (RBC) membrane. The RBCs limited
                  metabolic activity is geared primarily to performing its
                  main function of transporting oxygen from lungs to
                  tissue. In this case, ion transport across the
                  membrane is I inked to the mode of energy production
                  in the cell and to very little else, and studies on this
                  system are relatively simple to interpret.
                  Experiments on most living cells are complicated by
                  complex biochemical systems, many of which have
                  multiple or alternate pathways.  Rather than isolated
                  functional changes, broad end points such as growth
                  or ^genetic changes are often  assayed. Here, the
                  investigator tries to use "biological amplification"
                  where, for example,  the  cell's growth  is highly
                  sensitive to disruptions of a critical but unspecified
                  metabolic step, or where a mutant appears that must
                  reproduce over several generations before sufficient
                  numbers  are  present to be  measurable.  The
                  investigator also recognizes that the redundancies in
                  a complex system like  a living cell may negate any
                  efforts to detect the putative effect.

                  This inherent complexity makes it difficult to ascribe a
                  detailed biochemical mechanism to effects found.
                  Such effects must be viewed as a starting point from
                  which a mechanism  may eventually be deduced. On
                  the  other  hand, effects on cells can also be used to
Table 5-1.   Classification of Cellular and Subcellular Experiments
Classification
                                   Type of System Studied
                                                Examples of
                                                Experimental Measures
Molecular

Molecular
Subcellular
organelles
Subcellular
organelles
Subcellular
organelles
Cellular

Cellular


Cellular
" Purified enzyme preparations

 Purified DNA preparations

 Membrane-bound enzymes

 Cell membranes, phospholipid
 vesicles

 Isolated mitochondria

 Red blood cells

 Bacteria


 Isolated neurons
Enzyme activity, bind-
ing to small molecules

DNA melting

Enzyme activity

Infrared, Raman spectra


Mitochondrial function


Ion transport

Growth, survival,
infectivity
Neuronal firing,
membrane potential
                                                 5-1

-------
rationalize or interpret—perhaps even predict—
effects in the intact animal.
Dosimetry for in vitro systems presents problems to
the researcher that are in many respects unique to
the study of RF-radiation interactions. In general,
average  SAR  in  the  sample is  relatively easy to
measure and  has become  the standard dose-rate
measurement  in  this research  area. The most
convenient methods are thermal measurements, i.e.,
analysis  of heating and/or  cooling data (Allis et al.
1977), and electrical measurements, where forward,
transmitted, and reflected power are recorded in  a
closed exposure  system. (See Sec.  3.4.) Thermal
measurements have become much more convenient
and accurate since the development of temperature
probes that do  not perturb the RF  field. A third
commonly used dosimetric  method is calculation of
SAR based on incident power or power density. This
approach is less  satisfactory than  the  methods
mentioned above because simplifying assumptions
must normally be  made to model  the  exposure
geometry. In this section on cellular and subcellular
systems, the SAR given for each study is the average
for the sample unless otherwise noted.

In most in vitro  as well as in vivo exposures, the
distribution of absorbed energy in the sample is not
constant throughout the sample. Most researchers in
the field  recognize this fact, and the standard
approach for treating the problem is to stir or agitate
the solution if possible. Normally this is an effective
way  to eliminate  thermal  gradients produced by
inhomogeneous absorption patterns. When stirring is
not possible, and occasionally even when it is,
potentially serious problems arise because the end
points measured for in vitro systems are often very
sensitive to temperature, inhomogeneities in
temperature within the sample can give misleading
results (Livingston et  al.  1979, for  example). The
potential  for this problem exists in some of the work
discussed later in this section.
Because of temperature sensitivity of biological end
points measured in many in vitro systems, researchers
are generally careful about maintaining temperature
control of the exposed sample and matching it to the
control or sham-exposed sample. Two methods are
normally  used. In one, the researcher establishes a
steady-state temperature in  the presence of the RF
radiation  and then begins the assay. In this case, the
exposed  sample is cooled  to  (or experiences an
ambient  temperature) below the temperature at
which the assay is conducted, with the RF energy
making up the temperature difference. The sham or
control  sample  is  matched to the steady-state
temperature and remains at thermal equilibrium with
its surroundings. Because the exposed sample is at a
higher temperature  than its surroundings, a
temperature gradient necessarily exists in the RF-
exposed sample that does not exist in the control. The
magnitude of the gradient depends on the energy
absorbed from the field and the sample geometry.
Among the more conventional methods of heating,
infrared exposure mimics  most closely the heating
patterns for in vitro samples exposed to RF radiation.
The second method of temperature control often used
is to allow the temperature of the exposed sample to
rise during exposure and then to mimic  the rise in
the control sample by applying conventional heating
techniques. This method has the same difficulty with
a gradient in the exposed sample as just described;
also, it is difficult to match precisely the kinetics of the
temperature rise due to exposure to  RF radiation
using conventional  techniques. However, properly
designed experimental apparatus and good techniques
can significantly reduce the difficulty.
One must also  recognize that  generally in  vitro
systems are primarily made up of a solvent medium in
which the cells or subcellular systems are suspended.
Temperature probes are macroscopic and necessarily
measure the temperature of the solvent, whereas the
measurement  desired  is that of the  temperature
experienced by the cell, or by the immediate
environment about the subcellular systems. On the
other hand, a more detailed analysis of the thermal
properties of a sample,  and  more  precise  and
accurate data, can be obtained from in vitro systems
than can be achieved with in vivo experimentation.
The foregoing discussion has presented the major
problems  with  dosimetry  for in vitro systems. In
general, however, these problems  are less severe
than those  encountered  with in  vivo work. The
investigation  of  mechanisms of interaction of  RF
radiation with biological systems is critical to a basic
understanding of the biological effects and is best
accomplished using  in vitro techniques.  If  careful
attention is paid to  dosimetry and if the difficulties
presented here are addressed, dosimetry need not be
a principal  concern in the  interpretation of the
experimental results.

The general conclusions that can be made at this time
concerning the effects of RF radiation on cellular and
subcellular systems are:
 • No  consistent biological effects have been
   demonstrated  for  molecular  or  subcellular
   systems exposed in vitro that can be attributed to
   RF-specific interactions.
- • No consistent effects have been demonstrated on
   growth and colony-forming ability of single cells
   that can be attributed to RF-specific interactions.
 • There is an indication that sodium and potassium
   ion transport across red blood cell membranes can
   be affected by exposure in vitro to RF radiation in a
   manner different from generalized heating.
 • The electrophysiological properties of single cells,
   especially the firing rate of neurons  in isolated
                       5-2

-------
   preparations, may be affected by RF radiation in a
   manner different from generalized heating.
5.1.1 Effects on Molecular Systems
To obtain insights into possible direct interactive
mechanism(s)  of  microwave  radiation at  the
molecular level, several investigators have attempted
to measure changes  in  biologically important
macromolecules exposed in vitro (Table 5-2). Enzyme
kinetics studies form a majority of the reports, but a
few examine changes in macromolecular structure.
In sum, these reports do not demonstrate consistent
effects on molecular systems exposed in vitro to RF
radiation that can be attributed to  RF-specific
interactions.

In one study of macromolecular structure, Hamrick
(1973) measured DNA melting curves after exposure
of calf thymus DNA to 2.45-GHz (CW) radiation for 16
h (SAR = 67 W/kg) and at dose rates to 160 W/kg for 1
h. Temperature was controlled during all exposures,
usually at 37°C, but for some experiments it was
maintained at 40, 45, and  50°C. The intent of the
study  was  to  determine  whether microwaves
disrupted hydrogen bonding between the DNA
strands of the double helix and thus affected  the
melting  curve. The assay was carried out after
exposure. In  one  experiment  the  investigators
preserved any  disruption of  hydrogen  bonding
between  the strands by performing the microwave
exposure with formaldehyde in the buffer. However,
all melting curves were virtually identical to those of
unexposed, temperature-matched controls.
                                                   In another report concerning macromolecular
                                                   structure  (Allis  1975),  the protein  bovine  serum
                                                   albumin (BSA) was exposed to 1.70- and 2.45-GHz
                                                   (CW) radiation (SARs ranging from 30 to 100 W/kg).
                                                   The author recognized that structural changes due to
                                                   microwave exposure may be reversible. In this study,
                                                   he attacked the problem by developing an exposure
                                                   apparatus in which  ultraviolet (UV)  and visible
                                                   spectrophotometric measurements could be per-
                                                   formed during  exposure. The difference  in  UV
                                                   absorption between the  exposed  and unexposed
                                                   samples was measured directly by a double-beam
                                                   spectrophotometer. Differences of  this type  reflect
                                                   small changes  in the surroundings of certain UV-
                                                   absorbing amino acids and, in this case, could be
                                                   interpreted as changes in the structure of the protein.
                                                   The  temperature of the  exposed  samples  was
                                                   controlled during exposure, and it was matched by
                                                   the temperature of the control sample. Temperatures
                                                   ranged from 24 to 32°C, depending on the SAR value.
                                                   Spectra were measured immediately upon beginning
                                                   exposure,  and again 30 min later with continuous
                                                   exposure.  The study results showed that no changes
                                                   in the UV  spectrum could be found over a variety of
                                                   structural  states of protein, so that structural changes
                                                   due to microwave exposure could not be inferred.

                                                   The ability of microwave radia~tion  to alter enzyme
                                                   activity has been studied by several workers.
                                                   Measurements  were performed during microwave
                                                   exposure by two groups, each using spectrophoto-
                                                   metric  measures of enzyme activity. Ward et at.
Table 6-2. Summary of Studies Concerning RF-Radiation Effects on Molecular Systems
Exposure Conditions
End Point Measured/
Effects
No change in UV difference
spectra measured over pH range
2.5-5.5
UV spectra and binding constants
for mononucleotides showed no
difference from controls
Experimental
System
BSA-
Ribonuclease
Frequency
(GHz)
1.70(CW)
2.45 (CW)
1.70ICW)
2.45 (CW)
Duration
(min)
30
30
Exposure
Facility
(type)
Waveguide
Waveguide
SAR
(W/kg)
30-100
39
Reference
Allis (1975)
Allis et a/. (1976)
No change in enzyme activity
No difference in melting curves


Inactivation of enzyme; probably
temperature inhomogeneity effect
at very high doses
                          Glucose-6 phosphate
                          dehydrogenase;
                          adenylate kinase; NADPH
                          cytochrome C reductase
                          DNA
2.45 (CW)  5
Waveguide  42
                                              2.45 (CW)
                          Horseradish peroxidase   2.45 (CW)
Heat inactivation of enzymes
found at highest SAR (T = 50 °C)
corresponded closely 'o heat-
treated controls
         60.
         960
         5, 10,
         20, 30,
         40

         4.5,
         18.5
Far field
Waveguide
67.
160

62,500-
375,000
                          Glucose-6-phosphate    2.8 (PW)
                          dehydrogenase; lactate
                          dehydrogenase; acid
                          phosphatase; alkaline
                          phosphatase
Heat inactivation of enzyme found Lactate dehydrogenase   3.0 (CW)   20
at SARs > 165 W/kg
Waveguide  -200-500
Ward era/. (1975)



Hamrick (1973)

Henderson et al. (1975)
         Belkhode et al.
          (1974a,b)
                                                               Waveguide  33-960    Bini et al. (1978)
"BSA = bovine serum albumin.
                                                                        5-3

-------
(1975) examined three enzymes (glucose-6-phosphate
dehydrogenase,  adenylate kinase,  and NADPH-
cytochrome c reductase) exposed to 2.45-GHz (CW)
radiation (SAR = 42 W/kg). All exposed and control
samples were maintained at 25°C. Exposure durations
were ~ 5 min, during which the enzyme activity was
measured. No differences between exposed and
control  samples were  found. Bini  et al.  (1978)
followed  the activity of lactate dehydrogenase
exposed to 3.0-GHz (CW) radiation (SARs between 33
and 960 W/kg). They demonstrated that the changes
found in the enzyme activity were entirely consistent
with calculations  of thermal inactivation of the
enzyme at the temperatures attained. The  sample
exposed  at 33  W/kg was  not different from the
unexposed control; all other exposures (SARs ranging
from 165 to 960 W/kg) showed evidence of enzyme
inactivation.

Belkhode et al. (1974a,b) reported the effect of 2.8
GHz  square-wave  modulated (1-kHz) radiation on
four enzymes: glucose-6-phosphate dehydrogenase,
lactate  dehydrogenase, acid phosphatase, and
alkaline phosphatase. Enzyme preparations were
analyzed after exposures (SARs ~ 200 to 500 W/kg
on average). Exposures were conducted at 37, 46.7,
and 49.7°C; enzyme activities were compared to the
activity  of sham-exposed samples  at the same
temperatures. Activities of the exposed enzymes at
each temperature were  indistinguishable from the
shams.
Henderson et al. (1975) reported a change in enzyme
activity that was interpreted as an indicator of direct
interaction on the  enzyme by microwaves.  In this
experiment, horseradish peroxidase was subjected to
2.45-GHz (CW) radiation (SARs between 62,500 and
375,000 W/kg) with the sample-exposed in a tube
(4.7-mm ID) that protruded through a waveguide. The
sample tube was surrounded by a concentric cooling
jacket, through  which  an organic  coolant  was
pumped continuously to maintain the temperature at
25°C. Thermocouples were placed in the sample tube
so that the sample temperature could be monitored
from positions just outside the waveguide. The total
volume of the exposed sample was —0.8 ml. A marked
decrease in enzyme activity was found at  62,500
W/kg after 30 min of exposure, and at 187,500 W/kg
after 20 min of exposure, even though the temperature
was reported never to exceed 35°C. It is possible that
the very high fields present .at these SARs could
produce field-specific effects. However, it  appears
likely  that very  high local heating occurred in the
sample that was responsible for enzyme inactivation.
This likelihood is  substantiated by the work of
Harrison et al. (1980), who performed liquid-crystal
thermography  under  similar exposure  conditions.
Temperature rises of as much as 0.3°C were recorded
within a micropipette suspended in a waveguide and
cooled with water circulating through the waveguide.

                       5-4
Henderson et al. (1975) exposed their samples at 5 to
10 times the levels used by Harrison et al.; also, the
latter researchers used water as a coolant, which
would attenuate the energy reaching the sample
much more strongly than would the organic solvent
used by Henderson et al. A study by Livingston et al.
(1979) graphically illustrates the effect of temperature
gradients during  microwave exposure.
The binding of small substrate-like molecules to the
enzyme ribonuclease was studied to determine
whether the binding relationship between an enzyme
and substrate could be affected by microwaves (Allis
et al. 1976). In this work, UV-absorption spectra were
measured during exposure to 1.70- and 2.45-GHz
(CW) radiation (SAR =39 W/kg). Measurements were
performed immediately upon beginning irradiation
and  after 30 min  of  exposure.  Neither  structural
changes in the enzyme-substrate complex nor
changes in the binding constants were  found.
5.1.2 Effects on Subcellular Organelles
There has been relatively little research on the effects
of microwave  radiation  on subcellular organelles
(Table 5-3). The  reports  included in this document
range from work with phospholipid bilayers (i.e.,
synthetic analogs of cell membranes) to experiments
on intact mitochondria, the energy-producing system
in eukaryotic  (e.g., mammalian) cells.  Most of the
work  to be discussed here has not demonstrated
effects of microwave exposure at dose rates ranging
from 1 to 430 W/kg. Two reports that have indicated
an effect do not meet all criteria for inclusion here and
are therefore  discussed with other reports that
present unresolved issues.

Two  reports describe work with  enzymes bound to
biological  membranes. The study by  Ward et  al.
(1975),  discussed above, focused on  the enzyme
NADPH-cytochrome c  reductase, which  is loosely
bound to the membrane of the endoplasmicreticulum
of rat liver cells. Allis and Fromme (1979) studied
adenosine triphosphatase  (ATPase) in RBC mem-
branes  and cytochrome  oxidase in the inner
mitochondrial membrane of rat liver cells. The latter
two enzymes are thought to  be integral parts of
membranes.  Conditions  of the  exposure were
identical for all three enzymes in that the essay was
performed spectrophotometrically during exposure to
microwaves. In the latter study the dose rate was 26
W/kg, and the 2.45-GHz radiation was sinusoidally
modulated at 16, 30, 90, and 120 Hz. Enzyme activity
was not measurably affected by these exposures; a
10- to 15-percent change in enzyme activity would
have been required to detect a reliable microwave
effect.
Ismailov (1977) investigated  the infrared (IR)
absorption spectra  of  proteins in RBC  membranes
exposed to 1.009-GHz fields (SARs up to 45 W/kg)
and maintained at 25°C. The samples were exposed

-------
for 30 min in aqueous suspension in a stripline and
were then dried to a thin film to obtain the IR spectra.
No  change  in  cr-helix  or /3-sheet content of
the membrane proteins was noted. However, when
D2O (heavy water) was added  to the suspension
before beginning exposure, application of microwaves
was found to increase the degree to which strongly
bound amide hydrogens were exchanged. This effect
was pronounced at SAR =45 W/kg but disappeared
when the SAR was below 10 W/kg. The increase in
accessibility of the  poorly  exchangeable amide
hydrogens indicates that the microwaves disturb the
relationship  of  the  protein to  its  neighboring
membrane lipid.
IR spectra of Escherichia  coli after  exposure to
microwaves were also measured (Corelli et al. 1977).
After 12 h of exposure to 3.2-GHz (CW) radiation (SAR
= 20 W/kg), £. coli were dried to a film, and spectra
were measured  in the protein and  nucleic acid
absorption regions. No differences were found. These
results are comparable with those of Ismailov (1977),
which indicated no changes, but  they do not address
the conditions for which Ismailov did report effects.

The functional  properties of the  microtubule
assembly system extracted  from rabbit brain cells
were studied  after exposure to 3.1-GHz fields by
Paulsson  et  al.  (1977).  The binding  of the  drug
                      colchicine to the  microtubule precursor  protein
                      tubulin  was measured  after exposure  to PW
                      microwaves for  15  min at average dose rates of 112
                      and 243 W/kg (pulse-repetition rate of 200 Hz, pulse
                      duration of 1.4  /js). Colchicine normally blocks the
                      formation of microtubules, which halts cell division.
                      The normal assembly of microtubules from tubulin
                      exposed for 10 min to PW microwaves, as above, at
                      430 W/kg was also studied. No noticeable effect on
                      either process was found. The data indicate that a
                      change of about 15 percent in the colchicine binding
                      and about 10 percent in the microtubule assembly
                      measurement would have been noted. Paulsson eta/.
                      also studied the migration of proteins within the axonal
                      membrane of the rabbit's vagus nerve. In this case the
                      samples were exposed for 24 h (the SAR estimated
                      from  their data was 10 to 100 W/kg) at a pulse-
                      repetition rate (PRR) of 100 Hz and a pulse duration 1.4
                      /js. The distribution of tritium-labeled protein in the
                      axonal  membrane  was  found to be the same  in
                      exposed and control samples. A difference of > 20
                      percent would have been required for detection in this
                      experiment.
                      Two papers (Elder and AM 1975; Elder et al. 1976)
                      present results of exposure of rat liver mitochondria
                      to microwave radiation. Both papers examine oxygen
                      utilization  by measuring respiratory activity (e.g.,
                      respiratory control ratio, ADP to oxygen ratio) under
Table 5-3.   Summary of Studies Concerning RF-Radiation Effects on Subcellular Systems

                                                       Exposure Conditions
End Point Measured/
Effects
                                Exposure
Experimental    Frequency   Duration  Facility     SAR
System         (GHz)      (min)    (type)       (W/kg)    Reference
Increase in exchange of strongly
bound amide hydrogens in membrane
protein measured by IRt spectra
for SAR > 10 W/kg, no change in
cr-helix or /3-sheet content of
proteins

No change in activity of membrane
bound enzymes measured
spectrophotometrically

No change in activity of membrane-
bound enzyme measured  spectro-
photometrically
RBC membrane*  1.0 (CW)
          30
Stripline    5-45
RBC membrane.
mitochondrial
inner membrane

Endoplasmic
reticulum
2.45 (SWH)  10
2.45 (CW)
Waveguide



Waveguide
          26
42
          Ismailov (1977)
          Allis and Fromme (1979)
Vvardef a/. (1975)
No difference in respiratory
activity
No difference in respiratory
activity
No change in formation of
microtubules
No change in migration of proteins
within axonal membrane
No changes in IR spectra of
proteins and nucleic acids in
E. coli exposed before drying
Mitochondria
Mitochondria
Tubulin
(rabbit brain)
Vagus nerve cell
Dried film of
E. co/i cells
2.45 (CW)
2-4 (Swept)
3.4 (CW)
3.1 (PW)
3.1 (PW)
3.2 (CW)
30 to
210
10
15
24 h
8, 10,
11 h
Anechoic
chamber
far field
Coaxial
airline
Far field
Far field
Waveguide
17.5,
87.5
1.6-2.3
41
112-430
-10-100
20
Elder and Ali (1975)
Elder et al. (1976)
Paulsson et al. (1977)
Paulsson et al. (1 977}
Corelli era/. (1977)
*RBC = Red blood cell.
tIR = Infrared.
#SW = Sine-wave modulated.
                                                                         5-5

-------
various conditions. These parameters are functional
indications  of  the energy production system  in
eukaryotic cells. The earlier paper tested mitochondria
kept at 0°C,  or inactive state, during exposure in the
far field at 2.45 GHz. The mitochondria! functions
were examined after exposure in the active state at
2^°^  CFxnosures for periods up to  3.5  h were
conducted at SARs  of  17.5 and  87.5 W/kg.) No
rhsoooc; jn mi*ochondrial activity were seen. In the
laiLi paper,  the disadvantage of exposing  inactive
mitochondria was overcome by use of a novel flow-
through system that coupled a coaxial airline and an
oxygen electrode. The mitochondria! suspension was
cycled continuously between  the airline,  where
exposure was  accomplished, and the  oxygen
electrode.  Samples were exposed to 2.45-, 3.0-, and
3.4-GHz (CW) radiation (SAR = 41 W/kg) and also to
radiation at swept frequencies between 2 and 4 GHz
(SARs from 1.6 to 2.3 W/kg). As in the earlier work,
no  effects of  microwave exposure were detected
under  any condition. In general, a five-percent
change would  have  been sufficient for detection.
Other workers  have  presented data on microwave
exposure  of mitochondria,  either in  a form too
incomplete for  inclusion or  in oral presentations.
Their findings do not differ from those described here.

5.1.3 Effects on Single Cells

Effects on single cells have been  investigated by
several researchers (Table 5-4). Most  of the work
discussed  in this section falls into three classes:
studies of  ion transport into or out of red blood cells,
growth or colony-forming ability of various lines  of
cells, and responses of single neurons exposed to RF
radiation. In the ion transport studies two papers have
demonstrated effects, but one of them has not been
replicated  despite two attempts. Only one  of the
reports involving cell growth  reported an  effect,
but sample heating may have been responsible for the
result. The potentially most significant effect is that of
the response of neurons. Two papers from one labora-
tory are discussed in this section, and a third report
from a second  laboratory is discussed under Unre-
solved Issues (Sec. 5.1.4). All show very similar results
for the firing rate of neurons exposed to RF radiation. In
sum, for cellular systems, effects have been found
that may prove  significant and should provide leads
for elucidating a new mechanism of action. However,
at present, these effects require additional documen-
tation.

Transport and related properties of RBC membranes
have been studied by five groups. In each case Na+ or
K+ transport  was used as an  end point. In the RBC,
active (or  energy requiring)  Na* and  K* transport
across  the membrane is  by the  enzyme  Na+-K+
ATPase (the  same enzyme discussed in Sec. 5.1.2,
Effects on Subcellular  Organelles),  and passive
transport  is  through channels in  the membrane.
Ismailov (1971) exposed human RBCs to 1.0-GHz (CW)
microwaves at 45 W/kg and  found an increased
efflux of K* and a concomitant increased influx of Na*.
The Na+ influx was twice as large, ion per ion, as the
K* efflux. Exposures were carried out in a coaxial
stripline for 30 min, with analysis of the ion content of
the supernatant performed afterwards. These results
indicate either a reversal of the normal  action of the
Na+-K+ ATPase,  or an  inhibition of  the enzyme,
which permitted ion leakage across the membrane to
change the NaVK* ratio. Hamrick and Zinkl (1975)
and Peterson et  at. (1979) have performed similar
experiments by exposing RBC's at 2.45 GHz in the far
field (SAR's were 3 to 57 W/kg, and ~ 200 W/kg,
respectively).  Neither study found  a difference
between the K+  efflux  from microwave-exposed
RBC's and conventionally heated RBCs with similar
histories of temperature elevation, although Peterson
et al. did find a difference between unexposed rabbit
cells maintained at 25°C compared with those
maintained at  37°C. This difference did not occur
with  human cells. Ismailov also used  temperature
controls but did not describe the solution parameters
for the cell  suspensions.  Under certain conditions of
chemical concentration, it is possible to reverse the
Na+-K*  ATPase; however,  Ismailov's  controls
behaved normally,  which  wou.ld  indicate that
chemical concentrations in the controls were  not
unusual. The origin of the discrepancy between these
studies is not clear.
Hamrick and Zinkl (1975) and Peterson et al. (1979)
measured other end points as well. The former looked
at osmotic fragility of the RBC's and concluded that
there was no difference.  The latter paper gives data
on hemoglobin release from RBC's, an indicator of
membrane fragility. Again, no differences were found
between irradiated and heat-treated RBCs; however,
as for K+ efflux, the unexposed rabbit cells released
less hemoglobin at 25°C than at 37°C.
Liu et al. (1979)  also studied K+ efflux, hemoglobin
release, and  osmotic fragility in red  blood cells.
Exposure to RF radiation  was conducted in  a
waveguide  system at frequencies of 2.45, 3.00, and
3.95 GHz.  A  1.2-cm-diameter polystyrene tube
containing a 0.6-ml suspension of RBCs was inserted
through the center of the waveguide with the long
axis  of the tube parallel to the electric field.
Temperature  was measured before and after
exposure by insertion of a thermistor. The temperature
of the exposed sample was allowed to rise, and heat-
treated controls were placed in a water bath at the
temperature equivalent to the maximum exposure
temperature. Time of treatment was the same for
exposed and heat-treated samples. No statistically
significant differences were found between exposed
and heat-treated samples in any of the measurements
performed.  Experiments included exposure of rabbit
RBCs at SARs of 22 to 200 W/kg (equilavent to 5.2 to
22°C rise in temperature); ouabain treatment of the
                       5-6

-------
rabbit RBCs  to  block the  Na*-K* ATPase;  and a
comparative study of rabbit, human, and dog RBCs
exposed at 3.00 GHz and an SAR of 173 W/kg.
In a  separate paper,  Ismailov (1978)  reported
increases in the electrophoretic  mobility of human
RBCs exposed under conditions identical to those in
his previously discussed study. The electrophoretic
mobility was measured at  10-mm  intervals after
cessation of exposure. The mobility was found to peak
30 min after exposure and to return to base line ~ 60
min after exposure. The peak mobility decreased with
shorter exposure  durations (30,  15, 8, and 4 min).
Also, the mobility change decreased as a function of
dose rate and disappeared altogether between 5 and
10 W/kg. Although a  change in  the  counter-ion
                         distribution around the cell and possible conforma-
                         tional changes  in the membrane proteins  were
                         discussed  or suggested as possible causes, it
                         remained unclear why these phenomena peaked 30
                         min after exposure.
                         Passive ion transport was examined by Olcerst et al.
                         (1980) after they exposed rabbit RBCs to 2.45-GHz
                         (CW) radiation (SARs at 100, 190, and 390 W/kg).
                         The cells were treated wth ouabain to inhibit active
                         transport of  Na+  and  K* by  Na*-K*  ATPase,  the
                         important enzyme discussed previously in several
                         papers. Exposures took place in a waveguide system
                         in which the sample was placed parallel to the E-field
                         in  a cylindrical  tube. An organic  coolant of low
                         dielectric constant was circulated around the sample
Table 5-4.    Summary of Studies Concerning RF-Radiation Effects on Single Cells

                                                          Exposure Conditions
End Point Measured/
Effects
Increase in RBC electrophoretic
mobility 30 min post-exposure
(SAR > 10 W/kg)
Increase in K* efflux and Na* influx
K* transport no different from
heat-treated controls; no change
in osmotic fragility
K' transport no different from
controls at corresponding
temperatures; no difference in
hemoglobin release
Experimental
System
RBC*
RBC
RBC
RBC
Frequency
(GHz)
1 .0 (CW)
1 .0 (CW)
2.45 (CW)
2.45 (CW)
Duration
(min)
4.8,
15, 30
30
60, 120,
1 80, 240
45
Exposure
Facility
(type)
Stripline
Stripline
Monopole
far field
Anechoic
chamber
far field
SAR
(W/kg)
5-45
45
3-57
200
Reference
Ismailov (1 978)
Ismailov (1971)
Hamrick and Zinkl (1975)
Peterson et al. ( 1 979)
Passive transport of Na* and Rb*
increased at transition temperature

No significant changes in K* efflux,
hemoglobin release, or osmotic
fragility

Rapid response in change of firing
rate of pacemaker neurons which
does not correlate with tempera-
ture changes in minority of trials
No change in growth or CFUt of
exposed cultures

No change in growth, CPU, of vari-
ous strains of exposed cultures
under several growth conditions

No change in survival curves
(measuring CPU) of exposed
cultures

Growth rate slowed; morphological
changes found

No change in light emission of
photoactive  bacterium

No effect on colony-forming
ability

Temporary decrease in virulence
(> 6 h) of bacteria for its host
cells; recovery within 24 h at
37 °C
RBC
RBC
2.45 (CW)


2.45 (CW)
3.00 (CW)
3.95 (CW)
60       Waveguide   100,190  Olcerst  et al. (1980)
                    390
20.180   Waveguide   22-200   Liu et al. (1979)
Isolated neuron    1.5,2.45      3
from Aplysia      (CW and PW)
E. co/i
P. aeruginosa

E. coli
2.45 (CW)
2.45 (CW)
            720
                            240
E. coli           2 45         1
B. subtilis spores

Chinese hamster   2 45 (CW)     20
lung cells, V79

P. tischeri        2.6-3.0 (CW)   -22
                     Stripline
         Par field


         Anechoic
         chamber
         far field

         Microwave
         oven
1-100    Wachtel et al. (1975);
         Seaman and Wachtel
         (1978)

29-320   Hamrick and Butler
         (1973)

0.0075-   Blackman et al. (1975)
75

—400     Goldblith and Wang
         (1967)
E. coli
2.6-4.0 (CW)  8 h
A. tumefaciens    10 (CW)      30,60,
                            230
Waveguide
Waveguide
Waveguide
Cavity
1059
660 to
5300
29
-1
Chen and Lin (197£
Barber (1962)
Corelli et al. (1 977)
Moore et al. (1979)
*RBC = Red blood cell.
tCFU = Colony forming unit.
                                                                            5-7

-------
in a  larger concentric cylinder  to  maintain the
temperature of the sample under exposure. The SARs
were computed from readings of forward and reflected
power. The RBCs were pre-incubated with radioactive
Na+ or Rb+.  (The latter is a K+ substitute.) Samples
were  exposed or heat-treated for  1  h, and the
suspending  medium was analyzed for radioactivity.
Graphs of the  logarithm of the efflux vs.  inverse
temperature were identical except at three transition
temperatures, where the slope of these plots changes
sharply (Figure 5-1). At these points, the exposed
samples exhibited considerably higher efflux than the
heated samples; however, no consistent difference
between exposure levels could be established. These
results  imply  that  the  cell membrane structures
responsible for passive ion transport are sensitive to
microwave  exposure at temperatures at which
transitions  between two states are taking place.
Living  cells depend  on closely regulated ion
concentrations for  many processes, and serious
disruption of these balances could be lethal. It should
be noted  that none of the previously discussed
studies of Na+or K+transport in RBCs were conducted
at the temperatures at which transitions were found in
this report.
Many studies focus  on a major function of a  cellular
species as a generalized end point.  The reasoning for
this  approach  is that  if  microwaves disrupt an
important metabolic step, the result will be a decline
in the ability of the cell to perform its major function.
As mentioned earlier, this reasoning assumes that no
compensating  mechanism will operate. A broad
spectrum of such end points has been investigated for
cells exposed in vitro.  Some of them  are discussed
elsewhere: phagocytosis and bfastic transformation in
lymphocytes in Sec. 5.2, Hematologicandlmmunologic
Effects; and the induction of the  antibiotic  colicin,
along with point mutations in single cells, in Sec. 5.8,
Genetics and Mutagenesis.
Perhaps the ultimate test of a cell's functional ability
is growth  and survival. Several workers  have
concentrated on this end point and have investigated
the frequency range between 1 and 4 GHz. In general,
the results have proved negative. Far-field exposures
were conducted (Hamrick and Butler 1973; Blackman
etal. 1975)onseveralstrainsormutantsof£.  co//and
on Pseudomonas aeruginosa. Samples were exposed
in T-flasks or  petri  dishes, principally to 2.45-GHz
(CW)  radiation. Growth was measured by assays of
colony-forming units (CFUs). In both experiments the
duration of exposure was sufficiently  long (12 and 4
h) for the average cell to divide at  least once.
Blackman et al. examined several growth conditions
such as lag,  log and  stationary phases  of growth, rich
and minimal media, and a "normal" as well as a
mutant amino-acid-requiring strain. In each case, no
differences  between the  exposed  samples  and
temperature-matched  control samples were found.
Hamrick's SARs ranged from 29 to 320 W/kg with
Figure 5-1.
Arrhanius plot of Na* efflux. Unirradiated tempera-
ature controls are represented by open symbols;
irradiated samples are represented by darkened
symbols. Specific absorption rates are expressed
as watts per kilogram (e = 100. •= 190, A = 390).
Control samples had an average standard error of
0.98 percent. Irradiated samples had an average
standard error of 4 percent (Olcerst et al. 1980).
     39.3
    29.8
Temperature, °C
    20.9      12.5
     3.2
                                 3.6
power densities  of  60  to  600 mW/cm2,  and
Blackman's from 0.0075  to 75  W/kg with power
densities of 0.005 to 50 mW/cm2.
Corelli et al. (1977) exposed f. coli at the end of a
waveguide to microwaves swept from 2.6 to 4.0 GHz
for 8 h at a dose rate of 29 W/kg. No effect of the
microwave  exposure was  found  on colony-forming
ability of the bacteria. Goldblith and Wang (1967)
exposed  E. coli and  Bacillus  subtil is spores  in  a
microwave oven at 2.45 GHz for periods to 1 min(SAR
estimated at 400  W/kg).  In this case,  microwave
irradiation and  conventional heating were found to
have identical effects on survival.

Chen and Lin (1978) exposed Chinese hamster lung
cells, V79, in a  waveguide fitted with a micropipette
that contained the cell  suspension. Temperature was
regulated by circulation of cooling water through the
waveguide around the micropipette. Samples were
exposed to 2450-MHz  (CW) radiation for 20 min, and
the cells were allowed to grow in cultures for 12 days
after exposure.  The exposed cells were observed to
divide at a lower rate and to exhibit a fibroblastic type
                       5-8

-------
of growth, in contrast with the controls. These cells
were exposed at 400 mW/cm2 (SAR = 1059 W/kg)
under conditions similar to those in which Harrison et
al. (1980) found temperature elevations in the sample
that were as much as 0.3°C higher than the cells in
the cooling  bath. Chen  and Lin (1978) state that
temperature-treated controls at 38°C (1 °C higher than
the coolant temperature during microwave exposure)
did  not  display  the changes  observed in  the
microwave-exposed cells. However, it is not entirely
clear whether the changes in the microwave-treated
cells were caused by elevated temperatures within
the micropipette.

The emission of  light by a photoactive bacterium,
Photobacterium fischeri, has been used as the end
point of one  study (Barber 1962), where the bacterial
suspension  was  circulated through a  waveguide.
The bacteria were undergoing exposure  in the
waveguide for approximately half the time during a
typical 43-min experiment. Bacteria were exposed at
several -frequencies between 2.6 and 3.0 GHz; the
assay was performed 24 h later. In spite of extremely
high dose rates, 660 to 5300 W/kg, there were no
differences between microwave-irradiated and
conventionally heated samples that received parallel
treatment.

In the single experiment conducted  at 10-GHz
frequency,   a transient  effect was found in  the
virulence of  Agrobacterium tumefaciens towards its
normal hosts, potato and  turnip disks (Moore et al.
1979). This bacterium produces a plasmid, which it
injects into the host cells and which is responsible for
turning these cells into uncontrolled tumor cells. In
this experiment, a suspension of A. tumefaciens was
exposed in a petri dish for 30, 60, and 230 min. The
longer exposures  produced a decrease of virulence
near 60  percent  with no essential change in the
number of viable cells. The effect was unchanged 6 h
post-exposure, but virulence returned to normal 23 h
post-exposure when the bacteria were maintained at
27°C. An additional experiment, in which treated and
untreated cells were added to the host simultaneously,
indicated that the exposed cells were able to compete
effectively with  untreated cells for binding sites on
the host cells.  The treated  bacteria -were always
maintained at or below 27°C during irradiation. From
the temperature data in the paper, a dose rate of ~ 1
W/kg and power  density of 0.58 mW/cm2 can  be
estimated. A possible explanation  for  this  effect,
which was not offered by the original authors, is that
the  plasmid DNA of  the A. tumefaciens  was
incorporated into the major DNA of the bacteria during
microwave exposure, preventing  injection into the
host. Normal growth may have subsequently allowed
the plasmid to return to its original state and activity.

In two papers, Wachtel and co-workers (Seaman and
Wachtel 1978, Wachtel et al. 1975) have examined the
effects of microwave irradiation on the firing rate of
 isolated neurons from the marine gastropod Aplysia.
 Neurons were exposed in a stripline to 1.5- and 2.45-
 GHzfCW and PW) fields (0.5- to 10-//S duration, 1000
 to 15,000 pulses/s).  The  firing rate of pacemaker
 neurons and the burst  rate of bursting  cells were
 measured during  microwave exposure at dose rates
 between 1 and 100 W/kg. Glass capillary electrodes
 filled with 0.5 M  KCI were used to inject  current
 and measure the firing  rate of the cells.  The
 electrodes had artifactual DC potentials of less than 1
 mV and currents of less than 10 pA during exposure;
 these values were judged by the authors to be too
 small to affect the performance of the cells. In the
 majority of cases,  the firing rate of pacemaker cells
 increased  with an increase in temperature,  and
 decreased with  a decrease  in temperature. In  a
 minority of cases, 13 percent, for the pacemaker cells, the
 microwave irradiation reversed the normal  change
 in firing rate; i.e., the rate decreased or stopped with a
 microwave-induced increase in temperature.  The
 authors were able  to detect slow and rapid
 components. The slow component, occuring in 30 to
 60 s, was correlated with the slow rise of temperature
 associated with  exposure. The  rapid component,
 occurring within 1 s, appeared to correlate with the
 presence of the microwave field. The rapid component
 was always found to be a decrease in firing rate in the
 presence  of the field and was never produced by
 convective heating. Similar but more variable effects
 were found for the bursting cells. The threshold for
 the slow component was ~ 7 W/kg, but in one case
 the rapid component was found at an SAR as low as 1
 W/kg. For all cases in which effects were found, the
 firing rates returned  to  normal when the radiation
 was terminated  and  when  the temperature was
 returned to normal. The authors hypothesized that at
 a dose rate of ~ 1 W/kg, conversion of 0.1 percent of
 the  microwave energy into  a polarizing current
 density across the cell  membrane would be sufficient
 to affect the firing  rate of pacemaker neurons.

5.1.4 Unresolved Issues
Several  issues remain unresolved in the area of
cellular  and subcellular effects of microwave
radiation. Nearly  all  of the relevant research is
concentrated in a  narrow frequency band, between
 1.0  and 4.0 GHz.  No acceptable studies have been
reported for a large portion of the frequency spectrum
of concern, 0.5 MHz to 100 GHz.

An active  area of research at present is concerned
with the role of structured water in the cell. Questions
such as how much water in a cell is structured to a
greater degree than "bulk" water, and whether this
structure plays an important role in cell metabolism,
are  as yet unanswered.  Three monographs (Alfsen
and Berteaud 1976; Drost-Hansen and Clegg 1979;
Grant eta/. 1978) summarize knowledge in this area.
Experiments attempting  to establish the presence
and  extent of structured water within biological
                                                                       5-9

-------
systems  are described,  and the  dielectric data
indicating a possible frequency range for the
structured-water resonance are presented. According
to the limited information now available, RF radiation
is most effective in modifying the state of structured
water at frequencies below 1 GHz (Grant etal. 1978,
pp.  160-165).  As yet no  experiments  have been
conducted that define whether absorption  of RF
energy by structured water leads to a measurable
change in a biological system.
Three of the reports discussed in Sec. 5.1.3, Effects
on Single Cells, also raise questions. The explanation
offered by  Ismailov  (1978) for  the  change in
electrophoretic mobility of exposed RBCs is specula-
tive and  does  not account for the peaking  of the
phenomenon 30 min post-exposure. These results
must be considered an effect of unknown origin until
more information  is available.  The  results of
experiments by Moore et al. (1979), in which the
virulence of A.  tumefaciens was decreased for more
than 6  h after exposure to microwave radiation,
suggest  reversible functional changes in the
organism. In this case, the implications for cellular
function after microwave exposure would be broad.
However, no other worker has noted a similar effect
with other single-cell organisms, and this experiment
has not been independently confirmed. Therefore, its
significance is  unknown at this time.

The results of Wachtel  and co-workers (Seaman and
Wachtel 1978; Wachtel et al.  1975) are potentially
highly significant because they indicate the possibility
of a direct interaction between the microwave field
and the functioning of the pacemaker neuron (i.e., the
rapid  effect). Three other workers have  obtained
supportive results. The documentation in  these
reports is not sufficiently complete for inclusion in the
preceding sections, but they bear mentioning here.

Yamaura and Chichibu (1967) found results strikingly
similar to those of Wachtel in ganglia of crayfish and
prawn exposed at 11 GHz. The regular firing rate of
the  ganglia  decreased rapidly during microwave
exposure,  rebounded to a higher than  normal rate
when the radiation was removed, and then returned
to normal. Temperature controls showed only an
increased firing rate as the temperature was
increased. The authors stated that the SAR was ~
100  W/kg but did not describe the  method of
measurement.

Arber (1976) found a hyperpolarization of the resting
potential  of giant neurons of the  mollusk  Helix
pomatia during exposure to 2.45-GHz(CW) radiation.
In this experiment, the ganglion was isolated and
mounted  in a stripline. The cell potential was
measured, as in Wachtel's experiment, by insertion of
microelectrodes into the neuron. Exposure to
microwaves (SAR near 15 W/kg) for 1 h produced a 5-
to 10-percent increase in resting potential, followed
by a stabilization or slight additional increase over 1 h
post-exposure. This result is presumably caused by a
change in the Na+-K+ balance in the cell. When Arber
treated the cells with ouabain (post-exposure), which
inhibits  Na*-K* ATPase, he found that part of the
hyperpolarization could be accounted for by the
action of this enzyme  under the  influence  of
microwaves.  The remainder was attributed  to
changes in passive ion transport.

Pickard and Barsoum (1981) have recently presented
results in which single cells from Chara braunii and
Nitella flexilis, plants from the family Characeae,
exhibited a  large step increase  in voltage when
exposed to 0.1 - to 5-MHz (PW) radiation (pulses were
of 250-ms duration, pulse interval was 6.3 s). These
authors also found a fast and slow component. The
slow component correlated with a temperature rise in
the sample;  the  fast component was frequency
dependent and disappeared abruptly at ~ 10 MHz.
The authors suggest that the fast component was
produced by rectification  of the oscillating electric
field by the cell  membrane.  The fast component
disappeared into noise at ~ 667 V/m and may have
been an effect of intense fields.

The results of these four studies indicate a possibility
of a direct microwave interaction  with the electric
potential across the cellular membrane of all living
cells.  This  kind  of  interaction would have broad
significance  in the functioning of all cells and,  in
particular, cells of the nervous system.

Additional studies have been conducted recently that
have not yet been sufficiently documented to include
in the previous sections. However, because of their
potential importance, they are discussed here. The
first two reports concern microwave effects  on
phospholipid bilayers, whereas the third is concerned
with changes in growth of the yeast Saccharomyces
cerevisiae exposed at frequencies between 41 and 42
GHz.
Tyazhelov et al. (1979a) have exposed a phospholipid
membrane formed between two chambers containing
solutions of NaCI or KCI. An antibiotic that forms
pores through the membrane was added to facilitate
passage of Na+ or K+. Conductance was measured
across the membrane while 4-s pulses of 900-MHz
radiation were delivered  at between  125 and 280
V/m  (field  strength in  the aqueous medium). The
results  showed  a change in conductance under
exposure that is consistent with temperature rises of
12°C, but  the temperature  of  the  NaCI or KCI
solutions did  not vary by more than 0.5°C. However,
insufficient information concerning  the  exposure
system  makes it difficult to judge whether serious
inhomogeneities in  energy deposition, and thus
similar inhomogeneities in the temperature distribu-
tion, were likely. One cannot estimate an SAR from
the information presented.
                       5-10

-------
Sheridan et al. (1979) have presented at meetings,
but so far have not published,  results of  Raman
spectroscopy of single- and multilamellar phospholipid
vesicles exposed to 2.45-GHz (CW)  radiation. No
change was found in  the Raman bands of  single-
layered vesicles. In contrast, the data for multilamellar
vesicles indicate that the  hydrocarbon  tails of the
phospholipids were undergoing a  temperature-
dependent phase transition at a point at which the
bulk temperature was too low for the transition to
have begun.  The change was reported to be
equivalent to a temperature difference of ~ 2°C at an
exposure of 25 mW/cm2. The bulk temperature of the
sample in these experiments was measured  by  a
unique method. Small ruby crystals were suspended
in the sample, and the shifts in the Raman bands of
the crystal were measured  and compared  with
calibration curves. This  appears  to be an accurate
method for temperature determination. The origin of
this effect is  unknown, but if a similar  effect were
found in  naturally occurring membranes, it could
have an impact on the functioning of the biological
membrane.

Keilmann  and co-workers (Grundler et al.  1977;
Keilmann 1978; Grundler and Keilmann  1980) have
demonstrated that S. cerevisiae exhibits an enhanced
or inhibited growth  rate when exposed at certain
closely spaced frequencies between 41.60 and 41.80
GHz. For  instance, they found  a  10- to  15-percent
increase in growth rate at 41.64 and 41.68 GHz, and a
20-percent decrease at 41.66 GHz. The experiments
were conducted with a unique waveguide  termination
that was dipped into a suspension of yeast cells. In a
typical experiment, 24 W was dissipated in the yeast-
cell  suspension, and the authors  estimated  a
maximum  exposure intensity of about 10 mW/cm2.
However,  because of the  unusual  nature  of  the
waveguide termination and the high attenuation of
high-frequency radiation by aqueous samples, SAR
values cannot be determined. Sample temperature
was monitored and was within  0.5°C of  the desired
32°C. Equivalent temperature  controls were per-
formed, and the authors believe that changes of the
observed magnitude could not be purely temperature
effects. This contention  appears  reasonable, since
the growth rate both increased and decreased at the
same levels  of incident energy but at different
exposure frequencies.  The decrease  is difficult to
explain,  based  on the authors' reports of no
observable temperature rise. However, if a localized
temperature > 37°C (> 5°C  above the controlled
temperature) were attained close to the waveguide
termination, then the decrease could be explained
solely on the basis of temperature. There are scanty
reports from the Soviet literature that present results
similar to those of Keilmann and co-workers. These
studies are important because they are suggestive of
a specific interaction  mechanism  between RF
radiation and  a biological system.
 In summary, the available literature on cellular and
 subcellular effects  of  microwaves does not  yet
 definitely establish  whether effects unrelated to
 elevation of temperature exist at dose rates on  the
 order of 1 W/kg. Several investigators have reported
 effects,  but a majority have not  found effects
 unrelated to temperature variations. In some cases,
 the results conflict. In other cases, the effects found
 are equivocal. Effects of elevated temperature may
 not  be clearly eliminated, or,  as in the case of
 neuronal firing rate, the change in the rate may occur
 only in a minority of cases (Seaman and Wachtel
 1978). The question of microwave-induced tempera-
 ture rises must always be carefully considered when
 one is dealing with the biological effects of exposure
 to  microwave radiation. In cellular  systems, a
 principal effect is  to  increase the rates  of  all
 biochemical reactions, including the rate at which
 denaturation of proteins and DNA occurs (Lehninger
 1975), at which bases are removed from DNA(Lindahl
 and Nyberg 1974), and at which mutations occur in
 DNA  (Bingham et al. 1976). When temperature is
 raised to a certain level, often ~ 43°C, cell functions
 become so disrupted  that the cell's capacity to repair
 the damage is exceeded and the cell dies. The effect of
 a rapid heating rate is less clear. Some disruption of
 cefl function may be expected from a  rapid
 temperature  rise,  even if the critical temperature
 (e.g.,  43°C) has not been reached. Whether this
 disruption is fully reversible has not been well docu-
 mented.  RF-radiation exposure can  produce  such
 high heating rates, especially from high-peak, low-
 average  power pulses.  Effect  of nonuniform heat
 deposition  is probably the  most difficult aspect to
 account  for when  one is evaluating the effects of
 exposure of RF radiation.

The effects presented in this section are sufficiently
well established to warrant continued concern and
effort. The  effects of microwave radiation on the
electrical properties of the  cell  are potentially the
most significant. Four  separate experiments have
demonstrated these effects. All cells use the electrical
potential  across  the cell membrane in  their life
functions, and perhaps the most important cells are
those of the nervous system. Microwave effects have
been found in the functioning of the nervous system
and in behavior, as will be discussed in the following
sections.

At this time,  nearly all the  effects documented  for
cellular and subcellular systems are observed  in
intact cells. This finding could imply that a living cell is
required for the necessary interaction with microwave
radiation to occur;  or,  it may be that the right
questions are not being asked about the subcellular
level,  perhaps because the  levels of understanding
and  instrumentation capabilities  are  as yet too
limited. Although  the  primary mechanism(s)  of
interaction  other than  heating  the water medium
                                                                      5-11

-------
have not yet been defined, some useful directions are
indicated by the  results  reviewed here. One might
also conclude that some of the effects noted at the
cellular level,  particularly the changes in nerve cell
function, may be correlated with effects at higher
levels of organization.
                        5-12

-------
5.2 Hematologic and Immunologic
Effects
Ralph J. Smialowicz

Over the years, a considerable number of reports has
appeared in the literature dealing with the effects of
RF  radiation on  the  hematologic and immunologic
systems of animals. For the most part, the responsible
investigators have been motivated by a concern for
the possible adverse health effects of exposure to RF
radiation. Studies in which animals have been
exposed at various frequencies and intensities have
shown  inconsistent  changes in  elements of both
biological  systems.  In  some instances,  a thermal
burden to the exposed animal has been credited with
the observed  changes,  whereas in  others,  a
"nonthermal" (i.e., lack of measurable elevation of
temperature) or direct (i.e., athermal or field-specific
effect) interaction of RF radiation with the blood  and
blood-forming systems  has been suggested as  the
causative mechanism for the observed effects. In any
case, the final interpretation of RF-induced changes
must consider  many  variables that  affect  the
interaction of RF radiation with the biological entity.
As has been noted previously, variables such as body
shape and mass, radiation frequency, duration of
exposure,  field  intensity, specific absorption rate,
energy distribution, orientation of the body  in the field,
ambient environmental  conditions, area of the body
exposed, and field modulation may all influence  the
final results. Variability of response among species
and strains as well as between sexes must also be
considered.
   been  examined, the effects have proven to be
   transient.
 • Some of the reported effects of RF radiation on the
   hematologic and immune systems are similar to
   those (a) resulting from a stress response involving
   the hypothalamic-hypophyseal-adrenal axis, or(b)
   following administration of glucocorticoids.
 • Several reports show an association between RF-
   induced thermal  loading or increased  core
   temperature and hematologic and  immunologic
   changes. Conversely, there is a lack of convincing
   evidence for a direct interaction of RF radiation
   with hematologic and immunologic systems in the
   absence of some form of thermal involvement.
 • That an increase  in rectal temperature is not
   observed after exposure to RF radiation does not
   preclude a thermal interaction that the animal is
   able to compensate for  and control.
 • There is presently  no convincing evidence from
   animal studies for adverse alterations in the
   hematopoietic or immune systems at RF-radiation
   intensities comparable to average environmental
   levels, i.e., 0.01 to  0.1 /uW/cm2 in the frequency
   range of 54 to 900  MHz, which encompasses the
   resonant frequencies for human beings.
For convenience, this section is divided into  two
general topics: hematologic effects and  immunologic
effects. These two topics are further subdivided into
reviews of studies in which cellular components of
these systems  have been exposed in vitro and of
studies dealing with in vivo exposures to RF radiation.
This section critically reviews the reported effects of
RF  radiation on the  hematologic and immunologic
systems of laboratory animals. On the whole, direct
comparisons between many of these studies are
difficult to make because of differences in species, in
exposure parameters, and in the biological end points
examined.  Furthermore, as mentioned above, the
reported responses  are inconsistent and highly
variable, with general trends in the data not readily
apparent.  Nevertheless, from this review  the
following generalizations may be fo/mulated:

• Partial or whole-body  exposure of animals to RF
  radiation may lead to a variety of changes in the
  hematologic and immunologic systems. Depending
  upon the exposure  conditions, species,  and
  parameters measured,  the  changes  may be
  stimulatory or suppressive.

• Transient changes in peripheral blood composition,
  possibly caused  by redistribution of blood cells
  and hemoconcentration, have been reported.
• In  those cases  where the reversibility of RF
  radiation  effects on the immune  system have
 5.2.1 Hematology
 Hematology is the study of the anatomy, physiology,
 and pathology of the blood and blood-forming tissues.
 The hematopoietic system is comprised of a variety of
 cells and cell products. In fetal life, the production of
 blood cells  occurs in  the liver, spleen,  and bone
 marrow. After birth this function is limited largely to
 the bone marrow,  which  produces  red  cells
 (erythrocytes), white cells (neutrophilic, eosinophilic,
 and basophilic granulocytes; lymphocytes; and
 monocytes), and platelets. Each of these  cell types
 performs specific functions that are essential to life.
 For example, mature erythrocytes transport Oz and
 C02 to and from tissues, granulocytes and monocytes
 phagocytize invading microorganisms, and lympho-
 cytes are involved in immune responses.  These
 functional cells are all descendants of progenitors
 (stem cells) that reside within the bone marrow.
 Blood-cell formation consists of two  essential
 processes, proliferation and differentiation; bone
marrow progenitor cells proliferate and differentiate
into red  and  white  cells.  As  the process of
differentiation  progresses, the capacity for cellular
proliferation decreases. Impairment of either of these
                                                                      5-13

-------
 processes may lead to dysfunctions in the hematologic
 system that may be life threatening.

 5.2.1.1 In Vivo Studies

 A paucity of information on the health effects of RF-
 radiation exposure from clinical and epidemiological
 studies  has led to  studies of  effects  on the
 hematologic systems of laboratory animals (Table 5-
 5). Many of the early investigations on the blood-
 forming system of laboratory animals employed
 power densities of 10 mW/cm2  and higher. For
 example, Deichmann et al. (1964) reported significant
 leukocytosis, lymphocytosis, and neutrophilia in rats
 following 7 h  of exposure to 24,000-MHz (PW)
 microwaves at an  average  power density of 20
 mW/cm2  (SAR estimated at 3 W/kg). One  week
 following exposure, peripheral blood values returned
 to  normal. Rats exposed for 3 h at  10 mW/cm2
 displayed the same changes and returned to normal
 after 2 days (SAR estimated at 1.5 W/kg). Increases
 in circulating erythrocytes, hemoglobin concentration,
 and hematocrit were observed in two of three strains
 of  rats (Osborne-Mendel  and CFN) exposed  to
 24,000-MHz fields at 10 or 20 mW/cm2. However, in
 Fischer rats exposed under the same conditions,
 there was a reduction in  the number  of circulating
 erythrocytes  and a  reduction  in hematocrit and
 hemoglobin concentration.  These differences  in
 Fischer rats are difficult to reconcile in light of the
 hematologic responses to thermal loads displayed by
 the Osborne-Mendel and CFN rats. In another
 experiment, Deichmann et al.  (1963) exposed two
 dogs to 24,000-MHz (PW) fields at an average power
 density of 24 mW/cm2 (SAR estimated at  1 W/kg).
 One dog was exposed for 20 months, 6.7  h/day, 5
 days/week; the second dog  was exposed for 20
 months,  16.5  h/day, 4 days/week. No significant
 changes were observed in blood volume, hematocrit,
 hemoglobin,  erythrocytes, total  and differential
 leukocytes, blood cholesterol, or protein-bound
 iodine. The only finding attributed to  the exposure
 was a slight loss of body mass.

Kitsovskaya (1964) exposed rats to 3000-MHz (PW)
radiation at  10, 40, or 100 mW/cm2 for various
periods of time (SAR estimated at 2, 8, and 20 W/kg,
respectively). No changes were found in rats exposed
at 10 mW/cm2; at 40 and 100 mW/cm2, however, the
absolute number of peripheral blood  erythrocytes,
total leukocytes, and lymphocytes decreased,  and
granulocytes increased. These blood changes did not
return to normal until several months after cessation
of exposure.
The apparent discrepancy between the results of
Deichmann et al. (1964) and Kitsovskaya (1964) may
be partially explained by the work of Michaelson era/.
(1964). These  investigators  reported  that  the
hematopoietic effects of 2800- and 1280-MHz (PW)
fields depend on the  frequency,  intensity, and
duration of exposure. For example, dogs exposed to
2800-MHz fields showed  a  marked  decrease in
circulating lymphocytes and eosinophils after 6 h at
100 mW/cm2  (SAR estimated at 4  W/kg). This
exposure resulted in a 1 °C  mean increase of rectal
temperature. Neutrophils remained slightly increased
at 24 h; eosinophils and lymphocyte values returned
to normal levels. After a 2-h exposure at 165
mW/cm2 to 28,000-MHz fields (SAR estimated at 6
W/kg), there was a slight leukopenia,  neutropenia,
and  definite  hemoconcentration. These changes
were accompanied by a rectal temperature rise of
1.7°C. Eosinopenia was  still evident 24 h after this
exposure. Changes in leukocyte counts were more
apparent following exposure of dogs to 1280-MHz
(PW) fields  or  to  200-MHz (CW)  radiation. After
exposure of dogs at 1280 MHz for 6 h at 100 mW/cm2
(SAR estimated at 4.5 W/kg),  a leukocytosis and
neutrophilia were observed. After 24 h the neutrophil
level was still increased above pre-exposure levels.
Lymphocyte and  eosinophil values were slightly
depressed following exposure, but at 24 h they were
slightly higher than initial values. A 6-h exposure to
200-MHz (CW) fields at 165 mW/cm2 (SAR estimated
at 25 W/kg) caused a marked increase in neutrophils
and a slight decrease in lymphocytes. After 24 h this
trend was more evident. Michaelson  et al. (1964)
suggested that the results indicated a stress response
of the exposed  animals in the hypothalamic and/or
adrenal axis that  was brought about by a thermal
stimulation from RF-radiation exposure.

Spalding et al. (1971) exposed mice to 800-MHz fields
at an average power density of 43 mW/cm2 (SAR
estimated at 10.7 W/kg) for 2 h/day, 5 days/week,
for a total of 35  weeks. These investigators found no
changes in blood erythrocytes, leukocytes, hematocrit,
or hemoglobin concentrations. It is interesting that
these investigators  did  not detect  changes in the
peripheral blood picture of exposed mice, despite the
thermal  burden that was  being placed orr these
animals. Failure to detect changes  may have been
due to the animal's ability, over the prolonged period
of exposure, to adapt to the RF-induced thermal load.
Four mice died from "thermal  effects" following the
33rd and 34th RF-radiation exposures.

Effects produced at levels at or below  10 mW/cm2
(SAR estimated  at 0.5 to 2.0 W/kg) have also been
reported. For example,  Baranski (1971,  1972a,b)
exposed guinea pigs and rabbits to 3000-MHz (CW
and PW) microwaves at an average power density of
3.5 mW/cm2 (SAR  estimated at 0.5 W/kg) for 3
months,  3 h  daily. At this  power  level, the body
temperature  of the animals  was  not  elevated.
Observations  were made of increases in absolute
lymphocyte counts in peripheral blood, abnormalities
in nuclear structure, and mitosis in the erythroblastic
cell series in the bone marrow and in lymphoid cells in
lymph nodes and spleen.  No  changes were observed
                       5-14

-------
Table 5-5. Summary of Studies Concerning Hematologic Effects of RF-Radiation Exposure*
Exposure Conditions
Effects
Increased: WBC. lymphs, PMN,
RBC. Hct, and Hgb
No change
No change
Decreased: RBC, WBC and lymphs
Increased: PMN
Decreased: lymphs and eosin
Decreased: WBC, PMN, and eosin
Decreased: WBC, lymphs, and eosin
Increased: PMN
Decreased: lymphs
Increased: PMN
No change
Increased: lymphs and mitotic
index of lymphoid cells
Increased: RBC, Hct, and Hgb
No change
Increased: eosmophils
Increased: WBC, CPU
Decreased: 59Fe uptake
Accelerated recovery following
x-irradiation; increased
erythropoiesis and myelopoiesis
Accelerated recovery from
x-irradiation
Increased: PMN and RBC
Decreased: lymphs
Accelerated recovery from
x-irradiation
Increased: lymphs
Decreased: PMN
(Not reproduced consistently)
No change
No change
No change
Decreased: Hct, WBC, and lymphs
No change
Decreased: lymphs
Increased: PMN
Decrease in CPU for erythroid
and granulocyte-macrophage series
Reduction in CPU granulocyte-
macrophage precursors exposed
in vitro
Species
Rat
Dog
Rat
Dog
Mouse
Guinea pig
Rat
Rat
Rabbit
Mouse
Mouse
Dog
Chinese
hamster
Rat
(Perinatal
exposure)
Rat
(Perinatal
exposure)
Rat
(Perinatal
exposure)
Quail egg
Rat
(young)
Mouse
Mouse
Mouse
Mouse
Frequency
(MHz)
24.000 (PW)
24,000 (PW)
3,000 (PW)
2.800 (PW)
2.800 (PW)
1,280(PW)
200 (CW)
800
3.000 (CW or
PW)
2,400 (CW)
2,400 (CW)
2.450 (CW)
2,450 (CW)
2,450 (CW)
2.800 (PW)
2,450 (CW)
425 (CW)
2,450 (CW)
100(CW)
2.450 (CW)
2,736 (PW)
2,450 (CW)
26 (CW)
2,450 (CW)
2,450 (CW)
Intensity
(mW/cm2)
10
20
24
10
40
100
100
165
100
165
43
3.5
10
5
10
100
100
100
60
10
5
46
30
24.4
30
8610
15
60-1000
Duration
(days x min)
1 x 180
1 x420
400 x 400-900
21 6 x 60
20x15
6x5
1 x360
1 x 1 20
1 x360
1 x360
175 x 120
120 x 180
30 x 1 20
90x60
180 x 1380
1 x5
1 x5
1 x3600
1 x30
47 x 240
57 x 240
57 x 240
1 x 1440
35 x 240
22x30

9x30
1 x 15
SAR
(W/kg)
3.0t
1t
2t
8t
20t
4t
6t
4.5t
25t
12.9t
0.5t
2t
It
1.5
70t
70t
4t
28t
3-7
1-5
2-3
14
5-25t
22
13t
10
120-2000
References
Deichmann et al. (1964)
Deichmann et at. (1963)
Kitsovskaya (1964)
Michaelson et at. ( 1 964)
Spalding et a/. (1971)
Baranski(1971 and 1972a)
Djordjevic and Kolak (1973)
Djordjevic et at. (1977)
McReeefa/. (1980a)
Rotkovska and Vacek (1 975)
Rotkovska and Vacek (1 977)
Michaelson et at. (1963)
Lappenbusch era/. (1973)
Smialowicz et at. (1982)
Smialowicz et al. (1979a)
Smialowicz et al. (1981 a)
Hamrick and McRee (1975)
Pazderova-Vejlupkova and
Josifko(1979)
Smialowicz et al. (1979b)
Liburdy (1977)
Huang and Mold (1980)
Lin et al. (1979b)
*WBC = white blood cell, PMN =polymorphonuclear leukocytes, RBC = red blood cell, Hct =hematocrit, Hgb = hemoglobin, andCFU colony-
 forming unit.
tSAR estimated.
                                                                                   5-75

-------
in the granulocytic series in peripheral blood. Shifts in
peripheral blood cells were found to correlate with
changes in the cellularity of the spleen and lymph
nodes.  An increase in the mitotic index and in the
percentage of cells incorporating 3H-thymidine was
observed in the spleen and lymph nodes of exposed
animals.
Djordjevic and Kolak (1973) exposed Vats to 2400-
MHz (CW) fields at 10 mW/cm2 (SAP estimated at 2
W/kg) 2 h/day for 10 to 30 days. Body temperature in
rats exposed under these conditions increased by 1 °C
within the first 30 min of exposure and remained at
this level throughout the exposure period. Hematocrit,
hemoglobin  concentration, and circulating erythro-
cytes  increased during  the  30-day exposure.
Fluctuations in the various leukocyte populations
were also observed. The  authors suggested these
changes were caused by the thermal effect of
microwaves. In a more recent study,  Djordjevic et al.
(1977)  found  no  significant difference in any of
several hematologic end points for  rats exposed to
2400-MHz (CW) microwaves at 5  mW/cm2 (SAR
estimated at 1 W/kg)  for 1 h/day during a 90-day
period.
Recently, McRee et al. (1980a) reported significant
decreases in eosinophils and a lowering of albumin
and  calcium in  blood from rabbits immediately
following chronic exposure to 2450-MHz fields. In
this study, rabbits were exposed 23 h daily for 180
consecutive days to 2450-MHz (CW) radiation at a
power density  of 10 mW/cm2 (SAR = 1.5 W/kg). No
change in hematologic parameters was observed 30
days after termination of exposure (i.e., depression in
eosinophils  seen  immediately  following exposure
had normalized); however, a significant decrease in
albumin/total globulin ratio was observed in the blood
of exposed rabbits at this time. The authors contend
that, because only 3 of the  41 blood-chemistry
parameters  measured immediately  after exposure
were significantly different (p <0.05), and because
this observation is close to that expected by chance,
further validation of these changes is warranted.

Rotkovska and Vacek  (1975) reported changes in
hematopoietic  cell populations  of mice following a
single 5-min exposure to 2450-MHz (CW) radiation at
an intensity of 100 mW/cm2 (SAR estimated at 70
W/kg). The response  of  microwave-exposed mice
was compared with that of mice placed in a warm-air
chamber at  an ambient temperature of. 43°C for 5
min. Both  treatments caused a  rise in  rectal
temperature > 2°C. A  leukocvtosis occurred in mice
under both conditions; however, the time course for
the leukocvtosis and the response of hematopoietic
stem cells differed between the two treatments.
Following RF-radiation exposure, a decrease in the
total cell volume of the bone marrow and spleen was
observed, and the number of hematopoietic stem cells
in bone marrow and  spleen,  as measured  by the
colony-forming unit (CFU) assay, increased. Incorpor-
ation of 59Fe in the spleen decreased 24 h after RF-
radiation exposure. On the other hand, the exposure
to heat caused a decrease  in the CFU's in bone
marrow and spleen and an increase in the percentage
of 59Fe incorporation. Rotkovska and Vacek concluded
that the different effects  of RF radiation and
externally applied heat on the hematopoietic stem
cells indicate that biological effects caused by high
intensities of RF radiation may not necessarily be
related only to increases in internal temperature. They
indicated that  their results  suggest  a possible
"direct" effect. This study is significant because it
demonstrates a  marked difference in  the kinetic
response of the hematopoietic system to two forms of
heat stress. Consequently, these differences must be
considered in the interpretation of RF-radiation-
induced changes in the hematopoietic system.

Subsequently, Rotkovska and Vacek (1977) studied
the effect of microwaves on  the recovery of
hematopoietic  tissue  following exposure to x-
irradiation. Mice exposed to X rays at 300 to 750 rads
were then exposed to 2450-MHz (CW) microwaves
for  5  min at 100 mW/cm2  {SAR estimated at  70
W/kg).  The combined  treatment resulted in  an
accelerated recovery of hematopoietic tissue, a
heightened erythropoiesis and myelopoiesis, and an
increased survival rate compared with x-irradiated
mice.  The increase in the number of endogenous
hematopoietic  colonies  in the spleens  of the x-
irradiated mice after  microwave exposure supports
Rotkovska and Vacek's earlier (1975) observation of
an  elevation  in the  number of stem  cells in  the
spleens of intact mice  after microwave  exposure
alone. These investigators suggested that  RF
radiation may influence the mechanisms  that
activate the pool of stem cells, either by improving the
repair of sublet ha I radiation damage or by increasing
the proliferative capacity of stem cells that survive x-
irradiation. The authors concluded that this accelera-
tion of the processes of repairing radiation damage in
hematopoietic cells after thermogenic doses of RF
radiation depended on the stage of intracellular repair
at the time of RF-radiation exposure. In earlier work,
Michaelson et al. (1963) reported that simultaneous
exposure to X rays and microwaves (2800 MHz, PW
modulated, 100 mW/cm2, SAR estimated at 4 W/kg)
caused an accelerated recovery of the hematopoietic
function in dogs. Thomson et al. (1965) reported that
pretreatment of mice with RF radiation (2800 MHz,
PW modulated, 100 mW/cm2, SAR estimated at 70
W/kg) reduced the mortality after x-irradiation (800
rads). The 30-day lethality  was  40 to  55 percent
among mice given single or multiple  RF treatment
prior to  x-irradiation, compared with 76 percent
lethality  in mice not pretreated with RF  radiation.
Exposure of Chinese hamsters to RF radiation (2450
MHz, CW, 60 mW/cm2, SAR estimated at 28 W/kg,
for 30 min) 5 min after X irradiation (725 to 950 rads)
                       B-16

-------
significantly increased the X-ray LD|8 dose compared
with exposure of animals to X rays only or with
exposure to RF radiation before irradiation (Lappen-
busch et at.  1973). Lappenbusch et al. reported that
the radio-protective effect of RF radiation appears to
be associated with a delayed decrease in the number
of circulating white blood cells, reduced period of
decreased cell density, and complete replenishment
of white blood cells within 30 days following the dual
treatment. Exposure to  RF radiation either alone or
combined with X-ray exposure increased the relative
number of neutrophils,  reduced the relative number
of lymphocytes, and slightly increased the number of
circulating  red blood  cells.  On the other hand,
animals exposed first to RF radiation and then to X
rays demonstrated a more severe leukocyte picture
than hamsters x-irradiated only;  in these animals,
leukocyte counts decreased faster, and the animals
developed leukopenia.

The effect of exposure to RF radiation on circulating
blood cells of developing rats has been studied by
Smialowicz et al. (1979a, 1982). Rats were exposed
pre- and postnatally to 425-MHz (CW) fields at 10
mW/cm2, 4 h  daily up to 41 days. Because  the
animals were growing, SARs ranged from  3 to 7
W/kg. No consistent changes in blood values were
observed in exposed compared to sham-irradiated
control rats (Smialowicz et al. 1982). Rats exposed
under the same regimen but to 2450-MHz (CW) fields
at 5 mW/cm2 (SAR estimated at 1 to 5 W/kg) also
showed no difference  in circulating erythrocyte
count, leukocyte and differential counts, or hematocrit
and hemoglobin concentration compared with sham-
irradiated  controls (Smialowicz et al. 1979a). Rats
exposed to  100-MHz fields  at 46 mW/cm2 (SAR
estimated at 2 to 3 W/kg) pre- and  post-natally as
above  showed no change  in  blood  parameters
compared with controls (Smialowicz et al. 1981 a).

Pazderova-Vejlupkova and Josifko (1979) reported
decreases in the hematocrits, number of leukocytes,
and absolute numbers of lymphocytes in young rats
exposed to  2736-MHz  (PW,  395 Hz, 2.6-fJs pulse
width) microwaves at 24.4 mW/cm2 for 7 weeks (5
days/week, 4 h/day). The means of body mass of rats
at the beginning and  at the end of  the  7-week
exposure period was 65  and 350 g, respectively (SAR
estimated  at  5  to 25 W/kg). These changes
disappeared within  10  weeks after termination of
exposure.  The activity  of alkaline phosphatase in
neutrophils increased  during the  first week of
irradiation  but decreased transiently after the
irradiation.  In a similar experiment by the same
authors (data not given), in which adult rats were
exposed for 14 weeks to 3000-MHz (PW, 300 Hz, 2.5-
fjs  pulse width)  microwaves  at  1  mW/cm2 (SAR
estimated at 0.2 W/kg), no difference was observed
in hematologic  parameters between exposed  and
Control rats.
Hamrick and McRee (1975) examined the effect of RF
radiation  on developing birds. Quail  eggs were
exposed for 24 h during the second day of incubation
to 2450-MHz (CW) fields at 30 mW/cm2 (SAR = 14
mW/g). At 24 to 36 h after hatching,  quail chicks
were examined for gross deformities, changes in organ
weight, and  hematologic changes. No significant
effects due to RF exposure were detected.
In another study, Smialowicz et al. (1979b) exposed
mice to 2450-MHz (CW) fields at 30 mW/cm2 (SAR =
22 W/kg) for 30 min on 22 consecutive days. These
mice showed no significant difference in circulating-
erythrocyte count, leukocyte and differential counts,
or hematocrit and hemoglobin concentration
compared with sham-irradiated controls.  In this
experiment,  mice  were maintained in  an environ-
mental chamber in which temperature, humidity, and
air flow  were continuously controlled. Under the
conditions of this study,  the RF radiation did not
significantly elevate rectal temperatures of exposed
mice.  In  contrast,  when  mice were  exposed to
thermogenic levels (2 to 4°C rise in rectal temperature)
of 26-MHz  (CW)  radiation,  8610  mW/cm2 (SAR
estimated at 13 W/kg), a decrease in the number of
circulating lymphocytes and an increase in circulating
neutrophils was observed immediately after exposure
(Liburdy 1977). These  mice were held in a chamber
that  lacked  a continuous turnover of  air.  Liburdy
(1977) reported that this shift reached  its peak 3 h
after exposure. The number of circulating lymphocytes
and neutrophils was reported to return to normal,
pre-exposure levels 55 to 96 h after exposure. On the
other  hand, mice exposed at high ambient tempera-
tures  (79°C)  in a vented, dry-air oven  showed an
increased number of  circulating lymphocytes and
neutrophils for a  12-h period after exposure. The
response of circulating leukocytes to  exogenous
thermal loading thus depends on the means by which
the body is heated. These results are similar to those
reported by Rotkovska and Vacek (1977) and indicate
that the heating properties of RF radiation differ from
those of conventional modes of tissue heating.
Recently  Huang and Mold (1980) reported that bone
marrow (cultured in vitro) from mice exposed to 2450-
MHz (CW) fields at 15 mW/cm2 (SAR = 10 W/kg) for
30 min on 9 consecutive days had significantly fewer
(p < 0.05) CFUs of both the erythroid and granulocyte-
macrophage series. No data were presented on the
peripheral blood counts of any of these blood  cells
that  would confirm  or expand the  information
gathered in the CFU assay.

5.2.1.2 In Vitro Studies
Lin et al. (1979b) reported a reduction in the number
of CFU's (granulocyte  and  macrophage precursor
cells) formed by bone marrow cells exposed in vitrolo
2450-MHz fields at 60 to 1000 mW/cm2 (SARs at
120  to 2000 W/kg). This reduction in  colony
                                                                     5-77

-------
formation was reported to be dose dependent and
occurred without a significant rise in the temperature
of the cell suspension. The authors indicate that their
results point to a direct effect of microwave radiation
on these  hematopoietic precursor cells. Although
these results are interesting, the in situ application
of fields as intense as those required to produce the
observed effects would certainly cause gross thermal
injury to the tissue.

In summary, levels of RF  radiation that cause an
increase  in body temperature elicit changes in the
hematopoietic system that can for the most part be
ascribed to a thermal stress response. Changes in the
blood of animals exposed to RF radiation at intensities
below those that cause an  increase in core
temperature suggest a  similar stress-response
mechanism. The failure to record an increase in core
temperature does not preclude the possibility that the
animal is compensating for the added thermal energy
by thermoregulatory mechanisms. Indeed, lack of a
temperature change indicates that thermoregulation
is operating. The response elicited by RF-radiation-
induced heating,  however, appears to differ from that
of conventional  heating because  of the differing
internal heating patterns of this form of radiation.
5.2.2 Immunology
The  immune system is comprised of myriad
mechanical, cellular, and humoral components that
act as the body's defense against various pathogenic
microorganisms, viruses, and neoplasias. The
immune system is divided into the humoral element,
i.e., antibodies  and  complement, and  the cellular
elements, which are composed of the lymphoid and
phagocytic cells. The cellular elements of the immune
system are also part of the hematologic system. The
phagocytic cells  responsible  for  engulfing and
digesting certain microorganisms are the neutrophils
or polymorphonuclear leukocytes (PMNs) and the
monocytes or macrophages. In the presence of
antibodies and complement, neutrophils are aided in
engulfing and digesting invading  organisms. The
monocyte is also a phagocytic cell. Monocytes move
into an area in which an infection has begun and then
differentiate into macrophages. Macrophages can be
"activated"  to  kill certain microorganisms (e.g.,
intracellular, facultative bacteria such asMycobacteria
tuberculosis   and Listeria monocytogenes, viruses,
and  fungi)  through  the  interaction of certain
subpopulations of  lymphocytes,  such  as  the T
lymphocytes.

The other cellular components of the immune system
are the lymphocytes. These cells are broadly divided
into  two groups, the  B  lymphocytes and the T
lymphocytes. Although  these cells  are similar
morphologically, they are different functionally; Band
T lymphocytes can be distinguished by the presence
of unique antigens or receptors on their membrane
surface. Both T and  B  lymphocytes are believed to
originate in the bone marrow and then to proceed
through various stages of development and differen-
tiation, maturing into functional cells of the immune
system.

The B lymphocyte, or bursa-equivalent  lymphocyte, is
responsible for humoral immune responses. The B
lymphocytes, after appropriate stimulation  by
antigens, proliferate and undergo morphological
changes and develop into plasma cells that actively
synthesize and secrete  antibodies.

The T  lymphocyte, or thymus  lymphocyte,  is
processed through the thymus after leaving the bone
marrow. Classically cell-mediated or  T-lymphocyte
responses include protection against viruses, fungi,
and several bacteria. T lymphocytes are also involved
in reactions such as delayed hypersensitivity  or
contact hypersensitivity and rejection of tumors and
foreign tissues such as  transplants (allografts). Cell-
mediated  reactions  are so named because these
reactions,  which operate by specifically sensitized T
lymphocytes, can be transferred by these cells  to
normal animals. B-lymphocyte-mediated humoral
responses, in contrast, are transferable by serum.


The recent availability of monoclonal antibodies has
made the  typing of lymphocyte subpopulations
possible on a routine basis.  Various functional
categories of  T cells can be recognized:  "helper"
(inducer/amplifier) T cells,  "suppressor" T cells,
"cytotoxic" T cells, and  NK (natural killer) cells.
Immature  thymocytes may be identified by surface
antigens distinct from  mature  peripheral T cells.
Cellular interactions between  the various T-cell
subtypes and other immunocompetent cells are vital
to the modulation of the immune response. Because
of their  pivotal role in enhancing antigen-mediated
immune responses, disorders involving T-cell
subsets  may result in immunodeficiency syndromes
involving either cell-mediated or humoral immunity.


Each element of the  immune system—the T and B
lymphocytes and macrophages—plays  a cooperative
role in  defending the  host  against  infection and
disease. A-delicate balance exists to prevent the
immune system from reacting to its own tissues  so
that autoimmune  reactions  are avoided.  The
alteration  or dysfunction of any of these  elements
may lessen the host's ability to combat infection or
may lead to autoimmune disease. However, because
of  adaptability and  redundancy  in  the  immune
system, the  host can generally survive subtle
perturbations. Consequently,  although  subtle effects
on the immune system may be generated by physical
                       5-18

-------
or chemical agents, all such effects may not lead to
clinically significant immune dysfunctions.

5.2.2.1 In Vivo Studies
A summary of in vivo studies concerning immunologic
effects of RF-radiation exposure  is  presented in
Table 5-6.

Effects on adult animals—One of the most consistently
found RF-radiation-induced changes in the hemato-
poietic system  is the  increase in  lymphocyte
                             formation and activity following exposure of animals
                             of  several  species  to RF  radiation  at  various
                             frequencies (Baranski 1971, 1972a,b; Czerski 1975).
                             There have been several studies of the effects of RF
                             radiation on  lymphocytes and the immune system. In
                             a  study reported  by  Czerski (1975),  mice were
                             exposed 2  h  daily  to  2950-MHz (PW  modulated)
                             microwaves at 0.5  mW/cm2 (SAR estimated at 0.5
                             W/kg) for 6 to 12 weeks. After 6 weeks, there was a
                             large increase in the relative number of lymphoblasts
                             in the lymph nodes of exposed mice. In another series
Table 5-6.    Summary of Studies Concerning Immunologic Effects (In Vivo) of RF-Radiation Exposure
                                                           Exposure Conditions
Effects-
         Frequency
Species   (MHz)
Intensity
(mW/cm2)
Duration
(days x min)
SAR
(W/kg)
                                                                                            References
Increase in lymphoblasts in
lymph nodes and increased
response to SRBC

Increase in "spontaneous"
lymphoblast transformation
of cultured lymphocytes

Increase in lymphoblasts in
spleen and lymphoid tissue

Increased transformation of
unstimulated cultured lymphocyte
and decreased mitosis in PHA-
stimulated lymphocyte cultures

Transient decrease and increased
response of cultured lymphocytes
to PHA.  Con A, and LPS

Increased mitosis of PHA-
stimulated lymphocytes

Increase in CR* Fc*. and Ig*
spleen cells. Increased response
to B-cell mitogens. Decrease
in primary  response to SRBC

Increase in CR* and Fc*, spleen
cells

Increase in CR" spleen cells.
strain specificity

Increase in CR* spleen cells

Increased lethality to endotoxin

Increase in response of
cultured lymphocytes to T-
and B-cell  mitogens
No change
No change

Increase in T and B
lymphocytes in spleen
Decrease in DTH

Reduction of lymphocyte
traffic from lung to spleen

Decrease in NK activity, increase
in macrophage phagocytosis

Decrease in NK activity

Increase in macrophage viricidal
capacity

Decreased  response to PWM

No change

Decrease in tumor development
                               Mouse
Rabbit
Mouse
Mouse
Mouse
Mouse
         2950 (PW)    0.5
         2950 (PW)
         3105 (PW)
                                                                42 x 120
                                                                              0.5t
                                 24-48x120     0.8t
Chinese   2450 (CW)    5.15,30
hamster               or 45
         2450 (CW)    5 or 15
Rhesus    10-27(PW)   1320
monkey

Mouse    2450 (CW)    —
         2450 (CW)    —
         2450 (CW)    —
            6-8700


            5x 15




            1-17x30
              2t


              2.3, 6.9,
              13.8 or
              20.7


              3.6 or 10
            1 x 30         0.4-2.0t


            1 or 3x30      14
                                 1 x 15
                                 1 x 30

                                 1 x 30
                          11.8
                          5

                          10-19
Mouse
Mouse
Rat
Mouse
Rat
Mouse
Mouse
Mouse
Hamster
Hamster
Rabbit
Quail
Mouse
2450 (CW)
2450 (CW)
2450 (CW)
425 (CW)
2450 (CW)
100(CW)
26 (CW)
2600 (CW)
2450 (CW)
2450 (CW)
2450 (CW)
2450 (CW)
2450 (CW)
2450 (CW)
40
20, 30
5
10
5-35
46
800
5 or 25
30
25
25
10
5
(near-field
application)
1 x30
1 x 1 20
57 x 240
47 x 240
(Perinatal exp)
1-22 x 15 or 30
57 x 240
1 x 15 or
10x 15
1 x 60
2 or 9 x 90
1 x 60
1 x60
180 x 1380
12 x 1440
11-14 day of
gestation or
11-14 & 19-45
x20
28
12, 18
1-5
3-7
4-25
2-3
5.6t
3.8 or 19
21
13
13
1.5
4.03
35
              Czerski (1975)



              Czerski (1975)



              Mnoetal. (1974)


              Huang eta/. (1977)




              Huang and Mold (1980)



              Prince et al. (1972)


              Wiktor-Jedrzejczak et al.
              (1977a, b, c)



              Sulekera/. (1980)


              Schlagel el al. (1980)


              Smialowicz et al. (1981C)

              Riddle et al. (1982)

              Smialowicz ef al. (1979a, 1982)


              Smialowicz et al. (1979b)
              Smialowicz et al. (1981b)

              Liburdy(1979)


              Liburdy(1980)

              Smialowicz et al. (1983)


              Yang et al. (1983)

              Rama Rao et al. (1983)


              McReeetal. (1980a)

              Hamrick et al. (1977)

              Preskorn et al. (1978)
                                                                                   5-19

-------
Table 5-6.   (Continued)
                                                 Exposure Conditions
Effects*
Decreased granulocytic response
Tumor regression and increase in
antitumor antibodies and anti-BSA
Tumor inhibition and immune
stimulation
Increased tumoricidal activity
in lymphocytes and macrophages
Tumor regression
Increase in lung cancer
colonies and inhibition of
contact sensitivity to
oxazolone
Decrease in response to BSA
Species
Rabbit
Rabbit
Rat
Mouse
Mouse
Mouse
Rabbit
Frequency
(MHz)
3000 (CW)
13.56
2450 (CW)
1356
3000 (CW)
2450 (CW)
1356
Intensity
(mW/cm2)
3
(near-field
application)
200W
600-900
40
50
(near-field
application
Duration
(days x min)
42 to 84 x 360
1 x 10-15
3 or 6 x 45
1 x 5
1-14x 120
4, 7, 10 or
14 x 120
3x60
SAR
(W/kg)
0.5
(local
hyperthermia)
(local
hyperthermia)
(local
hyperthermia)
28t
36t
(local
hyperthermia)
References
Szmigielski et al. (1975)
Shah and Dickson (1978b)
Szmigielski era/. (1978) '
Marmor et al. (1977)
Szmigielski et al. (1977)
Roszkowski et al. (1980)
Shah and Dickson (1978a)
*SRBC = sheep red blood cells, PHA = phytohemagglutinin. Con A
immunoglobulin positive, Fc* = Fc portion of immunoglobulin. DTH =
tSAR estimated.
= concanavalin A, LPS = lipopolysaccharide, CR* = complement-receptor positive, Ig* =
delayed-type hypersensitivity, PWM -pokeweed mitogen, BSA = bovine serum albumim.
of experiments (Czerski 1975), rabbits were exposed
2 h/day, 6 days/week for 6 months to 2950-MHz
(PW) microwaves at 5 mW/cm2 (SAR estimated at 0.8
W/kg). After culturing for 7 days in vitro, peripheral
blood lymphocytes from these animals were found to
undergo an increase in "spontaneous lymphoblastoid
transformation." Maximal increases occurred after 1
to 2 months of exposure; the transformation rate then
returned to base line and rose again 1 month after an
irradiation had  been terminated. Miro et al. (1974)
continuously exposed  mice  to  3105-MHz (PW)
microwaves over a 145-h period at an average power
density of 2 mW/cm2 {SAR estimated at 2 W/kg). No
description was given of how animals were fed  or
watered. An  increase in lymphoblastic cells in the
spleen and lymphoid  areas of exposed mice was
observed.  A somewhat similar  response was
observed (Huang et al. 1977) in lymphocytes cultured
from Chinese hamsters that were exposed to 2450-
MHz (CW) microwaves for 15 min on 5 consecutive
days at 5 mW/cm2 (SAR =  2.3  W/kg) without a
detectable rise of rectal temperature. Increased
transformation to lymphoblastoid forms in the
absence of mitogens was maximal  in cultures from
hamsters exposed at 30 mW/cm2 (SAR = 13.8 W/kg).
Irradiation at  this power density caused a 0.9°C rise
in rectal temperature of exposed hamsters. Mitosis of
lymphocytes cultured in the presence of the mitogen
phytohemagglutinin (PHA), however, was depressed
in cells obtained from hamsters exposed at 5,15, 30,
or 45 mW/cm2. These effects were reported to be
transient and reversible; control levels  were again
observed after 5 to 10 days (Huang et al. 1977).

More recently,  Huang and Mold (1980) reported an
oscillating  response of spleen  cells  to  PHA,
concanavalin A  (Con A), and lipopolysaccharide (LPS)
from mice  exposed  to 2450-MHz fields at 15
        mW/cm2 (SAR = 10 W/kg) for 30 min. Mitogen
        responsiveness decreased significantly after 2 days
        of exposure,  returned  to  normal after  4 days of
        exposure,  and was significantly increased for all
        mitogens after 9 days of irradiation. Responsiveness
        to mitogens tended to return to normal or to fall to
        subnormal levels after 17  days of exposure. In
        contrast, when mice were exposed at 5 mW/cm2 for
        30 min on 5 consecutive days, a significant increase
        in the response to LPS was observed, whereas no
        change  was  observed in LPS responsiveness to
        exposure at 15 mW/cm2 for 5 days. When macro-
        phages were removed from spleen-cell suspensions
        of mice irradiated at 15  mW/cm2 for 9 days, the
        responsiveness to LPS was greater than the already
        increased spleen-cell responsiveness without macro-
        phage removal. However, addition of macrophages
        from the RF-irradiated  mice  to spleen-cell cultures
        from  nonirradiated mice caused a significant
        decrease in responsiveness to LPS. The authors
        suggest that  macrophage activation (macrophages
        from irradiated mice displayed increased spreading
        and increased phagocytosis of latex particles in vitro)
        by microwaves may be responsible for inhibiting LPS
        responsiveness.

        Prince et al. (1972)  reported the opposite effect in
        rhesus  monkeys. These  investigators  found an
        enhanced  mitotic  response  of  peripheral  blood
        lymphocytes  stimulated  in  vitro with  PHA from
        monkeys 3 days following a 30-min exposure to 10.5-
        MHz (PW) radiation at 1320mW/cm2 (SAR estimated
        at 0.4 W/kg).  Enhancement of mitosis  of cultured
        lymphocytes from  monkeys similarly exposed to
        19.27- and 26.6-MHz were also reported.  Also
        reported were  increases in circulating lymphocytes
        that ranged from 4 to 47 percent above pre-exposure
        levels. At a frequency of 26.6 MHz (SAR estimated at
                       5-20

-------
2 W/kg), the rectal temperature of monkeys following
exposure was reported to increase by 2.5°C above pre-
exposure levels.

The particular susceptibilities of lymphocytes to  RF
radiation have led to an examination of the effects of
this form of radiation on the immune response. For
example, Czerski (1975) reported that mice exposed
for 6 weeks to 2950-MHz (PW) microwaves at 0.5
mW/cm2 (SAR estimated at 0.5 W/kg) had signifi-
cantly greater numbers of antibody-producing cells
and higher serum antibody tilers following  immuni-
zation with sheep red blood cells (SRBCs). Interestingly,
mice that were exposed for 12 weeks did not show
this increased responsiveness. This return to normal
responsiveness may have resulted  from the  animal's
ability to adapt to the RF exposure after 12 weeks.
More recently, Wiktor-Jedrzejczak  et al. (1977a,b,c)
exposed  mice in a rectangular waveguide to 2450-
MHz radiation for 30 min at an average dose rate near
14 W/kg. At 3, 6, 9, and  12 days after a single or
multiple  exposures, mice were tested for (1) the
relative frequency of T and B splenic lymphocytes, (2)
the functional capacity of spleen cells to respond toT-
and B-cell-specific mitogens, and (3) the ability to
respond to SRBCs  (a T-dependent antigen)  or
dinitrophenyl-lysine-Ficoll (DNP-lys-Ficoll, a T-
independent  antigen).  A single 30-min exposure
induced a significant increase in  the proportion of
complement-receptor-positive (CR*) lymphocytes in
mouse spleens that peaked 6 days after exposure.
This effect was further enhanced by repeated (three)
exposures, which also  produced a significant
increase in the proportion of immunoglobufin-positive
(Ig*) spleen cells  (Wiktor-Jedrzejczak 1977a). A
significant increase in the proportion of Fc-receptor-
positive (FcR*) cells in the spleens was also observed
7 days after a single 30-min exposure (SAR = 13.7
W/kg). However, no change in the number of lg+ cells
in spleens of these mice was observed  (Wiktor-
Jedrzejczak  1977c). The  type and combination  of
surface receptors (CR, Ig, Fc) expressed on splenic B
cells represent different maturational stages in  B-
cell development. Wiktor-Jedrzejczak era/. (1977a,b,c)
were unable to demonstrate any change in  the total
number of  theta-antigen-positive (0*) T cells in the
spleens of mice following a single or multiple exposure
to 2450-MHz microwaves.  No change was detected in
the in vitro spleen-cell response to stimulation by the
T-cell-specific  mitogens PHA  and  Con  A or  by
pokeweed mitogen (PWM), which stimulates both T
and B cells  (Wiktor-Jedrzejczak et al.  1977a).
However,  the response to the B-cell-specific
mitogens lipopolysaccharide (LPS), polyinosinic
polycytidylic acid (Poly I  C),  and purified protein
derivative (PPD)  of tuberculin was significantly
increased over controls following a single exposure.
These results clearly correlate  with  the observed
changes  in the proportion of cells  bearing different
surface markers. Wiktor-Jedrzejczak et al.  (1977a)
noted that RF irradiation did not stimulate lymphoid-
cell proliferation per se but rather appeared to act as a
polyclonal B-cell  activator, which  led to an early
maturation of noncommitted B cells. They also found a
significant decrease in the primary immune response
to SRBCs, a thymus-dependent antigen, in mice that
had been immunized just prior to the first exposure to
RF  radiation. They suggested that this  decreased
response may result from the nonspecific stimulation
of some cells by RF radiation to mature before they
are  activated by antigen (SRBCs), so  that  the
proportion of unresponsive cells was increased.

Recently, Sulek  et  al.  (1980) corroborated this
increase  in CR+ and Fc* spleen cells in mice 6 days
following  a  30-min  exposure to 2450-MHz fields
(average SAR = 12 W/kg). The kinetics for increased
frequency of CR* cells in the spleens of irradiated
mice showed  an initial increase 3  days following
exposure that  persisted for 5 to 6 days and then
returned to normal within 9 to 10 days. The authors
determined the threshold for this effect by varying the
time of exposure with a constant forward power (0.6
W), or by maintaining a constant time of exposure
while varying the forward power (0.1 to 0.78 W). It
was shown that a minimum of  a single 15-min
exposure (11.8 W/kg) or a single 30-min exposure (5
W/kg) caused significant increases in CR* cells 3 or 6
days  after exposure. The  effect  of  absorption of
multiple subthreshold quantities of microwaves was
found to be cumulative only if the exposures occurred
within 1 h of one another. The increase in CR+ cells
was found at dose rates ranging from 10 to 18 W/kg
for mice within a weight range of  18 to 25 g. The
rectal temperature of exposed mice was found to be
elevated  no more than 0.6°C above that of sham-
irradiated mice. No change in 0* (T lymphocytes) cells
was observed. In contrast, Huang  and Mold (1980)
reported  a significant increase in  6* cells but  no
change in B cells (Ig*) of mice exposed to 2450-MHz
fields at 15 mW/cm2 (SAR at 10 W/kg) for 30 min on
9 consecutive days.
Smialowicz et al. (1979b)  exposed BALB/c mice (H-
2d) to 2450-MHz (CW) fields under far-field conditions
for 15 or 30 min daily for periods to 22 consecutive
days at power densities from 5 to 35 mW/cm2 (SARs
at 4 to 25 W/kg). Splenic lymphocyte function was
assayed by the in vitro mitogen-stimulated response
as measured  by 3H-thymidine  incorporation after
culture in the presence of T (PHA, Con A, PWM) or B
(LPS, PWM,  PPD) mitogens. The proportions of T (0*)
and  B (CR*) splenic cells  and the  primary immune
response of mice to  SRBCs were also studied.  No
difference in the response to mitogens or to SRBCs or
in the frequency of  T or B cells in spleens was
observed  in RF-irradiated  compared with  sham-
irradiated mice.
Subsequent to  the  report  by Smialowicz et  al.
(1979b),   Schlagel et al.  (1980)  reported that
                                                                      5-21

-------
sensitivity to RF-induced  CFT cell increases was
under genetic control.  Mice that  were responsive*
were found to be of the H-2k haplotype (i.e., CBA/J),
while other mouse strains of the H-2", H-2b, H-2d(i.e.,
BALB/c) and H-1'5 haplotypes  did  not display
increased CR+ cells following RF exposure. The age of
thp mice was also found to be critical for expression
of increase  CR* cells following 2450-MHz (SAR = 14
W/krO  RF-radiation exposure. Mice less than 12
wt  -o uid did not respond.

In light of the findings by Schlagel et al. (1980) that
indicated that RF-induced increases in CR+cells were
under genetic control, Smialowicz et al. (1981c) re-
examined the effect of RF  radiation on CR* cells in
CBA/J mice (H-2k). Increases in the frequency of CR+
spleen cells of CBA/J mice 6 days following a single,
2450-MHz exposure (SAR =28 W/kg), under far-field
conditions in an anechoic chamber, were reported by
Smialowicz et al. (1981c). This increase in CR+ cells
was observed only in 16-week-old mice and not in
younger mice, a result similar to that reported by
Schlagel et al. (1980). However, unlike the results of
Schlage et  al. (1980), an increase in CR* cells was
accompanied by a significant decrease in the number
of nucleated cells in the spleens of these mice that
were obviously under thermal stress. No increase in
CFT cells was observed in mice irradiated at SAR
values of 10,14, or21 W/kg, SARs at which Schlagel
et al. (1980) observed  effects. Smialowicz et al.
(1981c>  concluded that the  age and  strain  of the
mouse, the RF-exposure characteristics (waveguide
vs. far field), and the environmental conditions are all
sources of variation that affect CR+ cell appearance.

Evidence that exposure of mice to far-field 2450-MHz
RF radiation at an SAR > 12 W/kg is thermogenic
was provided by Riddle et al. (1982). Mice were
injected with various doses of lipopolysaccharide
(LPS) and exposed to 2450-MHz CW RF radiation, and
the 50-percent  lethal  dose (LDSO)  of  LPS was
determined. A significant decrease in the LPS LD5o
dose was observed at SAR values of 12 and 18 W/kg.
High ambient temperature (37°C) also potentiated
the lethal  effect of endotoxin, indicating that RF
heating was responsible for the observed effect. The
SAR values used in this study were comparable to
those used in studies by Wiktor-Jedrzejczak et al.
(1977a,b,c), Schlagel et al. (1980), and Sulek et al.
(1980) in which  augmented  CR* spleen cells were
observed. This further supports the hypothesis that
increases in CR* spleen cells following RF exposure
may be due to an RF-induced thermal response.
Further  evidence  for  an  RF-induced thermal
mechanism in  spleen cell  changes comes from
Liburdy (1979), who reported that changes in splenic
lymphocyte populations similar to those observed by
Wiktor-Jedrzejczak et al. (1977a,b,c) can be produced
by exposure of mice to thermogenic levels of 26-MHz
radiation. When mice were exposed to 26 MHz at an
intensity (800 mW/cm2, SAR  at  5.6  W/kg) that
produced a  2 to 3°C rise in rectal temperature, a
relative increase in splenic T and B lymphocytes was
observed. Similar responses (i.e., increase in T and B
cells) were induced following administration of the
synthetic glucocorticoid methyl prednisolone sodium
succinate. These results indicate that these  RF-
radiation-induced changes  might  represent  some
form of stress. It is difficult to understand how a 2 to
3°C rise in rectal temperature would occur in mice
irradiated  for 15 min at an SAR  of 5.6 W/kg. A
possible explanation is  that  these  mice were
restrained in perforated acrylic cages and held in an
exposure chamber in which the air was not circulated
during this period of irradiation.
In examining further possible mechanism(s) for the
shift in lymphocyte populations in the spleens of mice
exposed to RF radiation, Liburdy  (1980) examined
the circulation of lymphocytes in microwave-exposed
mice. Mice  injected with 51Cr-labeled syngeneic
spleen cells  were exposed for 1 h at  either 5 or 25
mW/cm2 to 2600-MHz  fields (SAR at  3.8 and  19
W/kg,  respectively).  Controls included sham-
irradiated mice, mice held in a 63°C warm-air oven
for 1 h, and  mice injected with methyl prednisolone
sodium succinate. The distribution of injected cells
was determined for the lung, liver, spleen, and bone
marrow at 1, 6, and 24 h after exposure. Exposures at
25 mW/cm2 caused  a 2.0°C increase in  core
temperature. This  regimen  led to  a  37-percent
reduction in lymphocytes leaving the lung and migra-
ting to the spleen. Also, a threefold increase in spleen
lymphocytes entering the bone marrow occurred in
this group of mice. A similar pattern  of lymphocyte
circulation was observed in the steroid-treated group.
No change in lymphocyte traffic was observed in mice
of the 5-mW/cm2 or warm-air groups.  Liburdy (1980)
concluded that these results suggest  that steroid
release associated with thermal  stress and attempts
by the animal to thermoregulate during  exposure to
RF radiation are responsible  for effects on  the
immune system.

More  recently, further evidence for an  association
between RF-induced thermal stress and effects on
the immune system has been reported. Thermogenic
doses of RF  radiation were found  to suppress the
natural killer (NX) cell activity of mice (Smialowicz et
al. 1983). Exposure of CBA/J mice to 2450-MHz (CW)
RF radiation at an SAR of 21 W/kg for 90 min on 2 or 9
days caused a  significant reduction splenic  NK
activity as determined using in vitro or in vivo assays.
No effect on NK activity was observed at  SAR values
of 3.5 or 10.5 W/kg. NK activity returned to normal
levels within 24 h following the last exposure at 21
W/kg. Treatment of mice with hydrocortisone also
caused suppression  of NK cell activity measured in
vitro and  in vivo.  Paradoxically,  a  concomitant
increase in  macrophage  phagocytic ability was
                       5-22

-------
observed in mice exposed to RF radiation at an SAR of
21 W/kg. These  results  were  essentially  in
agreement with work reported by Yang et a/. (1983)
and Rama Rao et  al.  (1983). Yang  er  al. (1983)
reported  that a  single, 2450-MHz exposure  of
hamsters at an SAR of 13 W/kg caused a significant
suppression of splenic NK activity that returned to
normal levels by 8 h. This exposure also caused a 2.0
to 3.5°C increase in colonic temperatures, as well as
elevated serum glucocorticosteroid levels. Exposure
of mice at 8 W/kg caused no  demonstrable effect on
NK cell  activity. In  another study Rama  Rao  et al.
(1983) reported  that exposure of  hamsters to
thermogenic levels (13 W/kg)  of 2450-MHz RF
radiation resulted in activation of peritoneal
macrophages. Macrophages from irradiated hamsters
were found to be  significantly more viricidal to
vaccinia  virus as compared to that from sham-
irradiated hamsters. In all of these studies (Smialowicz
etal. 1983; Yang era/. 1983, Ra mo Rao era/. 1983) a
strong association exists between RF-radiation-
induced  immune system  effects and RF-induced
thermal stress.

Effects on young animals—RF-radiation effects on the
development  of the immune response have  been
studied. Smialowicz etal. (1979a) exposed  rats on day
6 of gestation through 41 days of age to 2450-MHz
(CW) fields at 5 mW/cm2 (SAR at 1 to 5 W/kg). The
young animals absorbed microwaves at a higher rate
(5 W/kg) than the adults (1 W/kg). In this study the
exposed rats  had  lymphocytes that  responded to a
significantly  greater extent than those from control
animals following  in vitro stimulation by T- or B-cell
mitogens. A similar increase in lymphocyte respon-
siveness  was seen  in  another study in which rats
were exposed pre- and postnatally to  425-MHz
radiation  (SAR at  3 to  7 W/kg,  with the neonates
absorbing at the latter  SAR)  for  periods to 41 days
post part urn  (Smialowicz er al.  1982).  However,
lymphocytes from rats  exposed perinatally to 100-
MHz fields at 46  mW/cm2 (SAR at 2 to 3 W/kg)
showed no change in mitogenic responsiveness
(Smialowicz  er al.  1981b). The  results of the two
former studies indicate that long-term  exposure of
developing (especially neonatal) rats to RF radiation
at absorption levels higher than those achieved in the
latter study may give rise to increased responsiveness
of cultured lymphocytes. These results are similar to
other reported changes in  mammalian lymphocyte
responsiveness following RF-radiation  exposure
(Czerski 1975; Prince er al. 1972; Wiktor-Jedrzejczak
er al. 1977a,b,c). Increases in lymphocyte activity can
also be  elicited by  conventional heating (Roberts
1979). Although  the benefits are not known
concerning this  increased responsiveness to
mitogens  by  lymphocytes from animals exposed
perinatally to RF radiation, a recent report by Preskorn
er al. (1978) indicates that irradiation at this time
during development may be beneficial for  increased
 immunosurveillance against tumors. These investi-
 gators exposed mice to 2450-MHz fields (SAR = 35
 W/kg) for 20 min either on days  11 through 14 of
 gestation, or on days 19 through 45 postpartum, or
 during both periods. On the 16th day postpartum, all
 mice were  implanted with a  lymphoreticular  cell
 sarcoma. Mice irradiated in  utero only (colonic
 temperature  increase of 2.2°C in dams) showed  a
 lower incidence of tumors (13 percent vs. 46 percent
 for sham-irradiated mice) 93 days  postpartum  In
 mice irradiated in  utero  and postnatally, tumors
 initially developed at  a  lower rate compared with
 controls; however, after 2.5 months, no difference
 was observed in tumor incidence between groups. At
 the end of 4 months, the tumor incidence in irradiated
 mice was slightly greater than controls (46 vs. 40
 percent, respectively). An interesting finding,
 however, was that both tumor-bearing and tumor-
 free animals that had been irradiated only in utero
 lived longer  on the average than their respective
 controls. This result is  somewhat similar to that
 reported by  Prausnitz  and  Susskind  (1962), who
 found that mice briefly irradiated hundreds of times
 by a highly thermogenic level of RF radiation survived
 longer than controls.

 McRee  er al. (1980a) reported that 30'days after
 termination of a 6-month 23-h daily irradiation to
 2450-MHz fields (SAR = 1.5 W/kg) spleen cells from
 rabbits showed a decreased responsiveness to PWM.
 Decreased responsiveness to PHA and Con A by these
 spleen cells was also reported, but responses to PHA
 and Con A were not statistically different from  those
 of controls. Although these results are interesting,
 they are not conclusive and are of questionable value,
 because only four exposed and four sham-irradiated
 rabbits  were employed.  Also,  both  irradiated and
 sham-irradiated rabbits were transported from one
 laboratory to another (University of Washington to the
 National Institute of Environmental Health Sciences
 in Research Triangle  Park, N.C.) between the
 termination of RF exposure and spleen-cell assay.

 Hamrick er al. (1977) examined the avian humoral-
 immune response in Japanese quail exposed  to RF
 radiation during embryogenesis. Fertile quail eggs
 were continuously exposed to  2450-MHz  (CW)
 radiation at 5 mW/cm2 (SAR = 4.03 W/kg)
 throughout the first 12 days of development. At 5 weeks
 of age, quail were immunized with SRBCs, and the
 levels of anti-SRBC antibodies were determined. No
 difference was observed in the antibody liters of
 exposed and sham-exposed quails. The masses of the
 bursa of Fabricius (site of B-cell production in  birds)
 and spleen were not altered significantly by exposure
 to RF radiation.
Effects on phagocytosis—RF-radiation-induced
effects on phagocytic  leukocytes of animals  have
been reported by Szmigielski er al. (1975). Rabbits
were exposed to 3000-MHz fields for 6 h daily for 6 to
                                                                     5-25

-------
12 weeks at 3 mW/cm2 (SAR estimated at 0.5 W/kg).
After the last exposure to RF radiation, rabbits were
infected with an intravenous injection of virulent
Staphylococcus aureus. At periods before and after
infection, functional tests of granulopoiesis were
performed. The investigators reported a decreased
production of mature granulocytes in infected, RF-
radiation-exposed rabbits, which was manifested as
a more serious illness in these animals.
Summary—Exposure of  laboratory  animals to  RF
radiation can  lead to changes in the functional
integrity of  lymphocytes. These cells play  an
important role in the immune-defense system of man
and animals. The significance of the changes caused
by RF radiation is difficult to interpret. Although some
studies  indicate that  RF radiation causes  an
increased responsiveness, of lymphocytes (Czerski
1975; Smialowicz et al.  1979a, 1982; Prince et al.
1972; Wiktor-Jedrzejczak  et al. 1977a,b,c) and a
potentiation  of the immune  response to antigen
(Czerski  1975),  others  indicate a depression in
responsiveness (Huang et al. 1977; Wiktor-Jedrzej-
czak et al. 1977a; Liburdy 1979; Szmigielski et al.
1975). In most cases  these  alterations  can  be
attributed to  a stress response, since qualitatively
similar but quantitatively more pronounced changes
are observed at obviously stressful thermogenic
levels of RF radiation (Huang ef al. 1977; Huang and
Mold 1980; Prince era/. 1972; Liburdy 1979; Riddle et
al. 1982; Smialowicz eta/. 1981 b) or after administra-
tion of glucocorticoids or detection of increased levels
of glucocorticoids (Liburdy 1979; Smialowicz et al.
1983; Rama Rao et al.  1983; Yang et al. 1983). It is
well known that stress alters physiological systems
that regulate immunological function. Both immuno-
suppression  and immunoehancement have been
observed to result from stress (Monjan 1981; Blecha
et al. 1982;  Bradley and  Michel! 1981; Palmbald
1981; Rogers and Matossian-Rogers 1982), which is
not inconsistent with the  reported effects of  RF
radiation on the immune system. However, stress-
induced modulation of the immune system is still
imperfectly understood. Recent data suggest that the
mechanisms  by  which stress affects the immune
system are more complex than previously recognized
and that, in addition to  adrenal-dependent pheno-
mena, adrenal-independent effects, which are as yet
poorly understood, may  be operative (Keller et al.
1983). The role of stress as a possible mediator of
observed RF effects on the immune system remains
speculative,  although  useful heuristically, at the
present time. Critical experiments (especially, with
careful  consideration  of other environmental
stressors and proper controls—i.e., adrenalectomized
animals) examining the relationship between stress
and RF-radiation effects  have yet to be done.

Effects caused by RF-induced hyperthermia—
Alterations in the immune system can be produced by
RF-radiation-induced hyperthermia. Whole-body
microwave-induced hyperthermia has been reported
to serve a therapeutic role either alone (LeVeen et al.
1976) or in combination  with ionizing radiation
(Nelson  and Holt 1978). In many cases, the  direct
destruction of malignant tissues by RF-radiation-
induced heating  is the ultimate goal. However, in
some cases, hyperthermia has led to changes  in the
immune response. For example. Shah and Dickson
(1978b)  reported that following local heating of VX2
(carcinoma) tumor-bearing rabbits by a  13.56-MHz
field, tumor regression and host cure were observed
in 70 percent of the rabbits. Intratumoral temperatures
of 47 to 50°C were achieved within 30  min. Along
with tumor regression, cell-mediated immunity—as
measured by skin reactivity  to tumor  extract and
dinitrochlorobenzene—markedly increased. A hund-
redfold  increase in serum levels of  antitumor
antibody and increased response  to the antigen
bovine serum albumin (BSA) were also observed. In
contrast, whole-body hyperthermia led to temporary
reduction of tumor growth, followed by a return to an
exponential  increase in tumor volume  and  rapid
death of the rabbit. This course of events following
whole-body hyperthermia was accompanied by
abrogation of the enhanced  cellular and humoral
immune  responsiveness, observed  following  local
RF-induced heating.
Szmigielski ef al. (1978) reported that local heating
(43°C) of the Guerin  epithelioma in Wistar rats by
2450-MHz (CW)  radiation inhibited tumors and
stimulated the immune reaction against the tumor.
Other immune reactions stimulated by this treatment
were the antibody response to BSA, high reactivity of
spleen  lymphocytes to the mitogen  PHA, and
increased serum lysozyme levels as a measure of
macrophage activity.  Tumor-specific  reactions
observed were increased cytotoxicity of spleen cells
and peritoneal macrophages to cultured tumor cells.
Similar results were reported by Marmor etal. (1977),
who exposed tumors in mice to focal   1356-MHz
radiation. The EMT-6 tumor was found  to be
extremely sensitive to RF heating. The cure rate was a
function of temperature and duration of exposure. A
5-min exposure at  44°C  reduced  the  tumors by
almost 50 percent. To determine the effectiveness of
RF-induced heating on tumor regression, tumor-cell
survival was studied  by the  treatment of EMT-6
tumors  in situ.  Cell  inactivation by RF-radiation-
induced heating was similar to that  for heating by a
hot water bath. The results indicated that direct cell
killing could not account for the observed cures, and
these investigators suggested that hyperthermia (RF
or convection-induced) may stimulate a tumor-
directed immune response.
Szmigielski  et al. (1977) exposed mice bearing
transplanted sarcoma-180 tumors to  3000-MHz
radiation, 2 h daily on the 1 st through 14th day after
transplantation, whole-body at 40  mW/cm2 (SAR
                       5-24

-------
estimated at 28 W/kg). This exposure led to a 3 to4°C
increase in  rectal  temperature  and resulted in a
reduction of tumor mass by ~40 percent, a reduction
enhanced when microwave hyperthermia was
combined with Colcemide, Streptolysin S, or both.
Colcemide enhances the inhibiting effect of hyper-
thermia on proliferation of cells//; v/fro(Szmigielski et
al.  1976), and  Streptolysin S is an  antineoplastic
agent. Szmigielski  et  al. (1978) suggested that
immunostimulation is  important in the complex
inhibition of tumor growth by increased temperature.

Although many investigators see local and systemic
hyperthermia as a possible cancer treatment, either
alone or in combination  with  drugs or ionizing
radiation, there is evidence that hyperthermia may
enhance the  dissemination of certain cancers and
abrogate the immune response. For example,
Roszkowski et al. (1980) reported that exposure of
mice to 2450-MHz radiation at 50 mW/cm2 for 4, 7,
10, or 14 days, 2 h daily (SAR estimated at 36 W/kg)
caused an increased number of lung-cancer colonies
and an inhibition of contact sensitivity to oxazolone (a
measure of  T-lymphocyte  activity) with  increased
duration of hyperthermia. Shah and Dickson(1978a)
exposed normal  rabbits either to  RF-radiation-
induced (13.56 MHz) or to watercuff-local hyperther-
mia of thigh muscles, which are maintained at 42°C
for  1 h on 3 consecutive days. No alteration in the
response to dinitrochlorobenzene challenge was
observed. However, the humoral immune response to
BSA was significantly depressed. This response was
independent of the method and  degree of heating.
The results indicate that B lymphocytes might  be
more susceptible to hyperthermic damage than are T
lymphocytes.

The above results indicate that if  the applied
microwaves are of sufficient intensity to cause heating
of tissue or  of the whole animal, changes in the
immune  system will follow.  With heating  to any
extent, the  hypothalamic-hypophyseal-adrenal axis
plays a major  role in the responses elicited. It is well
known that  endogenous  or exogenous adrenal
glucocorticoid hormones affect the immune response.
In addition, heat alone may affect immune function.
For example,  heat  has been shown to  affect the
response of mitogens in vitro (Ashman and Nahmias
1978; Roberts and Steigbigel 1977; Smith et al. 1978;
Gutman  and  Chang  1982),  which suggests that
elevation of temperature per se may mediate some
of the observed RF-radiation-induced changes. But
what of the  RF-induced responses reported in the
absence of measurable temperature increases or at
rates of energy absorption well  below that of the
resting or basal metabolic rate?  Observed changes
in the immune response under these conditions are
more difficult to explain on a thermal-stress basis,
primarily because of a lack of sensitive techniques to
detect subtle stress responses. However, no increase
 of rectal temperature after exposure to RF radiation
 does not mean that the animal might not compensate
 for  added thermal energy by thermoregulatory
 mechanisms.  RF radiation may also  cause focal
 heating (thermal "hot spots") in organs critical to the
 immune response.

 5.2.2.2 In Vitro Studies
 Among the studies in this area (Table 5-7), several
 have involved attempts  to determine if in vitro
 exposure of lymphocytes to RF radiation leads to
 "direct" changes in the metabolic or functional state
 of these cells. In an early study, Stodolnik-Baranski
 (1967) exposed  human lymphocytes in culture to
 3000-MHz (PW) microwaves at 7 or 14 mW/cm2.
 Some  lymphocytes  were  irradiated 4  h/day at 7
 mW/cm2 for 3 to 5 days, while those exposed at 14
 mW/cm2 were irradiated 15 min daily for 3 to 5 days.
 After 5 days in culture, the microwave-exposed cells
 were found to have undergone a fivefold increase in
 lymphoblastoid transformation  compared with
 controls. Czerski (1975) attempted without success to
 repeat this experiment. But, in a more recent study,
 Baranslki and Czerski (1976) reported that exposure of
 human lymphocytes to 10,000-MHz fields at power
 densities between 5 and 15 mW/cm2 could induce
 lymphoblastoid transformation  (SAR not given). At
 power densities below 5 mW/cm2, this effect was not
 observed, whereas at power levels above  20
 mW/cm2, cell viability decreased.  The induction of
 blastic transformation depended on termination of
 irradiation (5 to 15 mW/cm2) at the moment when the
 temperature of the medium reached 38°C. These
 results indicate that the microwave-induced blastic
 transformation might be caused by  a thermal
 mechanism.

 Similar increases in the lymphoproliferative response
 of cells exposed to temperatures > 37°C have been
 reported. As  mentioned.  Ashman and Nahmias
 (1978) reported that  human lymphocytes, when
 cultured  at 39°C with the mitogens PHA or Con A,
 showed an enhancement and  earlier onset of 3H-
 thymidine  incorporation  compared with  cultures
 incubated at 37°C. In  a similar study, Roberts and
 Steigbigel (1977) reported that the in vitro human
 lymphocyte response to  PHA  and the common
 antigen streptokinase-streptodornase was enhanced
 at 38.5°C  relative to 37°C. Smith et al. (1978)
 reported that the in  vitro response  of human
 lymphocytes to PHA, Con A, PWM, and allogeneic
 lymphocytes was markedly enhanced by culture at
40°C compared with 37°C. These studies demonstrate
the need to monitor and to control the temperature of
cultures exposed to RF radiation. Without adequate
temperature data, it is virtually impossible to accept in
 vitro effects as due to RF radiation  itself.

The proliferative response of lymphocytes exposed in
 vitro to RF radiation appears to be related to culture
                                                                    5-25

-------
Table 5-7.   Summary of Studies Concerning Immunologic Effects (In Vitro) of RF-Radiation Exposure*

                                               Exposure Conditions
Effects
Increased blastogenesis of
exposed lymphocytes in vitro
Increased blastogenesis
No change in mitogen
response to PHA, Con A
orLPS
No change in mitogen
response to PHA
No change in viability
or growth
Decreased macrophage
phagocytosis
Liberation of intracellular
hydrolytic enzymes and
increased death
Species
Human
lymphocytes
Human
lymphocytes
Mouse spleen
cells
Rat blood
lymphocytes
Human
lymphoblast
cell lines
(Daudi and
HSB:)
Mouse
macrophage
Rabbit
granulocytes
Frequency
(MHz)
3.000 (PW)
10.000
2,450 (CW)
2.450 (CW)
2,450 (CW)
2,450 (CW)
3,000 (CW)
Intensity
(mW/cm2)
7
14
5-15
10
5. 10 or 20
10-500
50
1 or 5
Duration
(days x min)
3-5 x 240
3-5x15
Observed effect
only when culture
temperature ap-
proached 38 °C
1 x60, 120, or
240
1 x240, 1440,
or 2640
1 x 15
1 x30
1x15. 30,
or 60
SAR
(W/kg)
t
t
19
0.7, 1.4
or 2.8
25-1200
15 J/min
t
References
Stodolnik-Baranska (1967)
Baranski and Czerski (1 976)
Smialowicz (1976)
Hamrick and Fox (1 977)
Lin and Peterson (1 977)
Mayers and Habeshaw (1 973)
Szmigielski (1975)
*PHA = phytohemagglutinin, Con A = concanavalin A, and LPS = lipopolysaccharide.
tUnable to calculate SAR.
temperature. Smialowicz (1976) exposed murine
splenic lymphocytes to 2450-MHz (CW) radiation for
1, 3,  or  4 h at  10 mW/cm2 (SAR  = 19 W/kg).
Immediately after irradiation the temperature of the
exposed cultures did not differ significantly from that
of controls, and cell viability was unchanged. These
cells were then cultured for 72 h in the presence of T-
or B-cell mitogens, and the proliferative response
was measured  by 3H-thymidine  incorporation.  No
difference was found in the blastogenic response of
microwave-exposed and sham-exposed spleen cells
to any of  the  mitogens employed.  In  a similar
experiment, Hamrick and Fox (1977)  exposed  rat
lymphocytes to 2450-MHz (CW) radiation for 4, 24, or
44 h at 5,10, or 20 mW/cm2 (SARs at 0.7,1.4, or 2.8
W/kg, respectively). The transformation of unstimu-
lated or PHA-stimulated lymphocytes was measured
with 3H-thymidine. No significant differences were
found in the proliferative capacity  of  lymphocytes
from exposed and control  cultures. The effects of RF
radiation on the  growth  and viability of cultured
human lymphoblasts was studied by Lin and Peterson
(1977). Human  lymphoblasts (cell lines Daudi and
HSB2) were exposed to 2450-MHz (CW)radiation in a
waveguide for 15 min at incident power densities of
10 to 500 mW/cm2; the corresponding SARs were 25
to 1200 W/kg.  No temperature increase was
observed, even  at the highest power density in the
capillary tube that held the cell suspension in the
waveguide. No change was observed in the viability
or growth  of microwave-exposed lymphoblasts
compared  with controls. These studies provide
further evidence that no change in lymphocyte
activity occurs following  RF-radiation exposure in
vitro when proper control of culture temperature is
achieved.
In vitro exposure of macrophages to 2450-MHz fields
has been reported by Mayers and Habeshaw (1973) to
depress phagocytosis. Monolayer cultures of mouse
peritoneal macrophages were perfused with suspen-
sions of human erythrocytes while being exposed to
2450-MHz radiation at 50  mW/cm2. The rate of
energy absorption in the sample was 15 J/min. The
phagocytic index of exposed cultures was significantly
lower than that of the control after  a 3*0-min
exposure. Macrophage phagocytic activity returned
to normal levels if RF irradiation was discontinued.
During irradiation, a 2.5°C temperature increase was
observed; however,  the  final temperature in the
culture vessel in any given experiment reportedly did
not exceed 36.2°C. The investigators concluded that
the observed  depression of phagocytosis in the
irradiated cultures was not thermally induced. The
2.5°C rise in temperature during  irradiation,
according to the authors, would have been expected to
enhance rather than depress phagocytosis, since it is
optimal at 38.5°C. The mechanism by  which the
observed effect occurs  is  not  known;  however,
heating effects are difficult to dismiss because of an
observed 2.5°C rise in culture temperature. Although
the temperature of the suspension medium did not
exceed  36.2°C, thermal gradients of much higher
temperature would be expected at the macrophage-
glass interface.
                       5-25

-------
An RF-induced effect on granulocyte integrity and
viability was reported by Szmigielski (1975). Rabbit
granulocytes were exposed  in vitro to 3000-MHz
(CW) radiation  at 1 or 5 mW/cm2 (SAR not given) for
15, 30, or 60 min. Cultures exposed at 5 mW/cm2 for
30 to 60 min showed increased numbers of dead cells
as demonstrated by an increase in nigrosine staining
and an enhanced liberation of lysosomal enzymes.
Exposure at 1 mW/cm2 did not cause increased cell
death but did lead to a partial liberation of hydrolase
enzymes. No change was observed in the temperature
of the irradiated cultures.  The  liberation of acid
phosphatase and lysozyme from  granulocytes was
observed in cell  suspensions  exposed at 1 or 5
mW/cm2; both suspensions  exhibited a time- and
dose-dependent response.
In summary, exposure of laboratory animals to RF
radiation may lead to changes  in the  functional
integrity of leukocytes, which play important roles in
the immune-defense system. The significance of the
changes caused by RF radiation is difficult to interpret,
since many observed effects are transient and
reversible.  Furthermore,  some studies indicate that
RF radiation causes  immunopotentiation, whereas
others indicate immunosuppression. In many cases
the observed alterations  in the immune system can
be attributed to thermal stress, because qualitatively
similar but quantitatively greater changes are
observed at obviously stressful (highly thermalizing)
levels of RF radiation or following the administration
of glucocorticoids. A  possible  explanation for the
immunomodulating effects of  RF radiation arises
from the timing of the  meausrements of immune
responsiveness after an animal is subjected to stress.
For example, corticoid-induced impairment  of
immune  responsiveness is commonly followed  by
homeostatic recovery, then  subsequent overcom-
pensation, which may be  associated with immunoen-
hancement.
As for reports in which measurable  elevations  of
temperature from RF-radiation-induced heating are
not detected, a possible role by RF-radiation-induced
thermogenesis cannot be dismissed. The failure to
detect a measurable increase in  tissue or core
temperature in RF-irradiated experimental animals
through the use of conventional techniques indicates
that the animal was able to compensate for the added
energy. The role that thermoregulating mechanisms
play in affecting the immune response needs further
study. There is at present no convincing evidence for
a direct effect of RF radiation on the immune system in
the absence of a thermal (heating) effect.

5.2.3. Unresolved Issues

Several issues relating to the effects of RF radiation
on the hematologic and immunologic systems remain
unresolved. Perhaps the  most perplexing question is
what to make  of the  many Soviet reports on these
systems. In most cases these reports lack sufficient
technical detail for adequate critical assessment of
reported results. Alterations in the hematopoietic and
immunologic systems have been reported in animals
exposed to RF radiation at and below 10 mW/cm2 over
periods of  weeks  to  months.  Nevertheless, no
convincing evidence has been presented to demon-
strate a direct effect of RF radiation in the absence of
thermal  involvement;  well-defined and planned,
chronic (months to years), low-level (< 1 mW/cm2)
studies have not been carried out to investigate Soviet
claims of possible immune alterations including
induction of autoimmune reactions following chronic
exposure. Investigations of the possible hematologic
•and  immunologic effects of PW vs. CW irradiation
have not been undertaken in response  to the claims
by Eastern European investigators that PW modulation
is more effective than  CW irradiation in  causing
alterations in these systems.

Investigations are lacking that would define  possible
synergistic effects of other agents or drugs with RF
radiation on the hematopoietic and/or immune
systems. RF radiation (at hyperthermic doses) has
been shown to provide a protective effect against
damage by  ionizing radiation  to the  hematologic
system. Although this  is a beneficial  effect of RF
radiation, it is not known whether the combination of
drugs or other physical agents with RF radiation is
detrimental.

Another issue is based on the  recent  work of
Szmigielski  et a/. (1980, 1982) who described the
increased incidence of cancer development in mice
exposed chronically to 2450-MHz (CW) fields. After
exposing the animals for several months at either 5 or
15 mW/cm2  (SAR  =  2  to  3 and 6  to 9 W/kg,
respectively),  these investigators  reported an
increased incidence of spontaneously arising mammary
tumors in C3H/HeA mice and increased skin tumors
in BALB/c mice whose skin was painted with 3,4-
benzopyrene.  These workers also reported that
chronic stress (i.e., overcrowding of mice) produced
an acceleration in tumor development comparable to
that found in mice irradiated at 5 mW/cm2, with still
greater acceleration  in tumor development in mice
exposed at 15 mW/cm2. These results suggest that
RF radiation at 15 mW/cm2 was more stressful (i.e.,
via thermal-induced stress) than  at  5  mW/cm2.
Szmigielski era/. (1982) suggest that the acceleration
in tumor  development in mice irradiated at 15
mW/cm2 may  be due to local thermal effects or RF-
induced "hot" spots.  Acceleration of  cancer
development  in  these mice  was suggested by
Szmigielski et a/, to be accompanied by  a lowering of
natural antineoplastic resistance, in particular effects
on  immunocompetent cells. These investigators
conclude that the effects on immunocompetent cells
may be due to a direct interaction of RF radiation with
the immune system or via a stress response. Based
                                                                     5-27

-------
on the results obtained from other laboratories (Rama
Rao et a/. 1983; Yang etal. 1983; Liburdy 1977,1979,
1980; Smialowicz et  al. 1981, 1983;  Smialowicz
1979; Riley 1981), as well as the results in the
Szmigielski et al. (1980, 1982) studies with  mice
stressed by overcrowding, it appears that the latter
hypothesis (i.e., RF-radiation stress-induced effects)
is the most plausible explanation. Nevertheless, this
effect warrants corroboration because of its potential
significance. Studies should be undertaken that
strive to  determine the  threshold for RF-radiation-
induced accelerated tumor growth.  Strict control of
both  the RF-exposure  parameters and ambient
environmental parameters is essential in any future
studies. Several  species should be employed to
determine whether this phenomenon occurs across
species.
                       5-28

-------
 5.3 Reproductive Effects
 Ezra Herman
 This  section  on the reproductive effects of  RF
 radiation is  organized  into three  categories:
 Teratology (5.3.1), where the treatment is adminis-
 tered to the pregnant dam and observations are then
 made  on  the  embryo, fetus,  neonate, or older
 offspring; Reproductive Efficiency (5.3.2), where the
 end point of the experiment occurs in the primary and
 secondary sex organs of the parent;  and Testes
 (5.3.3), where testicular morphology and function of
 testes are examined for alterations.

 RF radiation has been examined intensively for its
 potential reproductive effects for a variety of reasons:
 (1) many reproductive toxicologic tests can be carried
 out in a simple and inexpensive experimental design
 after  insult by RF radiation, (2) the  laboratory
 techniques used are  conventionally acceptable as
 toxicologic assays, and (3) detrimental reproductive
 changes carry not only an emotional impact but have
 potential for early and long-term consequences of a
 serious nature.

 5.3.7 Teratology
 The science of teratology and its underlying principles
 must be kept in mind if we are to make judgments on
 the potential of RF radiation for teratologic manifesta-
 tions. Wilson (1973) has developed general principles
 of teratogenesis. These principles are rephrased here
 to familiarize  the reader with the guidelines one
 should keep in mind when evaluating the available
 literature, so that their relevance  to  RF radiation
 teratologic investigations  may be more easily
 understood.

 (1)  Susceptibility to  RF radiation  teratogenesis
    depends  on  species,  strain, and stage  of
    development at the time of exposure.
 (2)  There are four indications of abnormal develop-
    ment: death, malformation, retarded growth, and
    deficient function.

 (3)  These indications increase in incidence and
    degree with increasing dosage.
Teratology was initially confined to the study of birth
malformations, monstrosities, and serious deviations
from normal. As teratology has matured into a branch
of toxicology, included  now in teratology  are toxic
manifestations in the fetus with lesser symptoms
than gross morphologic  changes. Such  fetotoxic
symptoms may be decreased fetal or birth weight, as
well as changes in function observed well after birth.
In this  discussion,  the terms "teratogenesis"  or
"terata"  refer to gross morphologic or monstrous
changes. Subteratogenic doses are those just below
teratogenic doses and  are not expected  to cause
 terata. Fetotoxic  symptoms include body-weight
 changes  without  terata. Functional changes are
 fetotoxic symptoms not readily apparent in the fetus,
 or even at birth, but are often seen later as postnatal
 maturation occurs. This discussion posits that the list
 of possible deviations caused in the conceptus is an
 order  of  decreasing  degree of severity (death  >
 malformation > growth  retardation >  functional
 changes). The significance of any laboratory animal
 or system as a model of human exposure may not be
 universally accepted among scientists.

 An attempt is made to derive from available data three
 aspects of teratologic toxicology of RF radiation: the
 presence of a teratologic effect, the generalization of
 that effect across species, and the dose-response
 character of that effect. From the evidence presented,
 we  believe that the  reader will agree with the
 following  statements:
 (1) RF radiation can cause teratogenesis in all the
    mammalian  species studied adequately so far if
    sufficiently high power  densities or  SARs are
    obtained.
 (2) Reduced fetal weight seems to occur consistently
    in  rodents exposed gestationally to teratogenic
    doses of RF radiation, or to  doses somewhat
    smaller  than those  which  cause  death or
    malformation.

 (3) There is evidence that gestational exposure toRF
    radiation  may cause functional changes later in
    life.
5.3.1.1 Nonmammalian Models

The potential for human teratology is usually sought
in mammalian  models. But other models that are
lower  on the phylogenetic scale or cell-culture
models are  also often informative. Workers at the
Bureau of Radiological Health of the Food and Drug
Administration have used an insect to demonstrate the
reproductive alterations due to RF radiation. Pay era/.
(1978) examined the egg production of female fruit
flies (Drosophila melanogaster) in response to RF
radiation. Using a 2450-MHz waveguide exposure
system housed in an environmentally  controlled
chamber (24°C, 50 percent relative humidity), these
workers determined the survival rate of D. melano-
gaster pupae at SARs ranging from 400 to 800W/kg.
Approximately 70 percent of the pupae did not survive
a  10-min  exposure when  the  rate of energy
absorption was 640 W/kg; the temperature of the
agar surface on which the pupae rested was45°C. If
pupae were incubated instead in a 44°C environment
without RF radiation, the death rate was less severe
(50 percent).

The potential for RF-radiation-induced teratology in.
birds has been examined by use of the fetal form of
                                                                     5-25

-------
birds, the egg, as a  model. This is a reasonably
popular model since  use of  bird  eggs is more
economical than use of mammals. Study of the bird
egg can provide insight into fetal effects independent
of maternal influence. The egg is an object that can be
placed  in  almost any desired position  by the
experimenter; it  will remain there, a distinct
advantage over free-ranging mammals. The egg is
symmetrical externally and so lends itself more easily
to estimates of local and total absorption of RF energy.

In theory,  mammalian models have no advantage
over avian  models in the  study of teratogenic
potential  of RF radiation because concepts  of
organogenesis apply equally to avian and mammalian
fetuses. It  would be  a  mistake to dismiss the
considerable research on the teratologic effects of RF
radiation in  birds'  eggs because the avian egg is
poikilothermic and the mammal's egg is homeother-
mic. The mammalian conceptus is also poikilothermic;
it  has little,  if any, control of its own temperature
since it is entirely surrounded by placenta! fluids. It
can dissipate thermal  energy only by radiating into
surrounding maternal tissues because its capacity to
rid itself of thermal energy depends entirely on the
gradient between itself and surrounding tissues. If
the temperature of the mammalian conceptus tissue
is increased because of RF-radiation absorption, it
also must  radiate that absorbed energy  into its
surroundings (dam),  just as  the  egg in  similar
exposure conditions must radiate energy into the air
of its incubator. If air or maternal temperatures are
detrimentally high, then the loss of thermal  energy
from the egg or mammalian fetus is affected similarly.

Carpenter et al. (1960a) have studied RF radiation as
an inducer of cataracts. These investigators have also
likened the  growth  in the lens to the process  of
embryonic development and growth, where prolifera-
tion and differentiation take place concurrently. To
further understanding of the RF-radiation-induced
effects that appear in the lens, Carpenter  et al.
performed experiments on the  egg of the domestic
chicken. They irradiated almost 500 chick embryos at
the 48-h stage and examined  the eggs 48  h  after
irradiation, scoring incidences of  survival and
structural  abnormalities. The irradiation  was
conducted in an anechoic chamber at 2450 MHz and
at power densities of 200, 280, and 400 mW/cm2.
Estimates of SARs for these power densities  are 70,
98, and 140 W/kg, respectively (Durney eta/. 1978).
Power densities were determined by calorimetry with
a  saline-filled egg. The exposure durations  ranged
from 1 to 15 min. The eggs were incubated at 39°C air
temperature  before, during,  and after irradiation.
However,  the temperatures  of the eggs are not
mentioned  in  the  report. At  the  96th hour  of
development, all eggs were opened, and the embryos
were removed, fixed, stained, and  examined as
whole-mounts.
Experimental results  relevant to this  section are
presented in Table 3 of the Carpenter et al.  1960a
publication.  We have used the data in that table to
calculate the relative summed incidences of dead and
abnormal 96-h  egg embryos for each experimental
group (power  density: time) and present the
incidences in Figure 5-2.

The teratologic end points observed by Carpenter et
al. (1960a) were death  and  abnormal morphology,
which can be viewed  as a continuum of the same
effect, interference with the normal processes of
development. When abnormal structure  is so severe
as to interfere greatly with the continuation of
development, the embryo cannot develop or maintain
sufficiently normal structural and  physiologic
systems. The relationships between these systems
collapse  and the embryo dies. As stated earlier,
indications  of  abnormal development  increase in
incidence and degree as the dosage increases. One
mechanism for the increase  in death is that  as the
dosage  increases,  more embryos  undergo   more
changes in  more systems, which  results in  more
deaths.

Figure 5-2 contains three curves, one for each group
of eggs exposed at dose rates of 140, 98, or 70 W/kg
for varying durations. The data are plotted against the
summed relative incidence of affected (dead or
morphologically abnormal) eggs. The slopes  of the
curves are steep,  so  that small increases in the
exposure duration cause disproportionate increases
of effect. For example, the incidence in the  group
exposed for 4 min at 140 W/kg was 9 percent. An
increase of  only 30 s in exposure  time caused  a
ninefold increase  (80 percent)  in  incidence. An
additional characteristic of  these slopes  is the
disproportionate spacing of the three curves along
the axis of exposure duration.
Figure 5-2.  Summed incidence of abnormal and nonviable
          chick embryo eggs (percent of total) exposed to
          2450-MHz radiation for varying durations at 70.
          98. or 140 W/kg (from data in Carpenter et •/.
          1960a).
        1  2 3 4 5 6 7 8 9 1011 121314
           Duration of Exposure, minutes
15
                       5-30

-------
Studies of RF radiation effects in the Japanese quail
have been conducted exclusively by workers at the
National Institute of Environmental Health Sciences.
In their first attempt to induce terata in the quail egg,
Hamrick and McRee (1975) exposed fertilized eggs in
an anechoic chamber to 2450-MHz  (CW) fields,
producing an SAR of 14 W/kg in the egg. The eggs
were exposed for 24 h  beginning 2 days after laying,
then were returned to their incubators until 24 to 36 h
after hatching. Several experimental replications
occurred; the  sample  numbers were  102  in
the control group and 110  in the RF-irradiated group.
After hatching, observations were made of cellular
elements of blood and of selected organ weights. No
significant differences between the two groups were
seen in hatchability, blood cell  parameters, body
weights, or organ weights.

The same authors also reported a similarly designed
experiment on  Japanese  quail eggs,  in which the
eggs were exposed during  the first  12  days  of
development  at SARs of 4.03 W/kg (McRee and
Hamrick 1977). As before, the eggs were allowed to
hatch  and were than examined for teratologic
alterations and  blood-cell parameters. When  the
experiment  was conducted  at an environmental
temperature of 35.5°C, there were  no significant
changes in body weights, organ weights, or blood-cell
parameters.

The authors also demonstrated the importance of
ambient temperature in experiments  in which RF
radiation is the agent under investigation. Absorption
of RF radiation energy produces increased tempera-
ture. In most studies of reproductive effects reviewed
in this document, fields  were at sufficiently high
power  densities to  cause  significant depositions of
energy. Temperatures  in the  animal subjects were
usually increased by 0.5°C or more. Therefore, the
ambient temperature in which the subject  is
maintained or treated can be a significant determinant
of a resulting trend.

Levels  of  ambient  temperature can be especially
relevant to  experimental subjects that have no
significant thermoregulatory capability. For example,
the report by McRee and Hamrick (1977) included the
results of an experiment  in which Japanese quail
eggs were in an environment at a temperature  only
slightly higher (+1.5°C)  than in the  experiment
discussed  above (35.5°C)  The authors began their
study at an ambient temperature of 37°C because
that is the temperature  at which quail  eggs are
conventionally incubated. But this ambient tempera-
ture, with the  addition of a 4  W/kg SAR, was
sufficient to  cause  a 93 percent death rate of the
fertilized eggs. The eggs that  were irradiated at the
same SAR, but at an ambient temperature lower by
1.5°C, sustained only a 42 percent death rate.  The
results of this seemingly slight (1.5°C) increase in
incubation temperature during RF irradiation shows
that for this poikilothermic model, the quail egg, the
temperature at which the egg is maintained during
experimentation is critical.
5.3.1.2 Mammalian Models

Morphologic teratology—The mouse  is commonly
used in the determination of the teratogenic potential
of toxic agents and also appears to be a useful model
for  determining the teratogenic  potential of RF
radiation. The mouse's size is conducive to this kind of
experimentation because the apparent resonance of
mice is 2450 MHz, a frequency at which irradiation
equipment is commonly and economically available.

Rugh  er  al.  (1974,  1975) conducted  several
experiments on  the mouse at  2450 MHz with a
waveguide exposure system to expose  individual
pregnant CF1 mice at air temperature of 25°C and 50
percent relative  humidity.  All  mice were exposed
once on the 8th day of pregnancy to CW fields at 2450
MHz for 2 to 5 min at a forward power that delivered a
measured average energy dose ranging from 2.45 to
8 cal/g;  there were no control or sham-irradiated
mice in this study.

These and a few other authors have reported dose
parameters in calories  or Joules per gram, which we
have converted to the  SAR unit (W/kg) used in this
document. Each of the texts of the two papers by Rugh
et al. (1974, 1975) refers the reader to the other for
more detailed  methodology, but neither adequately
describes exposure factors to convert accurately the
dose values to dose rate (SAR). Our best solution was
to convert to a range.  The range of doses given by
Rugh et al. is 10.3 to 33.5 J/g delivered over 2 to 5
min. The four  possible combinations (Table 5-8) of
two extremes of dose  (10.3 and 33.5 J/g) and two
extremes for exposure duration (2 and 5 min) result in
SARs ranging from 34 to 280 W/kg for this study. The
most probable  range of values is 85 W/kg (10.3 J/g
for 2 min) to 112 W/kg (33.5 J/g for 5 min).

The authors clearly demonstrated the teratogenicity
of RF radiation in mice, especially the potential for
death and generation of anomalies. By determining
the  energy  dose in each  pregnant mouse,  and
Table 6-8.
Average Dose
(J/g)
103
10.3
33.5
33.5
Conversion of J/g to W/kg*
Exposure Duration
(min)
5
2
5
2

SAR
(W/kg)
34
85
112
280
*W/kg = J/g x 1000 -r seconds (data from Rugh et al. 1974.1975).

                    5-31

-------
 because varying doses were absorbed during the 2 to
 5 min of exposure, the effect observed in each litter
 could be assigned to a definite whole-body average
 dose. These are the strong points in the Rugh et al.
 studies.

Rugh et al. (1974, 1975) showed that resorption of
fetuses is a characteristic  response to RF-radiation
exposure.  Resorptions  appeared  at a dose  rate of
approximately  85 W/kg. Both  the  incidence  of
resorptions within litters and the incidence of litters
with resorptions  increased  as  the average dose
increased. Near  the  maximum dose  (33.5 J/g,
approximately 110 W/kg) almost all the litters were
affected. Even at 25 J/g, there were litters with 100
percent resorption rates.

Anomalies "...were various, but the one used for this
study, and which occurred  most frequently, was
exencephaly (brain hernia)..."  The report also
included  data on resorptions, dead, and stunted
fetuses. In lieu of numerical data, the incidences of
brain hernia (also termed encephalocele, exencephaly,
or cranioschisis)  are given by the authors  in one
figure. Data, as percentage of affected liners, were
grouped into cells 8.4 J/g in width that ranged from
10.9 to 35.9  J/g. Approximately 30 percent of the
litters given an average dose of 15 J/g (approximately
90 W/kg) had fetuses with brain hernias. Of those
litters given an average dose of 31.8 J/g (estimated
as 109 W/kg) almost all contained fetuses with some
form of brain  herniation.

Though sham,  cage-control, or  historical control
value for the incidence of brain hernia in this mouse
strain (CF1) is not given, the authors state "...since
this aberration rarely if ever occurs without trauma, it
is significant when even a single exencephaly
occurs."  However, one report  of the  incidence of
spontaneous  brain hernia in CF1  mice showed an
occurrence of 11 exencephalic litters in a total of 90,
approximately 12 percent of litters (Flynn  1968).
Apparently, spontaneous  incidences  of anomalies
are quite variable.

A report by Rugh and McManaway (1977) included
data on mice exposed once for 4 min to 2450-MHz RF
radiation (SAR = 100 to 114 W/kg) on days 0 to 11 (10
mice per  day).  In this group of mice  (we  cannot
differentiate these data from those reported in Rugh
et al. 1975 and Rugh 1976a,b) are uneven incidences
of stunted fetuses (irradiated on day 4 of gestation)
and death (irradiated on day 3, 8, or 10 of gestation).
Rugh considers  small fetuses as stunted and believes
that this is a teratologic symptom.

 Rugh's experiments included using a single, but high
field intensity for a very short time (i.e., 4 min on the
8th day of gestation). A longer period (100 min daily)
was used by Berman et al. (1978) in the exposure of
pregnant CD1 mice to 2450-MHz CW fields at power
densities of 0 to 28 mW/cm2 in an anechoic chamber.
Between 15 and 25 mice were irradiated simultane-
ously, each  in a small ventilated container. Mean
rectal temperature was 38.2°C at the end of exposure
to the  highest power density (28 mW/cm2). SARs
were determined by twin-well  calorimetry  and
averaged 2.0 W/kg at 3.4 mW/cm2,  8.1  W/kg  at
approximately  14 mW/cm2, and  22  W/kg at 28
mW/cm2.

This study incorporated  the examination of several
variables of fetal  response to toxic agents. Values
were tabulated by power density for pregnancy rates,
incidences of live and dead fetuses, live fetal weights,
and incidences of anomalies. Values were expressed
that used the litter as the experimental unit. This is a
conservative approach in experiments in which the
potential fetal toxicity of  an agent is determined
(Atkinson 1975). Berman eta/. (1978)also "corrected"
the mean fetal weight  of each litter by a factor
representing the potential influence that litter size
had upon fetal weight. The technique is also endorsed
by Atkinson, because it allows greater confidence in
the interpretation of data.
The study by Berman et al. (1978) demonstrated that
daily exposure  to 2450-MHz  fields at intensities
producing an SAR of 22  W/kg during the entire
period  of organogenesis caused a  10-percent
decrease in body weight of mouse  fetuses.  The
practical significance of this result is  difficult  to
determine without experimental confirmation of the
permanency or lack of reversibility of the smaller fetal
weight. Until it can be shown that  the  effect  is
permanent (stunting) or temporary (delayed growth),
the effect remains hygienically unresolved. The effect
will  have more practical significance if it can be
shown that stunting is the result.
Berman et al. (1978) also reported on the incidences
of anomalies. There did not appear to be a significant
increase of individual or total anomalies at any SAR.
The C3H-HeJ strain  of mouse  was used by
Chernovetz et al. (1975) in an attempt to determine
the effects of fetal irradiation  with 2450-MHz RF
radiation.  Pregnant mice were  irradiated  in a
multimodal cavity for 10 min on the 11th, 12th, 13th,
or 14th day of gestation at an SAR (determined by
calorimetry) of 38 W/kg with or without the
administration of cortisone, or given cortisone without
irradiation.  The fetuses  were examined near term
(near the end of the gestation  period) without any
difference seen between the RF-irradiated and the
sham-irradiated groups of fetuses. The variables
observed in these fetuses related to the morphologic
and the lethal aspects of the RF irradiation. Fetal body
weights were not observed in these animals.

The scientific reports on the teratologic potential of
RF radiation in the rat represent the major fraction of
                        5-32

-------
rodent work. Perhaps the main reason is that many of
the studies that have looked for morphologic changes
have also included functional teratology as one of the
end points.  For that kind of study, especially
behavioral studies, the rat appears to be preferred
over the mouse.
Chernovetz  has  probably published most on the
teratologic potential of microwaves  in the rat. She
reported  (Chernovetz et  al.  1977)  the results of
exposures of pregnant rats to 2450-MHz fields in a
multimodal cavity (most likely the same device used
to irradiate mice in experiments discussed  above in
Chernovetz et al. 1975)  at  an SAR calculated by
calorimetry of approximately 31 W/kg. Exposures
were conducted  during 1 of 7 days, from the 10th
through 16th day of gestation, and the fetuses were
examined on the 19th day of gestation. Each single
exposure lasted  20 min. The experiment included a
total of 74 time-bred female rats, each assigned to 1
of 7 days of gestation and to 1 of 3 treatment groups
(RF-irradiated, infrared-irradiated, or sham-irradiated).
The result is an experimental design with 21  different
cells into which 74 rats were distributed.
By the end of the 20-min exposure period, the mean
rectal temperature in the surviving pregnant rats was
42°C. Of the original group of 30 bred females that
were used in the  RF-irradiation  group,  7 (or 23
percent) died; none died in the sham-irradiated group.
Therefore, the application of 2450-MHz RF radiation
to bred rats for this duration at this SAR was lethal for
a significant portion of the animals involved.
No numerical data on the gross teratologic conse-
quences of this RF irradiation were given in the paper.
However, comments were made in the text regarding
the lack of any structural abnormalities. The number
of resorptions was significantly higher (approximately
6  times) in  the  RF-irradiated than in the sham-
irradiated fetuses, especially in the  dams that had
been  irradiated on the 11th  day of gestation.  Fetal
weights were also altered by the irradiation regimen;
there was a small but statistically  significant
decrease in fetal  weight. The alterations seen in the
fetuses, in  this  case  decreased fetal  weights and
death (in  the form  of resorptions),  reflect the
teratogenic potential of RF radiation'exposure in the
rat. These fetal  alterations  were found from RF-
radiation exposure regimens where rectal temperatures
in the dam rose above 40°C.
Chernovetz et al. (1979) reported on the effect of
2450-MHz RF radiation in pregnant rats, this time at
SARs of 0,  14, or 28 W/kg. The same multimodal
cavity was used, through which air  at 22  ± 1.5°C
flowed at 0.75 m/s. Pairs of bred female rats were
exposed for a 20-min period on the 8th, 10th, 12th, or
14th day of gestation and then were examined on the
18th day of gestation. With cage controls added to
each of these cells, the complete experimental design
used 4 different days of treatment and 4 treatments
(cage control, 0 W/kg [shams],  14 W/kg,  and 28
W/kg), and 6 bred females per cell, for a total of 72
bred  females.  Those rats irradiated  at  28 W/kg
developed rectal  temperatures of approximately
42°C, temperatures that were similar to the authors'
previous experiments in which bred female rats were
dosed at 31 W/kg. The rats irradiated at 14 W/kg had
a mean peak rectal temperature near 40°C.
The values used by the authors in the  statistical
analyses were individual fetal values, not litter mean
values. At the higher SAR level (28 W/kg), only ~ 10
percent  below the sublethal level used in their
previous report, the authors did not  demonstrate
lethality  in the dams due to RF-radiation  exposure.
The 20-min exposure period at other levels of  SAR
produced no gross morphologic alterations in the
fetuses or any severely edematous fetuses; there
were also no statistically significant changes  in
resorption rates.
The results of fetal weights are most interesting.  In
the previous study (Chernovetz  et  al. 1977), the
authors  reported  that  exposure for 20  min  at
approximately 31 W/kg on odd-numbered  days  of
gestation produced a significant  decrease  in fetal
weight.  In the 1979 study, a similar result was also
obtained from exposure  at 14  W/kg  on even-
numbered days (12 and 14), or 28 W/kg on  the 8th
day of gestation. Unexpectedly, exposure at 28 W/kg
on the 12th or 14th day produced an increased body
weight. The unexpected increased body weight is an
interesting observation, especially when it is viewed
against the decreased fetal body weight documented
in the same authors' previous paper and other reports
of RF-radiation-induced fetal alterations in rats and
mice.
Berman  et al. (1981) were not  able  to  elicit  any
fetotoxic or teratologic responses  (body weight;
numbers alive and dead; external, visceral, or skeletal
morphology) in fetal rats  irradiated daily (100
min/day) during gestation, even with a large number
of litters. The dams were exposed to 2450-MHz
radiation at power densitites of 0 or 28 mW/cm2 (4.2
W/kg).  These  conditions caused a  mean rectal
temperature of 40.3°C by the end of  the 100-min
exposure.
Kaplan (1981) did not find differences in the survival
of squirrel monkeys up to 10 months of age after daily
in utero  and postnatal  exposures to 2450-MHz  in
multimodal cavities. Pregnant squirrel monkeys were
irradiated  for 3  h daily,  beginning  in  the first
trimester; after birth, the daily irradiation of the dams
and their young were continued, and  from 6 to 10
months of age only the young monkey was irradiated.
The SAR was determined  by calorimetry  to be 3.4
W/kg. In 21 irradiated and 22  sham  live-born,
differences were not seen  in sex ratio, the number
dead by  10 months of age, or in  the  mean age at
death. Postmortem examinations  did not reveal any
                                                                      5-33

-------
tendency to specific causes for death. The results of
this study did not confirm the results of suspected
increased lethality in squirrel  monkeys from a
previous study (Kaplan el al.  1982).
Functional teratology—Functional teratology is a
branch of reproductive toxicology that has allowed
increasingly greater insight  into the toxicology of
environmental agents. The  field contains a wide
variety of specialists who examine the functional
capacity of neonatal or growing animals after in utero
exposure.  Except for  the following  discussion of
viability and  of body weight, assays of functional
teratogenesis after in utero exposure to RF radiation
are discussed in  the  section dealing with each
relevant discipline.
The  survival  rate  of fetuses can  be used as  an
indicator of immediate effects on the fetus. The rate of
survival of neonatal, infant, or older off-spring can be
also used as an indicator of delayed alterations in
functional capacity long after insult during the fetal
stage. For example, Kaplan et al. (1982) reported a
study of the postnatal effects  of in utero exposure (at
dose rates to 3.4 W/kg) to 2.45-GHz RF radiation in
squirrel monkeys. One effect the authors observed
was an increase in neonatal deaths. However, after a
more intensive examination, Kaplan (1981) did not
demonstrate any change in death rate and could not
confirm the results of the study done earlier (but
reported later).
A simple test to determine whether RF irradiation in
utero would alter the survival of mice after birth was
conducted by Rugh (1976a). In this experiment,
pregnant CF1 mice were exposed on the 9th, 12th, or
16th day of gestation to a regimen of approximately 4
min of irradiation at 2450 MHz, which resulted in a
mean dose of ~ 25 J/g (104 W/kg). The survivors of
this sublethal exposure were allowed to give birth and
their young to mature to 2 months of age. At that time,
the offspring were again irradiated in the same device
until each one was dead. The time-to-kill and the
mean dose-to-kill was determined as a measure of
the radiosensitivity  of  the offspring  irradiated
originally as fetuses.  In this waveguide system,
measurement of absorbed dose was monitored by
power meters and calculated  by the difference of for-
ward power, transmitted power, and reflected power.
While the subject  in the  waveguide was alive, the
values of transmitted  and reflected powers were
constantly changing. When the subject died, these
values became constant.  At  the time of death, the
rectal temperatures of these mice, measured just
after the waveguide was opened, ranged from 40 to
51°C.
Because of body weight differences in the sexes at 2
months of age, the analyses were carried out by sex.
Time-to-kill was  shortened in  males irradiated
originally during  the 12th or 16th day of gestation.
The males irradiated originally on the 12th day of
gestation had a lower mean dose-to-kill. Therefore,
Rugh (1976a) has shown that RF irradiation on select
days in utero can alter later (sometimes much later)
effects of re-irradiation with microwaves.

As noted, one effect observed commonly in fetuses
exposed to RF radiation is decreased body weight.
There  are  difficulties associated  with interpreting
altered fetal weight as an  indicator of toxicity,
especially when the permanency of decreased fetal
weight (delayed growth  vs. stunting) is  not yet
determined. But Rugh (1976a) gives evidence that
mice irradiated as fetuses weigh less at 2 months of
age. In that study, the mean weights of 2-month-old
male offspring of dams irradiated during the 9th,
12th, or 16th day of gestation were all lower than
concurrent sham-irradiated males (p < 0.05).
Females irradiated during  the 16th day of gestation
were also smaller than their controls (p <0.05), but
not females from the 8th or 12th day groups. These
results are  evidence  of. a permanent change
(stunting) in mice caused by RF irradiation.

Chernovetz et al. (1975) carried out an experiment to
observe alterations in functional capabilities after in
utero irradiation. Pregnant mice were exposed once
during  the  14th  day of  gestation at an SAR  of 38
W/kg, and the offspring were examined until
weaning for survival. The  RF-irradiated and sham-
irradiated groups  each contained  15 dams. The
statistical analyses,  done by litters, show that there
was a slight (approximately 12 percent) increase in
the survival  rate  of the  RF-irradiated litters as
compared to the sham-irradiated controls.

Some experimental protocols continue irradiation past
birth. In one example of chronic adminsitration at 2450
MHz, reported by Smialowicz et al. (1979a), pregnant
rats were irradiated daily  for 4 h/day at a power
density of 4 mW/cm2 from the 6th day of gestation to
term; irradiation of the offspring continued fftrough
40 days of age. The SARs were determined by twin-
well  calorimetry for several  ages of the  animals.
Pregnant rats weighing 300 to 350g had a mean SAR
of 0.7 W/kg; offspring 1 to 5 days of age and 6 to 10 g
in weight absorbed  approximately 4.7 W/kg. There
was no significant difference between the mean body
weights of the males (female offspring were not used
in this  experiment) in the 12 sham-irradiated litters
when compared  to  the  mean body weights of the
males in the 12 RF-irradiated  litters.
Shore et al. (1977) reported decreased body weight and
decreased brain weight in  postnatal rats exposed in
utero to 2450-MHz RF radiation. In this experiment,
pregnant rats were exposed  at an  average power
density of 10 mW/cm2 (SAR estimated, from Durney
et al. 1978, to be 2.2 W/kg) for 5 h per day repeatedly
on days 3 through 19 of gestation. Rats were allowed
to deliver  naturally and were observed frequently
thereafter. Rats were  grouped  by E- or H-field
orientation during  exposures; only the 3-day-old
                       5-34

-------
offspring that had been exposed parallel to the E-field
were significantly different, having lowered body and
brain weights. We cannot easily explain this
preferential decrement on the basis of either age or
orientation during exposure except that the data were
repeatedly analyzed using unadjusted r-tests.  Our
expectation is that there is no significant difference
due to orientation in wholebody SAR in rats at this
frequency (Gage et al.  1979).
In experiments conducted  by Michaelson et al.
(1978), some apsects of functional teratology of RF
radiation were explored in Long-Evans rats. In these
experiments,  rats were exposed to 2450-MHz RF
radiation at power densities of 10 mW/cm2 for 1  h
on the 9th and 16th day of gestation, or 40 mW/cm2
for 2 h on the 9th, 13th, 16th, or 20th day of gestation.
According to data based on water-calorimetry, an
exposure  at 40 mW/cm2 produced an  SAR of
approximately 10 W/kg and approximately 2.5 W/kg
at 10 mW/cm2. Exposure at 40 mW/cm2 caused an
increase  in the mean of  rectal temperature of
approximately 1 to2°C over that of the sham-exposed
animals. The exposure at 10 mW/cm2 (2.5 W/kg) on
the 9th or 16th day of gestation caused a significant
increase (approximately 0.5 to 1 °C) in temperature of
dams at the 16th day of gestation. There were no
statistical  differences  in the sizes  of litters  as  a
consequence of exposure  during gestation. There
also appeared to be no difference in the growth rates
up to 21 days of age of the rat pups that were exposed
to RF radiation as compared with shams, nor in their
relative brain weights.
At the symposium where Michaelson described his
work, Johnson et al. (1978) reported on a study, then
in progress, on the functional teratologic effects in
rats exposed to 918 MHz for 20 h per day for 1 9 days
of gestation at  a power density of 5 mW/cm2.  The
experiment is especially interesting because it was
conducted at a  frequency close to the resonance of
the experimental subject (rat) and because it is the
only reported study conducted with almost continuous
exposure.  The SAR in this study is approximately 2.5
W/kg, determined by calculations from thermographic
scans. The  eight RF-irradiated  and eight sham-
irradiated bred dams  were  maintained  in  the
waveguide exposure apparatus under environmental
conditions of 22°C, 45 percent relative humidity, and
ad libitum access  to  rat chow and water.  The
exposure was begun on the first day of pregnancy, the
day on which the copulatory plug was first seen. In
this  experiment, the dams remained in individual
waveguides or  sham  condition until day 20 of
gestation,  at which time they were removed  and
placed in delivery cages. At 4 days of age, the litters
were culled to four males and four females per litter,
which were then observed through 91 days of age.

There appeared to be  no  differences in the  litter
means for the number of pups born or for number
 dead during the first day after birth. There were no
 pups in any of these litters (including RF-irradiated,
 sham-irradiated, and cage-controls) with any visible
 physical defects. Body weights at birth, at 28 days of
 age,  and  at 91  days  were analyzed for significant
 differences  on an  individual pup basis. The  only
 statistically  significant differences in body weights
 were found  between the RF-irradiated and cage-
 control females at 28 days of age. A difference also
 existed in males at 91 days  of age, when the RF-
 irradiated and the  sham-irradiated males weighed
 less than  the cage-control males.
 Another parameter of development used as an assay
 of functional teratology in this study was the age at
•eye-opening. The phenomenon of  eye-opening is a
 little-understood  but  frequently used  indicator  of
 developmental maturation. There appeared to  be a
 significant shift to  earlier eye-opening in animals
 exposed to  RF radiation  during gestation, and this
 maturation was earlier by approximately 1 day.

 Jensh reported a series of experiments designed to
 examine pre- and postnatally the effects of 915-MHz,
 2450-MHz, and  6-GHz  irradiation of rat fetuses
 (Jensh et al. 1979; Jensh  1979,1980). These studies
 included irradiation up to 8 h daily during most of
 gestation with power densities not sufficient to cause
 increased core temperatures (915 MHz, 10mW/cm2,
 2 W/kg; 2450 MHz, 20 mW/cm2,3 W/kg; 6 GHz, 35
 mW/cm2, 3.5 W/kg; SAR estimates from Durney et
 al.  1978,  p.  96). Offspring were examined pre- and
 post part urn in a complex series  of observations
 meant  to locate  behavioral  or morphological
 alterations. None was noted.

 In lieu of any other workable hypothesis supportable
 by  experimental data, the most plausible available
 theory that  can  account  for RF radiation effects is
 that,  upon absorption, microwaves deposit energy
 that is  converted to thermal  energy. There remain
 differences among theories that attribute individual
 effects to a general input of energy, or the heating of
 local (hot-spot) areas, or some more subtle contribu-
 tion of microwaves not characteristic of conventional
 methods of heating animals with infrared radiation,
 convective heating, or immersion in  water or oil
 baths.
 Comparisons between RF heating and infrared or
 convective heating have been  conducted in the study
 of hyperthermia as a teratogenic agent. Edwards is a
 frequently  cited investigator of the  effects of
 hyperthermia on the development of mammals. His
 extensive work on this agent as a cause of congenital
 malformations is perhaps best summarized in one of
 his reviews (Edwards  1974), which describes his
 experiments  on  the teratologic manifestations of
 hyperthermia in the guinea pig. The guinea pig is an
 unusual laboratory  animal in its  long gestational
 development,  approximately  65 days.  During the
 latter two thirds of  gestation  in the guinea pig, the
                                                                      5-35

-------
unborn offspring is in a stage of development similar
to that which occurs after birth in the mouse and the
rat. A long gestation period does not in itself prevent
use of the guinea pig as a model for teratogenesis. But
as far as  teratologic experimental  methodology is
concerned, the administration of the agent at 40 or 50
days of gestational  age  in  the guinea pig has no
equivalent in  rodent fetuses; this stage of develop-
ment occurs postpartum  in mice and rats.
Studies of the teratogenic potential of conventionally
induced hyperthermia in  almost all the species used
show some general changes that can be seen in RF-
radiation-treated animals as well. The symptom  of
hyperthermia, whether induced by  RF radiation  or
otherwise, causes varied anomalies (such as kinked
tail,  microphthalmia, exencephalia,  cleft palate,
general edema). This is evidence that hyperthermia is
a general teratogen. Perhaps the only case where RF-
radiation-induced  and conventionally  induced
teratogenesis  is not the same is in the development of
smaller brain  weights in fetuses heated by non-RF-
radiation-heating  agents. Except for  Shore et al.
(1977), perhaps  the literature  in RF-radiation-
induced teratogenesis has not been sufficiently
developed to  show  this  effect of decreased  brain
weight.
Conventionally induced hyperthermia,  like RF-
radiation  hyperthermia, appears to affect  most
species if  sufficient energy is applied and proper
timing of  the  agent is  obtained. A sufficiency  is
usually manifested by  a rise in maternal rectal
temperature, usually over 40°C, but often in a range
of 41 to 43 or  even 44°C. At these temperatures, the
teratogenic potential of heat  applied by conventional
means or by RF radiation does not appear to show any
significant  differences. The authors of  one study
attempted  to differentiate  teratogenic effects  of
microwaves from the more conventional methods of
heating (in this case, infrared irradiation). Chernovetz
et al. (1977)  exposed pregnant  rats to  2450-MHz
fields at an SAR  of approximately 31 W/kg, or  to
infrared radiation. The rectal  temperatures of the RF-
irradiated  and the infrared-irradiated  groups  were
similar (approximately 42°C). The results of the
exposures showed little difference in effects due to
infrared heating or RF-radiation heating.

The symptom of decreased body weight of fetuses is a
general developmental  effect.  This  response  to
microwaves is commonly seen in experiments, even
those experiments that have not otherwise demon-
strated specific morphologic changes (e.g., O'Connor
1980). Body weight changes alone  have also been
seen in mice fetuses at temperatures <40°C (Berman
et al.  1978).

Body  temperatures  reported in the above studies
were those of the dam, usually measured in the dam's
large intestine. Measurements of fetal temperatures
after RF-irradiation have not been reported. A report
by Morishima  et al.  (1975) provides some under-
standing of the physiologic changes that occur in the
fetus as a result of conventionally induced hyperther-
mia. Morishima et al. used pregnant baboons and
simultaneously monitored temperature and physio-
logic changes in  the dam and the fetus during
hyperthermia. The temperatures in these dams were
raised to 41 to 42°C by infrared lamps and warming
pads. While the pregnant dam was maintained in an
analgesic state using  nitrous oxide inhalation
anesthesia and intravenous succinylcholine chloride
infusions,  thermistors and catheters were placed at
comparable positions in the dam and the fetus. Two
groups of animals were used: one group was kept at
38°C; the temperature of the other was allowed to
elevate gradually to 41 or 42°C. At the more normal
temperature, 38°C, there was a slight (0.5°C) but
steady increment in the temperature of the fetus over
that of the dam. When the temperature was
increased to 40°C in the dam, the fetal temperature
rose accordingly and remained 0.75°C higher than
the dam's temperature. The hyperthermia produced
increased  uterine activity and some acidosis  in the
dam, and a profound acidosis and hypoxia in the
fetus. If conditions  comparable to those seen in the
baboon  are also seen in the mouse, then we might
expect the mouse fetus to have temperatures greater
than that of its dam.

That the teratogenic potential  of RF radiation might
depend only on the  deposition of energy as heat was
shown  experimentally by Rugh and McManaway
(1976).  They  exposed  pregnant mice to highly
teratogenic (high incidence of fetal death) levels of RF
radiation with and without pentobarbital anesthetic.
The  mice exposed during anesthesia had normal
rectal temperatures and normal incidences of fetal
death.  Therefore, the anesthetic protected against
the primary temperature effect of the radiation by
reducing  "...the body temperature to a degree
equivalent to the rise in temperature expected from
the conditions used." This result  is clear evidence
that the increased fetal abnormalities are strongly, if
not solely, associated  with   increased  maternal
temperature.

Further evidence for a direct  relationship  between
fetal effects and hyperthermia  induced by  RF
radiation is shown in several  reports (Dietzel  and
Kern 1970; Dietzel  et al. 1972; Dietzel 1975). In these
experiments, groups of pregnant rats were subjected
to 27-MHz radiation to raise rectal temperature to
42°C. From these data, the thresholds for increased
malformation rates appear to be between core
temperatures of 39.0%. maintained for  5 min and
40.5°C for 10 min.

Observed teratological effects are summarized in
Table 5-9.
                       5-35

-------
Table 5-9.   Summary of Studies Concerning Teratologic Effects of RF-Radiation Exposure

                                                         Exposure Conditions*
Effects
30% survival of pupae
Embryonic LDso


Decreased postnatal survival
Teratogenesis
No change in teratogenesis
Increased postnatal survival
Maternal lethality, resorptions,
decreased fetal weight
Decreased fetal weight
No change post-hatching: hatchability.
hemogram, body or organ weights
No change
No change
No change
Teratogenesis
No change
No change
No change
No change

No change
Decreased body and brain weight
Species
D. melanogaster
Chicken


Mouse
Mouse
Mouse
Mouse
Rat

Mouse
Japanese quail

Rat
Rat
Mouse
Japanese quail
Rat
Rat
Squirrel monkey
Rat

Rat
Rat
Intensity
(mW/cm2)

200
280
400






28



40
3. 4-14

5
28

10-35

5
10
Duration
(days x min)
1 x 10
1 x 12
1 x7
1 x4
1 x4
1 x2-5
1 x 10
1 x 10
1 x20

12x100
1 x 1440

1 x20
1 x120
12x100
12
Many x 240
12x100
Many x 180
Many

19x1200
16x300
SARt
(W/kg)
640
70
98
140
104
85-112
38
38
31

22
14

14
10
2-8
4
0.7-4.7
4.2
3.4
1-3.5

2.5
2.2
Reference
Payers/. (1978)
Carpenter et al. (1960a)


Rugh(1976a)
Rugh eta/. (1974, 1975)
Chernovetz et al. (1975)
Chernovetz et al. (1975)
Chernovetz et al. (1977)

Berman et at. (1978)
Hamrick and McRee (1 975)

Chernovetz et al. (1979)
Michaelson et al. (1978)
Berman et al. (1978)
McRee and Hamrick (1977)
Smialowicz et al. (1979a)
Berman et al. (1981)
Kaplan (1981)
Jensh et al. (1979)
Jensh (1979, 1980)
Johnson et al. (1978)
Shore et al. (1977)
"Frequency used was 2450 MHz. except for Jensh (914, 2450, and 6000 MHz), and Johnson et al. 1978 (918 MHz).
tFrom report or estimated from Durney et al. (1978)
5.3.2 Reproductive Efficiency
This section discusses aspects of reproduction where
the conceptus is not "insulted" directly by the agent,
although the fetus may be involved in demonstrating
the effect.  Generally, reproductive efficiency is the
capacity of the dam or sire to effect conception and to
bear and rear offspring. Changes in  this capacity
might be due to alterations in behavior, physiology, or
morphology.  For example,  reproduction  efficiency
might be affected by changes  in maternal cyclical
hormone secretions. Tests for reproductive efficiency
are not conducted as  frequently as those for
teratologic  effects,  usually because the  male and
female  reproductive systems require  considerable
alteration by  a toxic agent to cause significant fetal
wastage or increased reproductive efficiency. The
testes as  factors in reproductive efficiency are
discussed separately in the next section.

We have previously discussed the effects of 2450-
MHz RF radiation on D. melanogaster reported by Pay
et al. (1978). The fruit flies that were the subjects of
that study were irradiated before production of ova.
The SAR used in the study  ranged from 400 to 800
W/kg. Reproductive efficiency,  in this case the
production of eggs, was significantly reduced in both
conventionally heated and RF-irradiated females as
compared with the shams. But there appeared to be
no significant difference in the production of ova from
females exposed to RF radiation compared with those
exposed to the conventional  source of thermal
energy.
The egg production of chickens can also be used to
provide an index  of reproductive efficiency or
reproductive wastage. A single intense exposure of
day-old chicks was made in the experiment reported
by Davidson et al. (1976). In one of four experiments
they describe the reproductive efficiency of 28-week-
old hens that had been irradiated as chicks on day 1 of
age for 4.5 s at a power density of 800 W/cm in a
multimodal cavity.  The estimated SAR was 2770
W/kg. No differences appeared between the control
group and the irradiated group in the production of
eggs during 100 days of  laying.  The authors also
stated that a dose rate of 2500 W/kg for 9 s at 2450
MHz is a  lethal dose in day-old chicks and that 42
percent of the chicks dosed for 6 s at 2810 W/kg died.
The animals that survived were unconscious for upto
5 min.  The day-old chicks  had approximately the
same sublethal SAR (2770 W/kg for 4.5 s) used in the
examination for latent reproductive effects, but there
were no deaths that could be directly attributed to the
RF-radiation exposure. From these experiments, it is
concluded that the massive doses of this experimental
regimen left  no latent  alterations in reproductive
efficiency.
Rugh  et al. (1975) reported an experiment in which
they examined differences in average lethal dose of
RF  radiation (2450 MHz) during the estrous cycle of
CF1 mice.  The weights of the mice ranged from 29 to
31  g.  There was a  prior 20-min acclimation to the
waveguide exposure  chamber, and environmental
conditions were 23.5°C, 50 percent relative humidity,
and an airflow of 38 liters/min (0.38 km/h). There
                                                                       5-37

-------
was a significant decrease in the average lethal dose
for females in estrus as compared to the average
lethal dose for those in  diestrus (P < 0.01). The
forward power was 8.24 W. This experiment shows
that changes brought about during the reproductive
cycle can affect radiosensitivity.

In summary, it appears that  the  efficiency of the
female reproductive system is not easily altered. Only
irradiation at extremely high  dose rates has made
changes in reproductive patterns. Small alterations
from normal  values, though detectable by modern
scientific and statistical methodology,  do  not seem
able to sway the general outcome of the reproductive
cycle. Observed reproductive effects are summarized
in Table 5-10.

5.3.3 Testes

Considerable scientific work has  been done to
determine effects of RF radiation on the testes. The
testes are so placed anatomically that they can be
conveniently  irradiated  without irradiation of the
remainder of the  body. Tests for testicular function
are also conducted easily. Quantitative changes in
sperm  concentration can be conveniently assessed
by repeatable laboratory techniques. It is known that
when the  mammalian testes,  which have a normal
temperature of 33 to 35°C, are heated to temperatures
approaching abdominal temperature (37-38°C),
sterility can  occur.  Sterility consequent to  high
testicular temperatures can be viewed as a purposeful
contraceptive agent or as an unintended toxic agent.

Table 5-10.    Summary of Studies Concerning Reproductive
            Effects of RF-Radiation Exposure
                       Exposure Conditions'
Effects
Decreased ova
production
No change in egg
production
Lethality changes
with estrous cycle
Species
D
melanogaster
Callus
Mouse
Duration SAR
Idav x mm) (W kg)
40O-80O
1 x 0 08 2770

Reference
Pay ft al (1978)
Davidson et al
(1976)
Rugh et al (1975
"Frequency has 2450 MHz in all cases
A report by Muraca et al. (1976) describes the results
of  irradiation  of rat testicles with  2450-MHz
radiation. Each animal was anesthetized, placed in an
anechoic chamber, then irradiated in a free field at a
power density of 80 mW/cm2. The SAR for adult rats
was 16 W/kg (Durney et al. 1978).

In the experiment, the authors irradiated rat testes to
produce  increases of mtratesticular temperature to
36,  38,  40, and 42°C;  the testes  were then
maintained at these temperatures. The technique of
irradiation included implantation of a thermistor into
one testis  of  the  anesthetized male, with  the
temperature of that testis acting as the control of the
on-off-on sequence of the RF irradiator. Both testicles
were irradiated until the temperature of the testes
was within  ±0.5°C  of the levels mentioned above.
The duration of each irradiation at a power density of
80 mW/cm2 varied from 10 to > 70 min. The animals
were irradiated once or repeatedly during 5 consecutive
days. Five days  after the treatment endedr the
animals were killed and their bodies were infused
with solutions to preserve the testis that was not
"invaded" by the thermistor.  The microscopic
appearance of the testicular  tissue was categorized
as follows: apparently normal; appearance of early
inflammatory or degenerative  changes  in the
spermatogenic epithelium without well-developed
necrosis; or severe degeneration of the majority of
seminiferous tubules with coagulated cellular
elements.
When male rats were exposed once  (80 mW/cm2;'
2450 MHz; 10 to 73 min; temperature of the testis
maintained at 40°C), no significant change was
observed in the incidence of apparently abnormal
testicular tissue. After a single exposure  of 10 min,
during which the temperature of the testis was
allowed to reach 42°C, the number of animals that
had some abnormal changes tripled.  Multiple (five)
exposures, even where the temperature reached only
36°C for 60 min, caused all the testes to have some
changes  in the  spermatogenic epithelium. When
temperatures were  allowed  to reach 40°C in the
testes for repeated short periods (10 to 27 min, five
times), severe degeneration in the spermatic tubules
was seen in all the testicular samples.
This study points to two important factors of irradia-
tion of the testes with 2450-MHz RF radiation: that a
minimum temperature (>40°C) must be reached in
an acute exposure, and that repetitive treatments are
much more effective than single  ones. These two
factors also interplay, so that an acute temperature
excursion, even as high as 40°C for over 70 min, is not
as deleterious as a lower temperature (36°C) reached
repeatedly (five times).

Fahim et al.  (1975) conducted an experiment  to
compare the contraceptive capability of microwaves
with that of other methods of heating. The authors
used male Sprague-Dawley  rats (Holtzman strain)
and irradiated them with 2450-MHz radiation (from a
diathermy unit of maximum output of 100 W) during
pentobarbital anesthesia. The applicator  of the
diathermy unit was 7.5 cm from the testicles of the rat
and provided near-field exposures that do not allow a
good estimate of the associated SAR. By varying the
power output  of the unit and by varying the exposure
time from 1 to 1 5 min, the authors developed four
subgroups  of animals: those  in  which testicular
temperature reached 65°C for 5 min, 45°Cfor 15 min,
and 39°C for  1 or 5 min. The  males were  allowed to
mate with normal females 24 h after treatment, and
every 5 days thereafter until pregnancy was observed
in the females. 'The endpoint for fertility was the
amount of time required for every surviving male in
the treatment group to impregnate a female." Later,
                       5-38

-------
the sexual organs were weighed, and histological
examinations were made of the testes and secondary
sex organs.
Raising the temperature of the testes to 65°C for 5
min or to 45°C for 15 min caused complete infertility
in the males for 10 months. When these testes were
examined histologically, there was  no  observable
spermatogenesis. When the temperature was raised
to only 39°C, 70 percent of the males retained normal
breeding capability, whereas the remaining 30 per-
cent recovered their fertility within 2 wk. Histological
sections indicated  normal  spermatogenesis. There
appeared to be no differences in testicular weights or
secondary sex-organ weights, even  in the group
(45°C,  15 min) that was sterile 10 months after the
RF-radiation exposure had ended.

This experiment demonstrates that a temperature of
45°C  caused  by RF-radiation exposure  must be
attained in the  rat testes to produce permanent
sterility. Temperatures that ranged to 39°C, some-
what above rectal temperature, for only a short period
(5 min) were  not effective  and produced only
temporary,  if  any,  infertility. Along with the
permanent sterility induced by the high temperatures
(minimum of 45°C), there was damage to the tubules
and a lack of spermatogenesis.
Sterility can  occur in  rats  following RF-radiation
exposure. At  high temperatures,  epithelial  and
tubular structures  of the testes can be damaged
permanently. At lower temperatures (approximately
normal  body temperature)  for short  durations, the
temperature in the  testicles  due to RF-radiation
exposure resulted  in temporary sterility. This last
condition implies  that the  spermatogenic  cells
themselves are not damaged permanently. The rat's
spermatogenic cycle has a duration of about 10
weeks,  i.e., it takes  about 70 days  from  the  first
meiotic divisions of sperm generation in the germinal
epithelium until the time when the sperm are fully
matured, in place, and ready for ejaculation. During
the last 2  weeks  of spermatogenesis, sperm are
located in  the tubules  of the testes and in the
epididymis and require very little additional matura-
tion. There is a constant stream of maturing sperm,
because various seminiferous tubules are in different
stages of this long cycle. The most mature stages of
the spermatogenic  cycle are most sensitive to heat
(Van Demark and Free  1970). The permanent (10-
month) sterility seen in the experiment above is due
not only to loss of the most mature and most heat-
sensitive stages, but also to the loss of sperm  at all
stages, including the stages involving the germinal
epithelium. The 2-week temporary infertility seen in
the less affected males was caused by the loss of only
the most mature and most heat-sensitive sperm.

A study to demonstrate an alteration in functional
fertility was conducted by Berman et al. (1980). The
 authors' report concentrated on chronic exposures of
 rats to 2450-MHz radiation in an anechoic chamber.
 There were three separate experiments:
(1) exposure at a power density of 5 mW/cm2 for 4
   h/day, daily, from the 6th day of gestation through
   90 days of age postpartum;
(2) exposure at a power density of 10 mW/cm2 for 5
   h/day for 5 days, beginning on the 90th day of age
   postpartum; and
(3) exposure at 28 mW/cm2 for 4 h/day, 5 days a
   week, for 4 continuous weeks, beginning on the
   90th day of age postpartum.
 The main purpose of the experiments was to evaluate
 potential mutagenic effects of RF radiation on the
 •germ cells of the male rat. Male rats  were bred to
 untreated female rats shortly after the end of the
 treatment period. No mutagenic effect was demon-
 strated. The breeding data of the assays used in these
 experiments were used to evaluate the effects on the
 spermatogenic function in rats.

 In the first experiment, the animals were exposed
 daily from the 6th day of gestation to 90 days of age, at
 a power density of 5 mW/cm2. In this  experimental
 group, the SAR varied inversely with the growth of
 the  animals, i.e., from approximately  4.5 WAg in
 neonates to approximately 0.9 WAg at 90 days of
 age. Twin-well calorimetry was not used in the other
 two experiments, and SARs are estimated at 2 W/kg
 at a power density of 10 mW/cm2, and 5.6 W/kg at 28
 mW/cm2.
 In this study, temperatures  of the testes of rats
 exposed up to 90  min to 2450-MHz fields at 28
 mW/cm2 showed an increase from  a  pre-exposure
 level of  34°C to  almost 38°C after  exposure.
 Simultaneously recorded rectal temperatures were ~
 3°C  higher than temperatures of the testes during the
 entire period. Sham-irradiated animals had testicular
 temperatures  ranging from almost 32°C to approxi-
 mately 35°C. The exposure at 28 mW/cm2during the
 90-min period produced a temperature in the testis
 equivalent to the normal rectal temperature.
 Chronic exposure at 5  mW/cm2 from the 6th day of
 gestation through 90 days of age appeared to have no
 effect on the reproductive efficiency of  the male rats
 when bred with normal females. Also, exposure at a
 power density of 10 mW/cm2 (SAR ~ 2 W/kg) at 90
 days of age for a period of 5 days had no effect on the
 reproductive efficiency of the  males. It was only the
 most severe regimen,  the  exposure of adult male
 rats at 28 mW/cm2 for ~ 4 weeks (estimated SAR ~
 5.6 W/kg), that caused any alteration in reproductive
 function. This exposure produced a severe decrease
 in the reproductive ability of the  males;  only 50
 percent of the females that were available  to the
 males for breeding became pregnant during the week
 immediately following the exposure period. The
 breeding  returned to normal beginning at the third
 week after irradiation (the next period of test breeding).
                                                                     5-39

-------
The animals bred normally thereafter. No examination
was made of the  testes of animals exposed at 28
mW/cm2. The temperatures reached in the testes at
the highest power density (28 mW/cm2) were similar
to those reported by  Fahim  et al. (1975), where
histological changes were seen.
Varma and Traboulay  (1975) exposed anesthetized
male mice to 1.7-GHz or 3.0-GHz radiation at power
densities up to 200 mW/cm2for up to 100 min. The
results of the reported experiments are (frequency in
GHz; power density in  mW/cm2; SAR in W/kg, and
estimated from Durney (1978); exposure duration in
minutes;  and  effects  observed):  1.7 GHz,  10
mW/cm2, 15 W/kg, < 100 min, no effect; 1.7 GHz,
10 mW/cm2, 15 W/kg, 100 min, necrotic germinal
and tubular tissue, and intact interstitial and Sertoli
cells; 1.7 GHz, 50 mW/cm2, 75 W/kg, 30 to 40 min,
all  tissues  necrotic; 1.7 GHz,  200 mW/cm2,  300
W/kg,  20 min, skin burns; 3.0 GHz, 50 mW/cm2, 50
W/kg,  20 min,  minimal (sic)  injuries. In this
experiment,  as in  the earlier work described in the
document, acute effects are very dependent on the
combination of exposure duration and field strength.

Cairnie  et al. (1980a) used oil-bath immersion to
modify core and testis temperature. When conscious
mice were  immersed  in oil baths at temperatures
from 34 to 42°C, both  core and testicular tempera-
tures appeared to  be physiologically  regulated,
paralleled  the bath temperature,  and reached
equilibrium.  Anesthetized mice, on the other hand,
did not appear to demonstrate any core or testicular
temperature modification, but appeared, instead, to
increase temperatures  equally with the temperature
of the oil. This study assumes importance because it
clearly demonstrates that  the  temperature in the
testes  is not normally  regulated in anesthetized
animals. The investigators also studied the effects in
mouse testes after a range of whole-body exposures
of up to 30 days for 16 h/day at 2450 MHz and power
densities of up to 36 mW/cm2 (average whole-body
SAR =  7 W/kg; average testicular SAR = 14 W/kg).
These exposures caused no measurable  increase in
testicular temperature. No changes were seen in the
number of  dead  testicular cells, the  number of
epididymal sperm, or in the percentage of abnormal
sperm  after exposure, even  up  to 8 weeks  after
exposure. Four strains of mice were used, and no
strain susceptibility was observed.

Cairnie et al. (1980b),  in preparation for their future
work in the RF-radiation effects in the  testes,
determined  the dosimetry of  2.45 GHz in mouse
testes.  They found that the absorption  in the
abdomen in the area of the liver was 11 times greater
than in the testes when the body is oriented parallel to
the electric-field vector. They also found that though
abdominal temperature may have been increased by
exposure to 50  mW/cm2 for  16 h,  testicular
temperature was not.
Cairnie and Leach (1980) examined the viability of
testicular and  sperm cells and the morphology of
sperm cells after exposure to hot-water baths of the
posterior torso in mice. Unanesthetized mice were
partly immersed in water at temperatures of 32 to
43°C for up to 4 h. The results of this experiment are
appropriate to this discussion in that they may reflect
the type of response that testicular tissue and sperm
would give to  exposure to  RF radiation; however,
caution  should be used  in attempting to equate
durations of water immersion and  RF radiation
exposure, as  the site and time-profile of heat
deposition may not be equivalent. Cairnie and Leach
observed that heat-damaged testicular cells  are
evident within 2 h after exposure (41 °C, 30 min), that
the incidences  of  the damaged  cells are at a
maximum from 4 to 12 h after exposure, and that the
incidences of damaged testicular cells are related to
combinations  of exposure duration and bath
temperature. They also observed that a 30-min
immersion in 43°C water, but not lower tempera-
tures, significantly decreased total epididymal sperm
counts during a 10-week-long postexposure  period,
and that the incidences of abnormal-appearing sperm
followed a similar pattern.

In the"studies reported by Ely et al. (1964), 2880-MHz
(PW)  radiation was  used  to irradiate only  the
testicular area  in groups of dogs. The  animals were
anesthetized during exposures while  temperatures
were measured in the testis by a thermistor in a metal
needle.  Irradiation of the testicular area continued
until  a  peak temperature was reached;  the RF
radiation was then turned off manually so that the
testes could cool, at which point the RF radiation was
turned on once again.  This cycling of exposure was
completed many times so that the animal's testes
could be kept  at a "steady" temperature for a
considerable length of time. The normal temperature
of the testes of the dogs was ~ 33°C, 5°C lower than
that of the rectum.

According to this report,  exposure  of the testes
caused an elevated testicular temperature of 36°C at
20 mW/cm2, 38°C at 33 mW/cm , and up to 40°C at
45  mW/cm2.  As 38°C is  the  approximate body
temperature in the dog, bringing the testes to body
temperature can  expect to produce sterility if the
elevation is sufficiently chronic. It appears, then,
that 45 mW/cm2 is required to produce this  type of
effect.

There is a  large  body of literature on the fertility
effects caused by temperature increases, such as
when the testicles are clothed to prevent the normal
thermal radiation. Even high environmental tempera-
tures, when sustained, can produce infertility in male
rats. One example of  this  is a report (Pucak et al.
1977) describing deaths in  a large rat-production
colony in which  room temperatures accidently
reached as high as 31.6°C for 2 days and as high as
                       5-40

-------
37.7°C in  individual cages. Under these conditions,
approximately 3,000 of 14,000 Sprague-Dawley rats
died from  heat prostration. When examined 18 days
after the incident, 25 percent of the surviving males
showed bilateral atrophy of the testicles, which were
approximately half the normal  size. Histological
examination showed atrophy of the spermatic
tubules  and  failure of spermatogenesis. The
proportion of affected testicles ranged from 50 to 75
percent. Five weeks after the incident the animals with
small testes were still sterile. In comparison with the
temperatures  seen  in this  study,  raising  the
temperatures of  testes to 42°C by RF-radiation
exposure  appears  to  be an  extreme experimental
regimen. Observed testes effects are summarized in
Table 5-11.

5.3.4 Unresolved Issues

5.3.4.1 Teratology
A threshold of teratogenic effects due to RF radiation
has  not been determined. The measurement of
threshold  will have to include the entire range of
teratogenic effects: lethality,  anomaly production,
decreased body weight in fetuses, and alteration of
postnatal function. Even at this stage of development
of the data relating to RF-radiation-induced terato-
genesis, a picture is appearing in which the degree of
response depends on  the level of whole-body SAR,
the duration of exposure, the timeliness of the
exposure,  and the species, all of which complicate the
threshold  determination. It is  not yet clear whether
threshold  can be based on a finite whole-body SAR
for all species or whether  some  adjustment or
proportion will be necessary for species size,
metabolic  rate, thermoregulatory capacity, etc. So far,
the only physiologic variable that can be supportably
associated with teratogenic  effects is the  colonic
temperature  of the dam during or at the  end of
exposure.  The mouse and rat, two extensively studied
species, show a minimum temperature of approxi-
mately 40°C in the dam is associated with teratologic
symptoms. However, whole-body SARs required to
cause temperature excursions like this are much
Table 5-11.    Summary of Studies Concerning Effects of RF
   higher in the mouse than in the rat. Therefore, an
   adequate relationship of teratogenesis to SAR alone
   is not apparent. The  maternal colonic temperature,
   then, is the only available indicator of a threshold for
   teratogenesis in mammals.

   Published reports that meet the criteria for considera-
   tion in this document have limited their examination of
   the fetal results of gestational exposure to a gross
   morphological change or one  that might be seen
   under low magnification (15X). There has been no
   organized  attempt  to examine in histologic detail
   fetuses that have been irradiated in utero. Authors of
   one study (McRee  et al. 1980b),  however, made  a
   detailed examination  of embryonic hearts but could
   not demonstrate changes in morphologic, ultrastruc-
   tural,  or  enzymatic  activity. The subjects were
   Japanese quail which had been exposed daily for  8
   days to 2450-MHz radiation at SARs of 4 and 16
   W/kg.

   There  are  classifications of  fetal  changes that
   represent no gross structural deficit, but nevertheless
   represent some variation of structure. An example of
   a variation that  may not be considered by all
   terajologists as a "deficit" is a small but normal fetus.
   However,  if  the  incidence  of  this variation is
   consistently increased by the application of a toxic
   agent, it could be considered an  expression  of
   embryotoxicity. The decreased body weight so often
   seen in offspring exposed to RF radiation (Berman et
   al. 1978; Chernovetz et al. 1977,1979; Rugh 1976a)
   might otherwise be  considered as a lesser category of
   "structural  variation  without deficit" if decreased
   weight were not so consistent in these experiments.
   Whether  this decreased fetal weight is temporary
   (i.e., only a delayed growth that will disappear in the
   neonatal stage) or permanent (i.e., a  stunting of the
   fetus that will persist) has not yet been resolved.

   There  is  one aspect of the literature on the
   teratogenic potential  of RF radiation that  deserves
   further discussion.  More than half the papers in this
   document on teratogenesis report experiments with
Radiation Exposure in Testes of Mice and Rats
                                             Exposure Conditions
Effects
No change
Abnormal germinal cells.
normal interstitial cells
All tissue necrotic
Scrotal skin burns
"Minimal" injury
No change in tissue.
sperm
Abnormal spermatogenic
tissue
No change
No change
Temporary sterility
Species
Mouse




Mouse
Mouse

Rat

Rat


Frequency
(GHz)
1.7




3.0
2.45

2.45

2.45


Intensity
(mW/cm2)
10
10

50
200
50
<37

80

5
10
28
Duration
(days x min)
1 x<100
1 x 100

1 x 30-40
1 x20
1 x20
many x 1 6 h

1 x 10-73
5x 10-73
many x 240
5x360
20 x 240
SAR
(W/kg)
15
15

75
300
50
<8

16
16
0.9-4.5
2
5.6
Reference
Varma and Traboulay (1975)





Cairnie et al. (1980a)

Muraca et al. (1976)

Berman et al. (1980)


                                                                      5-41

-------
the rat. Chernovetz et al. (1977) raise an interesting
point  about using  the rat in determining the
teratogenic  potential of RF  radiation. The authors
argue  that because  levels of microwave exposure
which  are associated with high mortality rates of
dams do  not also produce fetal structural abnormal-
ities in the rat, "...that the teratogenic threshold of
microwave  radiation  is higher  than the  dam's
threshold of mortality," and hence"... that maternal
mortality is more probable than malformation of the
fetus, irrespective of the dose."
The problem of using animal models in determining
teratogenic potential in humans is that there is no
assurance that any of these models is a real estimate
of effects in humans. The concepts supporting the
use of  animals as models of human beings require
that the  response be demonstrated in a number of
species, so that the resultant generality can be more
confidently extended to humans. Therefore, we seek
among species some generality of effects of
microwaves on the fetus. Mice and rats are the two
laboratory animals upon which rest almost all of the
data of RF-radiation-induced teratogenesis. Any
difference  between  the  two species in  their
teratogenic response  to RF radiation, therefore,
becomes important.
the energy could not reach the prostate. This appears
to deserve additional attention and exploration.

The literature we have cited does not lend itself to
extrapolation of effects of RF radiation that cause only
small increases in the temperature of the testes. The
lack of data on RF-radiation effects at lower power
densities (which cause lower testicular elevations of
temperature than  have been cited in  the articles
above) is especially important in light of suggestions
in the popular media that thermal energy, in the form
of absorbed  RF  radiation, can be  used as a
contraceptive in men.
One report by Rugh (1976a) contains data on survival
in RF-radiation fields that are different in males and
females. In this study, mice of both sexes and ofthree
ages  (weanlings,  young  mature, and  aged) were
irradiated with 2450-MHz RF radiation until dead. Of
the three variables (age, sex, body weight) examined
for their contribution to RFrradiation lethality, Rugh
found that "The overall conclusion would be that no
matter at what age...absorbed  dose to  death...is
different for the two sexes. Females are slightly more
radiosensitive..." These results are  not explained
easily. Although they are not apparently related to
sexual function, they were included here on the basis
of sexual differences.
5.3.4.2 Reproductive Efficiency and Testes

The testes contain, besides sperm-producing tissues,
interstitial cells that secrete testosterone, the male
hormone. Gunn et al. (1961) described the effect of
24-GHz radiation on the morphology and function of
the interstitial cells in rats. They exposed rats once at
this frequency for a period of 5 min at a power density
of 250 mW/cm2, which  caused the temperature to
rise to 41 °C in the  testes.  As  a result, there were
scrota! burns, severe edema, and spermatic tubular
degeneration, but not interstitial-cell pathology.
The testosterone secreted by the interstitial cells of
the testes regulates zinc uptake by the dorsolateral
prostate. When Gunn examined zinc  uptake in RF-
irradiated rats that had no interstitial cell pathology,
he  found  a decreased  uptake. Gunn  related the
decreased function of this  secondary sex organ to
decreased testosterone secretion.

The  effects seen by Gunn were produced at  a
frequency at  24 GHz. At such a short  wavelength
there should be no significant penetration to affect
the dorsolateral prostate directly. The tests using the
dorsolateral  prostate as  an indicator of sexual
function in the rat have been commonly used by Gunn
and others in other types of  experimental situations.

It is not known why and how the dorsolateral prostate
lost its capacity to function normally when there was
no observable change in the interstitial  tissues, the
other testicular damage appeared to be minimal, and
                        5-42

-------
5.4 Nervous System
Michael I. Gage
Ernest N. Albert
Communication with the external environment and
modification of the operation of other organ systems
are functions of the nervous system. Signals from
stimuli in the  external  environment,  such as
chemicals, pressure, temperature, or radiant energy,
are  received  by specialized  receptor cells and
transduced into nerve impulses that are measured as
a complex pattern of electrical and chemical events.
These impulses are combined with other signals from
the internal environment, such as chemical secretions
or feedback from intermediate neurons or neurons in
efferent paths of the nervous system, at junctions
called  synapses. Information  is thus transmitted
through the  nervous system  until  some effector
change occurs (usually a chemical event at  a site
called  a neuromuscular junction) that activates or
modifies some gland or muscle tissue, and brings
about  an  integrated response to changes  in the
internal and external environment of an organism.
Detailed  descriptions of  the organization and
functioning  of the  nervous system are  available
(Crosby et al. 1962, Kandel  and Schwartz  1981;
Carlson 1980;  Cooper  et al.  1982; Scientific
American  1979).

Two types of cells are found in the nervous system,
neurons and glia. Neurons consist of cell bodies and two
types of peripheral processes known asdendritesand
axons.  They  are considered responsible for carrying
the information throughout the nervous system. The
glia (of which there are several types) are considered
to support the neurons, to nourish them, and react to
damage of nearby neurons.

The  nervous system can  be  divided into several
anatomical  subsystems,  that are not mutually
exclusive  but have been  formulated to unify a
particular function or anatomical similarity. One such
division is  the central nervous system (CNS, i.e., the
brain and spinal cord), as opposed to the remainder of
the nervous  system called  the peripheral nervous
system (PNS). The brain may be subdivided into areas
which  may  be further subdivided  into  nuclei or
cortices (the gray matter containing the cell bodies
and frequently the dendrites of neurons) and tracts (the
white matter containing the axons which are covered
with a white fatty tissue called myelin and in the PNS
are wrapped by Schwann cells). Neurons in nuclei
and cortices and axons in tracts are positioned so that
each area  contains  a topographic projection  of the
part  of the  peripheral  location connected by the
neurons.

There  is much  redundancy  within  the  nervous
system. Multiple pathways carry similar information
between the periphery and  the  brain,  as well  as
within the brain itself. Some of these paths, such as
primary sensory tracts, have few neuronal connec-
tions; others, such as the reticular formation, have
many intermediate (or internuncial) neurons. Input
neurons synapse  with intermediate and  output
neurons at numerous levels in the nervous system.

At synapses specific chemicals known as neurotrans-
mitiers  function  in complex series of reactions to
carry signals between  neurons. Packets of  neuro-
transmitter substance are released at the synapse by
one neuron and  taken  up at  the  postsynaptic
membrane of an adjoining neuron. Much evidence
indicates that norepinephrine, dopamine, acetylcho-
line, serotonin, and x-aminobutyric acid (GABA) are
neurotransmitters. The reactions at synaptic
connections  are  different from  another type  of
chemical activity that takes place at the membrane
along the axon at breaks in the myelin cover called
nodes, enabling  conduction  of electrical  impulses
along the axon. At these nodes influx and efflux of
ions of  potassium, sodium,  and  calcium  establish
changes in electrical potentials.

The  fine topographic  organization throughout the
nervous system, coupled with multiple connections
and  pathways between input and output and the
transmission of  specific chemicals  at particular
synaptic junctions, enables the specificity  of
response to ambient changes for which the nervous
system is renown. This exquisite  organization and
complexity within small areas in  the brain  is also
responsible for a variety of changes that  might  be
seen if RF radiation were to affect the nervous system.

A review of the literature of the effects of RF radiation
on the nervous system permits the following general
statements to be made:

• Acute or chronic CW or PW radiation of animals at
   SARs >  2 W/kg can  produce morphological
   alterations in the central nervous system. These
   changes are qualitatively similar after acute  or
   chronic exposure and  at different  SARs, but
   quantitatively  more alterations  occur  in the
   affected  neuronal  structure  after exposure  at
   higher SARs  and after chronic exposure. The
   changes are found less frequently if the animals
   are allowed to survive several days to weeks after
   exposure. Acute exposures above 2.5 W/kg alter
   electrophysiological responses in thethalamusof
   the cat brain.
   Currently, no information conclusively shows RF
   radiation affects the blood-brain barrier (BBB) at
   SARs below 2 W/kg. Initially it appeared that RF
   radiation at  low power densities  altered the
   permeability of the BBB in experimental animals.
                                                                     5-43

-------
   Later experiments indicated such alterations were
   due to thermal effects. Effects formerly attributed
   to changes in BBB permeability are now thought
   to be due to increased blood flow in the brain.

• RF radiation appears to have a potentiating effect
   on drugs that affect nervous system function. Most
   of the relevant studies have used pulse-modulated
   waves only.  Whether RF radiation also has an
   inhibiting effect on neuropharmacologic drugs is
   not certain.

• There are no data on the effects of RF radiation on
   neurotransmitters at SARs < 2 W/kg. However,
   after exposure at SARs > 8 W/kg the release of
   neurotransmitters appears to be affected.
   Whether there is  an  increased or a  decreased
   release depends on the specific neurotransmitter.

5.4.1 Morphological and Physiological
Observations
The nature of morphologic changes in the nervous
system of exposed animals depends on the frequency,
power density, duration, and modulation characteristics
(e.g., PW or CW) of the radiation. Gordon (1970) and
Tolgskaya  and Gordon (1973) reported severe
damage in the brain of rats after short exposure to RF
radiation of various  frequencies at high power
densities (> 40 mW/cm2, SAR estimated > 8 W/kg,
assuming  the  worst  case  condition which would
occur if the rats had their long body axis parallel to the
E-f ield vector) which produced rectal temperatures of
42 to 45°C. These changes consisted of hemorrhages,
edema, and vacuolation of  neurons after a 40-min
exposure to 3,000 or 10,000 MHz PW or CW radiation
at 40 to 100 mW/cm2. At 20 mW/cm2 similar but less
severe effects  were observed.  Further,  3000-MHz
radiation (SAR estimated about 4 W/kg) produced
more marked  changes  than  10,000 MHz (SAR
estimated about 3.2 W/kg) at equal power densities.
However, Austin and Horvath (1954) did not observe
similar  changes  in brains of rats that  became
convulsive and hyperthermic (rectal temperature
increase of 2.3°C, or  brain temperature increase to
about 43.7°C) during a single, short irradiation. The
exposure to 2450 MHz lasted for a maximum of 7 min
or until the onset  of convulsion (which was usually
less than 2 min). Exposure was of high intensity and
was presented only to the head (60 or 90 W applied
2.5 cm above or on the head). These authors observed
only mild pyknosis and hyperemia in some areas of the
brain,  mostly  in the  pyramidal cell  layer of the
hippocampus.

Albert and DeSantis (1975) did  not observe
hemorrhage, gliosis,  or focal necrosis in adult
Chinese hamsters  exposed to 2450 MHz  (CW) at 50
mW/cm2 (SAR estimated at 15 W/kg) in the far field
in an anechoic chamber for 30 min or single periods
lasting up to 24  h, but they did observe swollen
neurons with frothy cytoplasm  in the  hypothalamic
and subthalamic regions located near the center of the
brain.  Such observations were not seen  in the
cerebellum, pons, or spinal cord, which are located
more posterior than the other areas. These changes
were less severe in hamsters allowed to survive for 6
to 10 days after treatment. In a  second study (Albert
and DeSantis 1976) similar histologic changes were
seen in the hypothalamus and subthalamus  of
Chinese hamsters following single 1700-MHz (CW)
exposures lasting from 30 to 120 min at either 10
mW/cm2 (SAR estimated at 3 W/kg) or 25 mW/cm2
(SAR estimated at 7.5 W/kg) in the far field in an
anechoic chamber. No lessening of the severity of
these  changes was  seen in  animals allowed  to
survive for 13 to 15 days following exposure.

Histologic changes in the rat brain have also been
reported after multiple (35  or more) 30-min
exposures to 3000 MHz at lower power densities(< 10
mW/cm2, SAR estimated at 2 W/kg). The histological
alterations included  cytoplasmic vacuolation  of
neurons, axonal swelling and beading, and swelling
in and decreased numbers of dendritic spinesf Gordon
1970; Tolgskaya and Gordon 1973). These changes
were seen to a lesser degree or not at all in animals
allowed to survive for 3 to 4 weeks after exposure
before sacrifice. Vacuolation of neurons but not glia
was also seen in the hypothalamic region of Chinese
hamsters exposed to 2450 MHz (CW) at 25 mW/cm2
(SAR estimated at 7.5 W/kg) for 14 h on each of 22
days (Albert and DeSantis 1975).

Baranski (1972b) reported that exposure of groups of
30 guinea pigs to 3000 MHz (CW and PW at 400 Hz) in
an anechoic chamber at 3.5 mW/cm2 (SAR estimated
at 0.53 W/kg) and exposure of 20 rabbits to 3000
MHz (CW or PW not indicated)  at 5 mW/cm2 (SAR
estimated at 0.75 W/kg) for 3  h daily for 30 days
resulted  in  myelin degeneration and increased
proliferation of glial cells in both the cerebrum and
cerebellum.  More alterations  were noted  in the
guinea pigs exposed to PWthan those exposed to CW.
Temperatures recorded at unspecified points in the
body were  reported never to increase more than
0.5°C. Single 3-h PW or CW exposures of guinea pigs
at 3.5 mW/cm2 had no effects, but single exposures
at 25  mW/cm2 (SAR estimated at  3.75 W/kg)
produced edema, hyperemia,  and small  necrotic
lesions indicative of damage due to heat.


Switzer and  Mitchell (1977) found 3 times as many
myelin figures in the dendrites of brain neurons of 15
female rats repeatedly exposed to 2450-MHz (CW)
fields in a multimodal cavity than in the 14 sham-
irradiated control rats. Exposures lasted 5 h, 5 days
weekly for 22 weeks (a total of 550 h at an average
                       5-44

-------
 SAP calculated to be 2.3 W/kg, with Ta = 24 ± 1.5°C,
 RH = 50 ± 10 percent) and were followed by a 6-week
 recovery period before the euthanasia. These authors
 did not  observe other changes such as gliosis,
 perivascular edema, or synaptic pathology. However,
 the irradiated rats  in this  study exhibited marked
 disruptions of behavior during the exposure  period
 (Mitchell et al. 1977; see also Sec.  5.5).

 Qualitatively, the morphological effects on the CNS
 are similar in the range of 10 to 50 mW/cm2  (SARs
 above 2 W/kg), but effects are quantitatively greater
 at the higher power densities. Many scientists  would
 consider that irradiation at these power densities can
 raise body  temperature. Soviet  scientists have
 reported similar morphological changes  at  power
 densities less than 10 mW/cm2 after chronic
 irradiation; they do not consider these alterations to
 be of thermogenic  origin.  Tolgskaya  and Gordon
 (1973)  and Baranski (1972b) further state that
 morphological effects are more marked after PW than
 after CW and after chronic exposure than after acute
 exposure. Most Eastern European studies  claim full
 recovery of irradiated animals in  1 to 3 weeks after
 exposure at less than 10 mW/cm2 (SAR of 2 W/kg).
 Albert and DeSantis (1975, 1976) found continued
 existence of  neuronal  cytopathology in animals 2
 weeks after exposure.  Perhaps  a  longer recovery
 period in the latter study would have shown complete
 reversibility. We  can  conclude  that RF  radiation
 causes morphological changes  in the CNS  of
 experimental animals following  acute or chronic
 exposures at SARs of 2 W/kg. In most of the lower-
 intensity exposure studies,  effects were  rarely
 observed or were less severe in animals allowed to
 recover for 3 or more weeks after exposure.

 Electrophysiological  recordings from the brain and
 spinal cord of animals have been made during and
 following RF radiation when precautions have been
 taken to minimize artifact due to electrodes  in the
 microwave field. Johnson  and  Guy (1972)  have
 shown that a metal electrode of the type commonly
 used in neurophysiology can greatly increase the
 amount of energy absorbed in neighboring tissue by
 several orders of magnitude. However, saline-filled
 glass electrodes did not alter thermograms showing
 RF power absorption in brain tissue. Averaged evoked
 potentials were measured with glass electrodes in the
thalamus to somesthetic stimulation of the contra-
 lateral  forepaw (Johnson  and  Guy 1972). Cats
(anesthetized with alpha-chloralose and immobilized
with gallamine triethiodide) were exposed to 918
MHz (CW) for 15-min intervals at power  densities
 ranging  from 1  to 40 mW/cm2 with a microwave
applicator located 8 cm from the head and directed so
maximum intensity was aimed at the thalamic region
of the brain. The SAR in the thalamus of a live cat was
 1.88 mW/cm3 (~ 1.88 W/kg) for  an incident power
 density of 2.6  mW/cm2. The exposures decreased
 latencies between the stimulus presentation and the
 peaks of the  later components of the evoked
 potentials as  a  function  of the power density
 (threshold of 2.5 to 5 W/kg). Latencies between the
 stimulus  and  initial thalamic response were not
 decreased by the RF radiation. This result indicates a
 change in the multisynaptic paths to the thalamus
 and within the brain with little change in the primary
 sensory pathway. Although  microwave exposure
 produced increases in thalamic and body temperatures,
 similar increases  in body temperature produced by a
 hot pad decreased latency  of both initial and later
 components of the thalamic evoked potential.

Monosynaptic ventral  root  reflex responses of the
spinal cord to electrical stimulation of the ipsilateral
gastrocnemius nerve of cats anesthetized with pento-
barbital and immobilized by gallamine  triethiodide
were measured with a polyethylene suction electrode
during and after brief exposure to microwaves (Taylor
and Ashleman 1975). Exposures of 3 min at 2450 MHz
(CW)  were delivered by  a  parallel-plate applicator
 surrounding the spinal cord.  Decreases in latency and
attenuation in amplitude to zero of the monosynaptic
 reflex  occurred during exposures with  an  incident
power of 7.5 W, resulting in an absorbed power of 1.6
W/cm3 (~  1600 W/kg) at the  spinal cord.  The
temperature in  a bath surrounding the cord  was
37.5°C. Similar  but  less  marked changes were
observed during 3.75-W exposures. These  changes
were  reversible with  the termination of exposure.
Increased amplitude  of  the  monosynaptic  reflex
produced by cooling the spinal cord was reversed by
 microwave radiation. Raising spinal cord temperature
by heating the bath surrounding  the cord produced
effects similar to those seen during  microwave
 radiation.

Takashima  et  al. (1979) exposed male  rabbits
between parallel  plates to RF fields having carrier
frequencies  ranging from 1 to 30 MHz that were
amplitude  modulated at either 15 Hz or 60 Hz during a
single 2-to 3-h exposure or  during 4 to 6 weeks of
chronic exposure.  The electric field strengths ranged
from 60 to  500  Vrms/m. Acute  and chronic EEG
readings were obtained from animals under sodium
pentobarbital anesthesia. There was no temperature
rise in the exposed animals. The EEG recordings after
acute exposure at field strengths of 60 to 500 Vrms/m
or after chronic exposure at strengths up to 70 Vrms/m
showed no difference  between control and experi-
mental animals.  However, chronic irradiation at
higher field strengths was associated with abnormal
patterns,  consisting of bursts of high amplitude
spindles at 90  Vrms/m, as  well as suppression of
activity at 500 Vrms/m.  All brain activities returned to
normal a few hours after irradiation. The results in
this study appear to be free  of electrode  artifacts
                                                                         5-45

-------
because chronic exposures  were made  without
electrodes, and all recordings were made when the
fields were switched off. The effects of anesthesia are
not clear, and the natural fluctuations of brain activity
also complicate  interpretation of the  results.
Although the occurrence of high-amplitude spindles
in irradiated animals in this study is similar to that
described by Bawin et al. (1973), comparison  of the
two  results is difficult because in the latter  study,
chronically implanted electrodes may have interfered
with the  imposed fields. Takashima et al. (1979)
confirmed the existence of artifacts during irradiation
when implanted electrodes were used.

Bawin et al. (1973) exposed  adult female cats
previously implanted with steel electrodes and held in
a fixed position by a wooden stereotaxic frame in a
parallel-plate  exposure system  to  147-MHz  fields
amplitude modulated at frequencies ranging from 1
to 25 Hz at intensities  of 1 mW/cm2 or less (SAR
estimated about 0.015 W/kg).  Five cats were trained
(by operant behavioral reinforcement techniques) to
increase the percentage of time during which specific
brain waves were within a narrow frequency band
specified by the experimenter. Then reinforcement
was  discontinued for a series of sessions until the
percentage of brain  waves of the  reinforced
frequency returned to preconditioning operant levels.
During these extinction sessions, three of  the cats
were exposed to the RF radiation modulated at 4.5 Hz
for one cat, 3 Hz for the second, and  14 Hz for the
third. The modulation frequency was within the range
reinforced for each cat. Irradiated cats required more
sessions for the brain activity  to return to precondi-
tioning levels than  nonirradiated cats.  Spectral
analysis of the brain waves of irradiated cats showed
a shift such that the predominant frequency centered
around the modulation frequency of the RF radiation.
No such spectra were seen in irradiated cats that had
no behavioral training. In the same study two other
cats were reinforced by amplitude-modulated
radiation. The production of specified frequencies of
brain rhythms was followed by irradiation. Increases
in the number of bursts of the reinforced frequencies
occurred,  but the duration of these bursts did not
increase. The number of bursts returned to precondi-
tioning levels when the irradiation was discontinued.
5.4.2 Blood-Brain Barrier Studies
A  separation between the  blood and the central
nervous system (including the ventricles containing
the cerebrospinal fluid) limits the ready passage of
certain substances from the blood into the nervous
system. Such a separation acts to protect the brain
from foreign and  therefore  toxic substances  but
allows entry of certain  molecules  necessary  for
metabolism. This separation is an anatomical entity
consisting of special cells that have tight junctions
between them, as well as a functional property of
some glial cells. A thorough discussion of the blood-
brain barrier has been presented by Rapoport (1976).
In the past few years, contradictory reports have been
published concerning the effects of RF radiation on
the permeability of the BBB. (For reviews, see Albert
1979a and Justesen 1980).

Frey et al. (1975) were the first authors in the United
States to report a permeability increase in the BBB of
the rat after RF-radiation exposure. They observed
that a 30-min  1200-MHz exposure  (CW) at 2.4
mW/cm2 (SAR estimated at 1.0 W/kg) resulted in a
statistically significant increase  in fluorescein in
brain slices of experimental animals over controls.
Most of the fluorescein appeared to be concentrated
in the vicinity of the lateral and third ventricles. Some
dye also was detected in the metencephalon. The
authors  also reported  similar  but heightened
alterations  in the permeability of the BBB when rats
were irradiated with PW radiation at 2.1  mW/cm2
peak and 0.2 mW/cm2 average power density (SAR
estimated at 0.8 W/kg). Their results also indicated
that PW radiation was more effective in altering brain
permeability than CW.

Merritt et al. (1978) were  unable to  replicate the
fluorescein studies of Frey  et al.  (1975). However,
increased brain permeation of fluorescein-albumin
(mol. wt. 60,000) was produced in rats heated to40°C
by hot air or by RF radiation. They concluded that
hyperthermia per se, and not field-specific effects of
RF  radiation,  is  the essential  determinant of
increased permeability. With sodium fluorescein and
Evan's blue, Lin and Lin (1980) also found no change
in BBB permeation after a  single  20-min  focal
exposure within the rat head at 0.5 to 10OOmW/cm2
(local SARs ranged from 0.04 to 80 W/kg) at 2450
MHz (PW). In 1982, these authors reported increased
BBB permeability in the rat exposed similarly btit at an
SAR of 240 W/kg (in the brain); the brain temperature
was 43°C (Lin and Lin 1982).

Electron microscopic tracer methodology has been
used to  follow  the  movement of horseradish
peroxidase into Chinese hamster brains after 2450-
MHz (CW) radiation in the far field at 10 mW/cm2
(SAR estimated at 2.5 W/kg) for 2 h (Albert 1977;
Albert and  Kerns 1981), and into rat  and  Chinese
hamster brains after 2800-MHz (CW) radiation at 10
mW/cm2 (SAR estimated at 0.9 W/kg for rats and 1.9
W/kg for Chinese hamsters) for 2  h (Albert 1979b).
Horseradish peroxidase  is  an enzyme (mol. wt.
40,000) that normally does not enter the brain. Focal
areas of dark-staining material indicating increased
permeability of peroxidase were seen in about 35
percent of the irradiated animals in contrast with
about  10 percent of the controls.  Oark-stainad
particles were seen in a variety of brain areas, but
                       5-46

-------
appeared with greater frequency in the thalamus,
hypothalamus, medulla, and cerebellum than in the
cortex or hippocampus (Albert and Kerns 1981).  In
two of the above studies (Albert 1979b; Albert and
Kerns 1981), fewer of the animals that were allowed
to survive for 1 h following irradiation and almost no
animals that were allowed to survive for 2 h following
irradiation showed evidence of peroxidase in areas
other than those that are leaky in all animals due to
the normal absence of  the  BBB. This finding
demonstrated complete  reversibility of  the RF-
radiation effect. The increased permeability of the
BBB  appeared to be  due  to increased pinocytotic
transpdrt of the tracer rather than to opening of the
endothelial tight junctions (Albert and Kerns 1981).


Sutton and  Carroll  (1979) produced a change  in
permeability of the BBB to intravenously administered
horseradish  peroxidase by 2450-MHz (CW) micro-
waves of sufficient power (10 to 20 W) to raise and
maintain brain temperatures of male rats to40,42, or
45°C for periods  lasting from 10 to 120  min.
Exposures were limited to the head region by the use
of type A or type B applicators and by shielding of the
remainder of the body with an absorbent collar.
Increased residual peroxidase activity in brain was
found after 10-min exposure at45°C, after 15 min at
42°C, or 60 min at 40°C. If body core temperature
was  maintained at  30°C  during exposure by
precooling  the rat  before head-only irradiation,
peroxidase activity increased in the brain after 15 min
at a brain temperature of 45°C, 30 min at 42°C, and
180 min at 40°C. Precooling the rats increased the
exposure time needed to eliminate BBB integrity, and
also increased the survival time of the rats irradiated
at the lowest level from 2 to 3 h. This study indicates
that severe hyperthermia  induced  by the radiation
produces the disruption in the BBB and that this
disruption can be prevented or retarded by perfusion
of the brain by blood cooled in passage through the
remainder of the body.
Oscar and Hawkins (1977) exposed rats injected with
radioisotope  tracers to 1300 MHz (CW or PW)
radiation for 20 min.  Using the technique of
Oldendorf (1970), they found after CW irradiation at 1
mW/cm2 (SAP estimated at 0.4  W/kg)  a greater
uptake of radiolabelled mannitol (mol. wt. 182) and
inulin (mol.  wt.  5000),  but not dextran  (mol. wt.
60,000), in brains of exposed animals. Similar, but
greater, uptake of these compounds was observed
after PW irradiation (average power density 0.3
mW/cm2; SAP estimated at 0.1 W/kg) than after CW
irradiation.  Uptake of  mannitol by the brain was
dependent  on power density, pulse width,  and
number of pulses per second.  Merritt et al. (1978),
who also used the Oldendorf technique, reported no
significant change in uptake of mannitol or inulin in
rats exposed to RF radiation under conditions similar
to those used by Oscar and Hawkins (1977).

Preston et al. (1979) used the Oldendorf technique in
a study of rats exposed to 2450-MHz (CW) fields for
30 min at zero or one of five power densities ranging
from 0.1 to 30 mW/cm2 (SAP estimated at 0.02 to 6
W/kg) in the far  field  in an anechoic  chamber
(Ta=22±1°C). No change in uptake of mannitol intothe
medulla,  cerebellum, diencephalon,  or cerebral
cortex of the brian was found. Intracarotid infusion of
propylene glycol, however, did  increase brain uptake
of mannitol.  They speculated that the changes
reported by Oscar and Hawkins (1977) may have been
due to changes  in blood  flow. Later, Oscar  et al.
(1981) measured the blood flow  in several  brain
regions during exposure  to 2800-MHz (PW) fields at
15 mW/cm2 (average) for 5 to 60 min and found
increased local blood flow. They then suggested that
previously reported BBB permeability changes(0scar
and Hawkins 1977) may be smaller than originally
indicated.

A later report by  Preston and Prefontaine (1980)
described  studies  on BBB permeability in  rats
exposed to 2450-MHz (CW) radiation in both the near
and the far field. The exposure in the far  field in an
anechoic chamber  was at 1  or 10 mW/cm2 (SAR
estimated at 0.2 or 2.0 W/kg)for 30min (Ta=22±1°C).
In the near-field exposure, a microwave  applicator
was placed on the rat  head  for a  single 25-min
irradiation at 7,28, or 140 mW forward power (SAR of
the head, estimated as 12.64 W/kg for each watt of
forward power, was 0.08 to 1.8 W/kg). In the near-
field study the exposure took place after radiolabelled
sucrose was injected into the animal so that BBB
function during irradiation could be examined.  No
change in permeation was found in either study.


In summary, some initial reports indicated that direct
effects of RF radiation in experimental animals might
result in increased permeability of the BBB(Freyef al.
1975; Oscar and Hawkins 1977; Albert and DeSantis
1976;  Albert  1979b).  Other reports indicated that
increased  permeability might be mediated  by
hyperthermia  induced by intense RF fields (Sutton
and Carroll 1979; Merritt et al. 1978; Lin and Lin
1980). Preston  et al. (1979) and Preston and
Prefontaine (1980) reported negative findings at
lower power densities. Some of these discrepancies
may be attributed  to differences in  techniques
employed to assess changes of permeability. These
methodologies  consisted  of gross examination of
brain  slices, fluorescence  observations, and mea-
surement of single-passage isotopic tracers and
electron microscopic tracers. All these techniques
have  some inherent shortcomings, either in
quantitation or  sensitivity. Some of the deficiencies of
these  methods  have been  recently  reviewed
                                                                     5-47

-------
(Blasberg 1979; Segal and Magin 1983). Thus, unless
the  changes of permeability  are diffuse and
significant,  positive results  may not  be readily
apparent. Further complications with the interpreta-
tion  of the data lie in  reports that average power
density,  peak  power,  and  pulse width may be
important variables affecting the BBB permeability
(Frey et at.  1975; Oscar and Hawkins 1977).
Therefore, one must consider the limitations of the
techniques  and the exposure parameters before
reaching conclusions regarding effects of RF  radia-
tion  on the BBB.

5.4.3 Pharmacological Effects
RF radiation  has been  reported to alter effects of
drugs that influence CNS functions.  Baranski and
Edelwejn (1968) noted altered EEG  tracings and
increased effects of Cardiasol, a CNS stimulant, in
persons working in microwave fields. To  understand
these observations better they conducted experiments
in rabbits. Administration of  4  mg/kg Phenactil
(chlorpromazine), a  depressant of cortical  activity,
followed  by 3000-MHz (PW) irradiation at 20
mW/cm2 (SAR estimated at 3.0 W/kg)  for 20 min,
resulted  in  desynchronization of the EEG, and
reversal of the  synchrony of  the EEG  seen  after
Penactil alone. According to Baranski and Edelwejn
(1968) this  result  indicated that microwaves
stimulated the brain-stem reticular formation (a path
of multiple short neurons lying in the middle of the
brain), which was inhibited by the chlorpromazine.
Administration of 3 mg/kg Cardiasol (pentylenetetra-
zole,  a  CNS stimulant and  convulsant) following
irradiation or sham exposure of these rabbits resulted
in an EEG change only  in the exposed animals. The
authors concluded that  microwaves potentiated the
effects of the pentylenetetrazole through pathways
through the center of the brain. Chronic  exposure of
rabbits (Baranski and Edelwejn 1968,1974) to 3000-
MHz (PW) radiation at 7 mW/cm2 (SAR estimated at
1.0 W/kg), 3 h/day for 23 to 26 days (a total of 70 to
80 h) resulted in convulsions that were more violent
than after single exposures or in control rabbits. This
result indicated that microwaves acted on the same
thalamic  reticular formation areas  as did the
pentylenetetrazole. Chronic RF exposure also
resulted  in  desynchronization and high-amplitude
recording  potentials  in the EEG. In these  studies,
thermal  effects of microwaves were  considered
unlikely.  However, it should be noted that the EEG
records  were  obtained with screw  electrodes
implanted into the skull of the rabbits and that the
pulse modulation characteristics were not specified
in these experiments.


Servantie et  al.  (1974) investigated the convulsive
effects of a  50-mg/kg intraperitoneal  injection of
pentetrazol (pentylenetetrazole) following chronic
exposures of mice to 3000-MHz radiation (PW, peak
power of 600 kW, average power of 350 W, pulse
duration of 1 yus, repetition rate of 525 Hz). The mice
were exposed in groups in the far field in an anechoic
chamber in front of a horn antenna for either 8,15,
20, 27, or 36 days (for an unspecified duration each
day).  The mean  power density measured in the
absence of the  animals was 5  mW/cm2 (SAR
estimated at 5 W/kg). RF radiation affected the .time
to onset of convulsions  and the mortality  rate.
Nonirradiated mice had a  biphasic distribution of
convulsion latency. Mice were equally distributed
between groups with a short and long onset time.
After  15 days of irradiation, a greater proportion of
mice had longer latencies to start pentylenetetrazole
convulsions. However, after 20,  27, or 36 days of
exposure,  a greater  number of mice had  shorter
latencies to convulsion.  Mice  irradiated for only 8
days were not  different from  control populations.
Increased incidence of mortality following convulsions
was observed in  the groups of mice irradiated for
more than 8 days.

Servantie et al. (1974) also investigated the effect of
curare-like drugs on rats and neuromuscular
preparations  of rats  irradiated for 10 to 15 days,
presumably under conditions described above (SAR
would be  1 W/kg). They found that irradiated rats
were less susceptible to the paralyzing drugs. Similar
findings were noted in  sciatic and phrenic nerve
preparatons  from exposed rats. Phrenic  nerves
isolated from irradiated rats were  paralyzed to a
lesser extent and  recovered sooner than those from
control rats.
Goldstein and Sisko (1974) investigated the gross
behavior and EEG of rabbits given pentobarbital and
then exposed for 5 min to 9300 MHz (CW) in an
anechoic chamber at power densities ranging from
0.7 to 2.8 mW/cm2 (SAR estimated at 0.1  to 0.3
W/kg, assuming the rabbits were oriented with the
long axis of their body parallel to the E-field vector).
Five minutes prior to irradiation or sham exposure the
rabbits were  injected intravenously with 4 mg/kg of
sodium pentobarbital. No difference was seen in EEG
patterns between control and  irradiated rabbits
during the 5-min exposure period. However, after a
latent  period lasting from 3 to 12  min following
exposure, periods of  EEG and behavioral arousal
lasting as long as 16 min and followed by periods of
sedation occurred in the  animals exposed to  RF
radiation after the barbiturate, but not in the controls.
Such effects were seen  after the  lowest  power
densities  investigated but were more  pronounced
after  exposure at 2  mW/cm2.  Similar effects  of
arousal after irradiation were  also seen in rabbits
injected with hallucinogens or  morphine in place of
the pentobarbital before exposure. In this study the
EEG was recorded with the aid of stainless steel
electrodes implanted before the irradiation.
                        S-48

-------
A related finding, that of decreased sleeping time in
rabbits  following injections  of  22 mg/kg sodium
pentobarbital during  2450-MHz or 1700-MHz (CW)
RF irradiations, has been  reported  (Cleary and
Wangemann 1976). Exposures were in the far field in
an anechoic chamber (Ta=22±0.6°C,  RH=40  to 60
percent). The exposure level for this effect was 5
mW/cm2 (SAR estimated at 0.8 W/kg) at 2450 MHz
and 10 mW/cm2 (SAR estimated at  1.3 W/kg) at
1700 MHz. Effective exposures were accompanied by
increased rectal temperatures. However, pentobarb-
ital-injected rabbits kept at an elevated  room
temperature (39°C),  which resulted in similarly
increased rectal temperatures (+1°C), showed an
insignificant reduction in sleeping time.

Recently, Thomas et al. (1979) reported that acute,
low-level (av 1 mW/cm2)2450-MHz(PW) radiation in
the  near  field  (SAR estimated at 0.2 W/kg)
potentiated the behavioral response to chlordiaze-
poxide  (a  tranquilizer) in rats.  This and  other
behavioral studies are discussed in Sec. 5.5.5,
Behavior, Interactions with Other Stimuli.

It can be concluded from pharmacologic studies that
low-level microwaves may  elicit a drug-specific
interaction on the nervous system. Such interactions
may prove to be potentially both useful and harmful
as more information becomes available.

5.4.4 Effects on Neurotransmitters
Specific neural  systems  that contain various
neurotransmitters are known to affect the inhibitory
or excitatory states of the brain. The  relative firing
rates of these neuronal systems are reflected  in the
turnover of their neurotransmitters. Since RF radiation
has been reported to stimulate and depress the CNS,
several scientists have investigated the effects of RF
radiation on CNS neurotransmitters.

Snyder (1971) made neurochemical measurements
in rats  exposed  to 3000-MHz (CW)  radiation.  He
observed that a 1 -h exposure at 40 mW/cm2 (SAR
estimated  at 8  W/kg) resulted  in a. significant
increase of 5-hydroxyindolacetic acid (5-HIAA) and 5-
HT (serotonin), indicating increased turnover rates of
serotonin in the brain. The opposite effect (that is,
reduced levels of 5-HIAA and  5-HT, indicating
reduced turnover rates of serotonin) was found in
rats  exposed at 10 mW/cm2 (SAR estimated at 2
W/kg) 8 h/day for 7 days. The body temperature in
the rats exposed at 10 mW/cm2 rose  by 1 to 2°C
during irradiation, and the animals showed signs of
moderate heat stress. Control  rats were similarly
handled and restrained but not irradiated. The effects
of the 10-mW/cm2 exposure were compared to those
of conventional radiant heat loads by elevation of body
temperature of rats by 1 to 2°C (by placement in an
incubator maintained  at 34°C  by thermostatically
controlled incandescent lights) for 8 h/day for 7 days.
No difference was found in the  turnover  rate of
norepinephrine or 5-HT, or steady-state levels of 5-
HIAA,  between conventionally heated and  control
animals. Snyder concluded that exposure to RF radia-
tion produced distinctly different effects on 5-HT and
5-HIAA in rat brains from effects produced  by
conventional heating.

Zeman et al. (1973) investigated the effects of acute
and chronic exposure of rats to 2860-MHz (PW, peak
power of 60 kW, average power of 300 W, pulse
duration of 1 fjs. repetition rate of 500 Hz) radiation on
brain GABA. Chronic exposures at 10 mW/cm2 (SAR
estimated at 2 W/kg) lasted for 8 h daily for 3 or 5 days
or lasted 4 h for 5 days per week for either 4 or 8
weeks. No  significant  increase in body temperature
was measured in randomly selected rats immediately
after  exposure during the chronic series. Acute
exposures at 40 mW/cm2 (SAR estimated at 8 W/kg)
lasted 20 min, and acute exposures at 80 mW/cm2
(SAR of 16 W/kg) lasted 5 min. Rectal temperature
increased by no more than 3°C, and signs of general
hyperthermia were observed after exposures to
either  of these acute conditions. There were  no
significant differences in whole-brain GABA levels or
in L-glutamate decarboxylase (the enzyme synthesiz-
ing GABA) activity between control  and irradiated
animals after either chronic or acute exposures.

Merritt et al. (1976) reported decreased norepineph-
rine, dopamine, and serotonin in discrete areas of rat
brains  after a whole-body exposure for 10  min to
1600-MHz  microwaves  at  80 mW/cm2  (SAR
estimated at 24 W/kg) that raised rectal temperatures
4.1°C.  Hyperthermal control rats were kept at 78°C
for 10 min, which was  sufficient  to raise  rectal
temperatures 3.7°C. Hypothalamic  norepinephrine
and dopamine were significantly  decreased  in
irradiated rats and were  decreased (although less
severely) in hyperthermic control rats, as compared
with normotheric controls. Levels of serotonin in the
hypothalamus  and striatum of the  brain  were
unchanged, but serotonin levels in the hippocampus
were decreased only in irradiated rats, and serotonin
levels in the cerebellum and cortex (areas usually low
in serotonin) were  decreased only in  hyperthermic
control rats.  Merritt et al. (1976) concluded that
hyperthermia  was responsible for these effects on
neurotransmitters.


In a separate study, Merritt et al. (1977) observed a
significant decrease  in  norepinephrine and an
insignificant decrease in dopamine in the hypothala-
mus of rats exposed for 10 min to 1600-MHz (CW)
radiation at 20 mW/cm2 (SAR estimated  at 6.0
W/kg). but not at 10 mW/cm2 (SAR of 3 W/kg). The
effective  exposure  produced increased brain and
rectal temperatures, whereas exposure at the lower
                                                                     5-49

-------
power density did not. Serotonin inthehypothalamus
was unaffected by exposures up to 80 mW/cm2 (SAR
of 24 W/kg).

Nervous-system effects are summarized in Table 5-
12.  There appear to be  ample data that  suggest
effects of high  power densities of RF radiation on the
nervous system of animals. It may be that most of
these effects are the result of thermal effects on the
tissues.  The effect of RF radiation on calcium ion
efflux from brain and other tissues is discussed in
Sec. 5.7.5.


5.4.5 Unresolved Issues

There are several difficulties in evaluating alterations
in function of  the nervous system that result from
exposure  to RF  radiation.  Minute changes in the
energy  distribution due to focusing or scattering
effects could produce changes that may be reflected
in various parts of the body. Electromagnetic energy
may affect the nervous system in many ways. The
electrical and magnetic  fields surrounding  the
neurons might be directly alterable by externally
imposed fields. Radiation may produce changes in
chemical reactions that could result in a variety of
effects.  In addition, small  differences in  size and
shape among individuals could  result  in  quite
different manifestations of exposure to a given power
density  of RF radiation. Observations of anatomical
changes must  be made at a time quite removed from
the exposure.  Moreover,  it is  labor  intensive to
localize where small anatomical damage might occur.
Neurophysiologists typically record patterns of
electrical activity in either  individual neurons or
groups of neurons by inserting metal electrodes fn the
nervous system or on the body and amplifying the
small signals recorded by sophisticated electronic
apparatus. The use of metal in electronic equipment
precludes easy measurement of neuronal functioning
in the presence of electromagnetic fields. Chemical
measurements made by neurochemists are only just
beginning to be understood as indicators of neuronal
functioning.  Techniques are only now becoming
available by which minute chemical changes known
to occur in the  nervous system as it responds to
specific inputs can be measured.

Investigations of effects of RF  radiation on  CNS
development have usually examined behavior rather
than the  nervous  system at  the  cellular  and
subcellular levels. Thus the effects of low-level
exposure to RF  radiation on the morphology  and
physiology of the adult and developing CNS have not
been adequately studied to permit  definitive
conclusions to be drawn. Two morphological studies
have been done by Albert et al. (1981 a,b). In the first
study,  he  reported permanent  loss of  cerebellar
Purkinje cells in rat pups after exposure of pregnant
dams to 2450-MHz (CW, SAR of 2 W/kg) and both the
dams and their offspring to 100-MHz (CW, SAR of
2.77 W/kg) radiation. These authors also reported an
apparent reversibility of a decrease in  cerebellar
Purkinje cells following irradiation of rat pups 6 to 10
Table 5-12.   Summary of Studies Concerning RF-Radiation Effects on the Nervous System*

                                                  Exposure Conditions
Effects
Desynchronized EEC
Greater effect of CNS
stimulating drugs
Biphasic effect of latency
to a convulsive drug effect
Decrease effect of paralyzing
drugs
Changes in EEC patterns of
anesthetized animals
Potentiation of drug response
Decreased hypothalamic NE. DA,
and hippocampal serotonin in
hyperthermic animals
Decreased hypothalamic NE,
DA
No effect on neurotransmitter
levels
No effect on GABA content
Swollen neurons in
hypothalamus and subthalamus
Swollen neurons in hypo-
Species
Rabbit
Rabbit
Mice
Rats
Rabbits
Male rats
Rats
Rats
Rats
Rats
Chinese
hamsters
Chinese
Frequency
(MHz)
3000 (PW)
3000 (PW)
3000 (PW)
3000 (PW)
9300 (CW)
2450 (PW)
1600(CW)
1 600 (CW)
1600(CW)
2860 (PW)
2450 (CW)
1 700 (CW)
Intensity
(mW/cm2)
20 (av)
7(av)
5
5
0.7-2.8
1 .0 (av)
80
20,80
10
80
40
10
10
50
25
10
Duration
(days x min)
1 x20
24-26 x 180
8-36 x
Unknown
10-15 x Unknown
1 x5
1 x30
1 x 10
1 x 10
1 x 10
1 xS
1 x20
5x480
40x240
1 x30
22x840
1 x 30-120
SAR
(W/kg)
3.0 (est)
1 .0 (est)
5
1
0.1 -0.3 (est)
0.2 (est)
24 (est)
.6-24 (est)
3.0 (est)
16.0 (est)
8.0 (est)
2.0 (est)
15 (est)
7.5
3 (est)
Reference
Baranski and Edelwejn (1 968)
Baranski and Edelwejn (1968)
Servant le era/. (1974)
Servantie et al. (1974)
Goldstein and Sisko (1974)
Thomas era/. (1979)
Merrittera/. (1976)
Merrittef a/. (1977)
Merritt et al. (1977)
Zeman era/. (1973)
Albert and OeSantis (1975)
Albert and OeSantis (1976)
 thalamus and subthalamus
                        hamsters
                       5-50

-------
Table 6-12.   (Continued)
Exposure Conditions
Effects
Myelm figures in dendrites
6 weeks post -exposure

Increased permeability of
BBS to fluorescein
Myelm degeneration and
metabolic alterations;
glial cell proliferation
Focal areas of increased
BBB permeability to
peroxidase
Increased peroxidase in brain.
absent after recovery period
Increased peroxidase in brain

Brain temperature elevation
(40-45°C|; increased perme-
ability of BBB
Increased permeability of BBB
(mannitol and inulin)
Decreased latency of late
components of thalamic somato-
sensory evoked potentials
Attenuation of monosynaptic
spinal reflex
EEC effects seen after chronic
but not acute exposures


Change of predominant EEG
frequencies
Reversible neuronal morphology
alterations
Mild pyknosis of hippocampal
neurons, increased brain and
rectal temperature
Increased brain serotonin
turnover rate
Decreased brain serotonin
turnover rate
No decrease in cerebellar
purkinje cells in offspring
Decreased cerebellar purkinje
cells after perinatal exposure
Species
Female rats


Rats

Guinea
pigs
Rabbits
Chinese
hamsters
Rats
Chinese
hamsters
Chinese
hamsters
Rats


Rats

Cat


Cat

Rabbit



Cat

Rat
Rat
Rat


Rat

Rat

Squirrel
monkey
Rat

Frequency
(MHz)
2450 (CW;
multimodal
cavity)
1200ICW)
1200(PW)
3000 (CW.PW)
3000 (CW.PW)
3000 (CW;PW)
2800 (CW)

2800 (CW)
2450 (CW)

2450 (CW)

2450 (CW)


1300(CW)
1300(PW)
918 (CW)


2450 (CW)

1-10 (AM)

1-10 (AM)

147(AM)

3000
3000
2450


3000

3000

2450

2450
100
Intensity
(mW/cm2)
10


24
0.2 (av)
3.5
25
5
10

10
10

10

SOW


1.0
0.3 (av)
2.6


3.75 W

60-500
V^/m
90-500
V™/m
1.0

10
10
60-90 W


40

10

10

10
46
Duration
(days x min)
110x300


1 x30
1 x30
90 x 180
1 x 180
90x180
1 x 120

1 x 120
1 x 120

1 x 120

1 x 10-30


1 x20
1 x20
1 x 15


1 x3

1 x 120-180

20-30 x
120-180
Varying

35x30
35x30
1 x 2.5-7


1 x 60

7x480

368 x 180

5x 1260
110x240
SAR
(W/kg)
2.3


1 .0 lest)
0.08 (est)
0.5 (est)
3.5
0.4 (est)
1.9

0.9
2.5

2.5

—


0.4
0.1
2.5


800

10'5 -10~"

10-" -10'3

0.01 5 (est)

2 (est)
2 (est)
head only


8.0 (est)

2.0 (est)

3.4

2.0
2.7
Reference
Switzer and Mitchell (1977)


Freyefa/. (1975)

Baranski (1972b)


Albert (1979b)


Albert and Kerns (1981)

Albert (1977)

Sutton and Carroll (1979)


Oscar and Hawkins (1977)

Johnson and Guy (1972)


Taylor and Ashleman (1975)

Takashima et al. (1979)



Bawinef a/. (1973)

Gordon (1970)
Tolgskaya and Gordon (1973)
Austin and Horvath (1954)


Snvder(1971)

Snyder(1971)

Albert et al. (1981 a)

Albert el al. (1981b)

"AM = amplitude modulation, NE - norepmephrme. DA - dopamme

days postnatally at 2450 MHz. Rats killed immediately
after exposure had a 24-percent decrease in these
cells, whereas those killed  40 days  after the
exposures had only  an insignificant 7-percent
decrease. No histological examination was made of
other  portions of the  brains of these rats. In the
second study (Albert et al. 1981 b), no difference was
observed in the  number of Purkinje cells in the uvula
of the cerebellum from  seven squirrel monkeys
exposed for 3 h,  5 days/week, both prenatally and for
9 months postnatally to a 2450-MHz (CW) field in a
multimodal cavity (SAR calculated  to be 3.4  W/kg),
when  compared to cell  counts in  seven  control
monkeys.  Therefore, the  extent of abnormal
development and the conditions of exposure that lead
to such development are still uncertain.
Synergistic and antagonistic effects of RF radiation
with chemical agents that affect the nervous system
have not been investigated in a systematic fashion. In
addition, existing data are quite old, and experiments
utilizing the latest neurochemical measurements and
techniques are lacking. Information on the effects of
chronic and  low-level RF-radiation  exposure on
neurotransmitters is lacking.  Data on microwave-
produced alterations in BBB permeability are
controversial;  the effects  may be simply due to
increased brain temperature or increased blood flow.
In summary, the state of the art does not permit one to
assume that exposure to low-level RF  radiation
produces  a  significant  effect  on nervous  system

                    5-51

-------
morphology,  the blood-brain barrier, CNS-active
drugs,  or neurotransmitters. Reports of long-term,
low-level exposure on the developing and adult
nervous system are conspicuously absent from the
Western European and the U.S. literature. Higher-
level, acute exposure  may alter  nervous system
structure and function, but the effects may not be
specific to  the  nervous system and could be the
consequence of body heating.
                       5-52

-------
5.5 Behavior

Michael I. Gage
Behavior has been defined as "anything an organism
does" (Catania 1968), or as "the actions or reactions
of persons or things under specified circumstances"
(Morris 1976). Behavior may also be defined as the
actions of an organism in relation to its environmental
stimuli.
Behavioral  effects of RF  radiation have been
extensively studied  for  several reasons. The first is
the reports of experiments (not fully documented with
methods and detailed data analysis) from the Soviet
Union  and  other  East European countries that
behavioral  effects of  microwaves are seen at
relatively low levels, i.e., 500 //W/cm2 and below at
2375t MHz in rats (Dumansky and Shandala  1974;
Shandala et at. 1977).  The second is that behavior
can serve as an index of how the whole organism is
functioning,  displaying  the  status of  the nervous
system and many other  organ systems of the body as
they act together. Behavior has often been defined as
the final common pathway of the nervous system.
Moreover, behavior can  be analyzed in a nonterminal
fashion, without resort to surgically or biochemically
invasive preparatory techniques.

Behavior may be separated into two major categories:
naturalistic and acquired. Spontaneous or naturally
occurring behavior may be innate and is  often
species-specific in frequency of occurrence. Examples
of naturally occurring  behavior include locomotor
activity, eating, drinking, and mating.  Elements of
these behaviors are often acquired.
There  are two general categories of acquired or
learned behavior—respondent  and operant, which
are distinguished by the procedures  used  in  the
acquisition or conditioning of the behavior. Respon-
dent conditioning occurs as  a  consequence of the
temporal contiguity between  stimuli. Stimuli paired
in time with  another  stimulus,  which  reflexively
elicits  a response,  gradually elicit the  response.
Examples of acquired respondent behavior  are
salivation and hunger pangs when one is passing a
restaurant  or  eye blinks  to acoustic stimuli.
Responses conditioned by respondent procedures are
usually measured by their  occurrence and their
magnitude.
Much complex behavior of human beings and other
higher animals in the course of daily activities can be
viewed  as  emitted  or  operant behavior. Operant
conditioning  occurs  as a consequence of reinforce-
ment  that follows  the emission of a response.
Reinforcers may be positive (such as food or water), or
negative (such as termination  of electric shock or
intense radiant energy). All reinforcers maintain or
increase the frequency  of response. By definition, a
positive reinforcer  is a  stimulus that increases or
maintains the probability of emission of an operant by
 its presentation. A negative reinforcer is a stimulus
 that increases  or  maintains this probability by its
 removal after emission of the operant. Reinforcement
 need not follow the occurrence of each response. It
 may be intermittent according  to a schedule (i.e., a
 schedule of reinforcement). Responses are said to be
 conditioned when they have  a  highly predictable
 probability of occurrence. This probability is  often
 expressed by the average occurrence within a period
 of time, or the response rate.  Examples of operant
 behavior include eating with a knife and fork, driving a
 car,  or writing  a  review on  behavioral effects  of
 microwaves.

Operant conditioning may be used to answer specific
questions about behavior that are known to be similar
across species. For example, by  reinforcing the
responses in the presence of light of one wavelength
but not reinforcing responses in the presence of light
of any other wavelength, animals  can learn to
respond selectively in the presence of light of the first
wavelength. By presenting light at wavelengths close
to the one that is reinforced,  one can obtain the
threshold for  discriminability of colors.

Specific types of behaviors investigated in behavioral
research  are  so numerous that no attempt can be
made to describe them all here. Descriptions will be
given of  behaviors that  have been  studied as  a
function of  microwave exposures.  For a  more
complete  description of behavior as studied by
ethologists and  psychologists, the reader is referred
to several standard books (Brown 1975; Hinde 1970;
Honig and Staddon 1977; Kling and  Riggs 1971;
Konorski 1967;  Pavlov 1960; and Skinner 1953).

Some general statements can be made regarding the
effects of RF-radiation exposure on behavior:

 • Some microwave effects have been reported for a
   variety of  animal behaviors. Most of the studies
   have used rats as subjects; only a few have used
   mice, squirrel monkeys, and rhesus monkeys.
 • Changes in  locomotor behavior have occurred
   after CW exposures at an SAR as low as 1.2 W/kg
   in rats (D'Andrea et a/. 1979). Changes in food and
   water  intake or body mass have  not been
   consistently  reported at such levels.
 • Decreases in rates of  ongoing operant behavior
   have been seen during exposures at SAR = 2.5
   W/kg  in  rats (de Lorge and Ezell  1980), and
   cessation of operant behavior has been seen at an
   SAR of 10 W/kg in rats (D'Andrea  et al. 1976).

 • Alterations in operant performance measured
   after exposure  was terminated also  occurred
   with SARs of 2.5 W/kg or more  in rats (Gage
   1979a).

 • The threshold for detection of microwaves may be
   as low as 0.6 W/kg in rats (King et al. 1971).
                                                                     5-53

-------
   However, it is not certain that animals avoid or
   attempt to escape from CW microwaves, even at
   very high power levels.
 • Drug  effects on behavior  in rats have  been
   augmented after PW-radiation exposures lasting
   0.5 h  at average SAR = 0.2 W/kg (Thomas et al.
   1979). Behavioral thermoregulation  has been
   altered after only several minutes of exposure at
   SAR = 1.0 W/kg in the rat (Stern era/. 1979)andat
   SAR = 1.0 W/kg in the squirrel monkey (Adair and
   Adams 1980b).
 • Although the same behavioral effects during or
   after microwave exposure of the same magnitude
   may not be consistently predictable, enough
   behavioral changes have  been reported after
   similar exposures to warrant the conclusion that
   behavior  is  disrupted by microwaves with  an
   energy input that approximates one quarter to one
   half of the resting metabolic rate of many animals.
 • When the persistence of  behavioral changes
   after  the  termination  of  exposure has  been
   investigated, the behavioral alterations reported
   were  reversible  with  time after the  exposure
   ended.

5.5.7 Naturalistic Behavior
Spontaneous  locomotor behavior has been studied
with both acute- and chronic-exposure  regimens in
the rat. Hunt et al.  (1975) found decreased initial
exploratory activity by male Wistar rats after a 30-min
exposure to 2450-MHz fields in a multimodal cavity
(Ta = 24°C, RH from 20 to 40 percent), with power
adjusted to produce an SAR of 6.3 W/kg. The activity
of exposed animals returned to control values within
a 2-h  period. The decrease in initial exploratory
activity was the same whether the rat was placed in
the apparatus immediately after or 1 h after exposure
was  terminated. Core body temperature  was in-
creased to 40.3°C at termination of exposure but
dropped to 37.8°C, within the normal range, 1 h after
exposure. Decreased swimming speed in water at
24°C was also seen in rats  that  were  practiced
swimmers after  30-min exposures to 2450-MHz
fields at SARs of either 6.3 or 11 W/kg. The effects
were seen only after 1.2 km of swimming following a
6.3-W/kg exposure. After an 11-W/kg exposure,
effects were  seen immediately for the  first few
meters and again after 0.6 km of swimming despite
the fact that water at 24°C would reduce persistence
of a microwave-related hyperthermia.
Robert! et al. (1975), on the other hand, did not see
changes in spontaneous motor activity, as measured
in a glass cage, in male Wistar rats after four different
exposure conditions in the far field  of an anechoic
chamber  (Ta = 22 ± 1°C, RH = 50 ±  5 percent). The
locomoter activity measured by Robert! et al. may
not  have required as much  physical  effort as
continuous swimming. The four exposure conditions
were (1)  10,700-MHz  (CW)  fields  at  0.6  to  0.9
mW/cm2 (SAR, based on single animal exposures,
can be estimated at 0.15 W/kg) for 185 continuous
hours (24 h daily for 7 2/3 days); (2) 3000-MHz (CW)
fields at 0.5 to 1.0 mW/cm2 (SAR can be estimated at
0.3 W/kg) for 185  continuous hours; (3) 3000-MHz
(PW) fields, 769 pulses/s, at 1.5 to 2.0 mW/cm2 (SAR
estimated at 0.6 W/kg) for 185 continuous hours; or
(4) 3000-MHz (PW) fields, 769 pulses/s at 24 to 26
mW/cm2  (SAR estimated at  8.3 W/kg) for 408
continuous hours (24 h daily for 17 days). After this
last exposure condition, no change was also observed
in running speed during forced runway running by
the rats.

Locomotor activity on a small platform was increased
(as compared with five controls) in five female
Sprague-Dawley rats during the course of repeated
exposures to 2450-MHz (CW) fields in a  multimodal
cavity  (Mitchell  et al.  1977). Exposures of 5-h
duration occurred 5 days weekly for 22 weeks (TB = 24
± 1.5°C). The average SAR was determined to be 2.3
W/kg, which is similar to the SAR at power density of
10  mW/cm2 in  a plane-wave environment. The
activity increased within the first week of exposure
and  remained high throughout the  course of  the
exposure period.

Decreases in activity, measured by visual observation,
were reported during repeated exposures of eight
Wistar male rats to 918-MHz (CW) fields in a circular
waveguide (Moe eta/. 1976). Each exposure lasted 10
h (from 2200 to 0800) and occurred nightly during the
dark and usually more active period of the circadian
cycle of a rat for 3 weeks for a total of 210 h (T,= 21.8
± 0.13°C). The range of whole-body average SARs
was measured as  3.6 to 4.2  W/kg (10 mW/cm2
average  power density). Most of  the  decreased
activity occurred shortly after  the microwave field
was activated. The exposed rats were stretched out in
a prone position more frequently than control rats in
the early morning  hours. In addition, exposed  rats
were reported to consume less food than did the
controls over the course of the exposure, even though
their body mass  was  not different from that of
controls.  A repeat  of this experiment, where eight
male Wistar  rats were exposed 10  h/night for 13
weeks at an average power density of 2.5 mW/cm2
(average SAR = 0.9 to 1.0 W/kg; T. = 21.1 ±0.15°C),
resulted in no differences in food consumption or in
activity measured during the eleventh week (Lovely et
al. 1977). The two  studies indicate that a dose-rate-
related threshold for these effects might be somewhere
between  1.0 and 3.6 W/kg.  Unfortunately, in the
circular waveguide, the power density is twice the
average on axis and falls off rapidly toward the wall,
which may  lead  to fluctuating and  uncertain
quantitative description  of SAR  at  specific  times
during exposure (Guy and Chou 1976).
                       5-54

-------
Effects on  spontaneous  behavior of rats were
reported in two other chronic-exposure experiments.
Both experiments used 15 exposed and 15 control
male Long-Evans rats irradiated during the illuminated
portion of the circadian cycle from 0900 to 1700(8 h),
5 days/week for 16 weeks in an anechoic chamber
with a central monopole antenna and ground plane
(Ta = 22 ± 2°C, RH from 20 to 40 percent). The rats
were adapted to the exposure and testing apparatus
for 4 or 8 weeks before irradiation. In one experiment
(D'Andrea et al. 1979), exposures were to 2450-MHz
(CW) fields at a power density of 5 mW/cm2 (average
SAR = 1.2 W/kg); and in the other (D'Andrea et al.
1980), exposures were to 915-MHz (CW) fields at a
power density of 5 mW/cm2  (average SAR = 2.5
W/kg).
In the  study at 2450 MHz,  rats showed decreases in
activity as measured on  a  stabilimetric platform
throughout  the course of exposure but increased
running-wheel activity overnight through the course
of exposure. (This latter effect was not significant.) No
significant differences were seen in food and water
intake and in body mass.  In the study at 915 MHz,
exposed rats showed increased activity as measured
both in the running wheels and on the stabilimeter.
Again, no  significant  changes in  food and water
intake or body mass were observed.
Rudnev et al. (1978) reported effects of exposure of
25 male albino rats to 2375-MHz (CW) fields at 0.5
mW/cm2 (SAR estimated at 0.1 W/kg for individual
animal exposure) for 7 h daily for 1 month. Open field
activity was measured in  irradiated rats and in 25
controls as the number of squares crossed in 3 min on
2 successive days. The count on the first day was
defined as exploratory activity, and on the following
day, as motor activity. Shock-induced aggression was
measured by  observation of battles between an
exposed and control rat. Maintenance of balance on a
rotating treadmill (dynamic-load endurance) and on
an inclined rod (static-load endurance) was measured,
as was the amount of food  consumed in 20 min after
23 h  of deprivation. Electrodermal skin sensitivity
was measured as the voltage of 100-Hz square-wave
electrical stimulus that was needed  to elicit  paw
withdrawal from the metal bars on the cage floor. The
above  measurements were made at the start of the
series  of exposures and again prior  to irradiation
periods on the 10th, 20th, and 30th days of exposure,
and every  15 days for 3 months after exposure.

A significant reduction in  food intake, time  on the
treadmill, and in time on the inclined rod was seen by
the 10th day of exposure.  Exploratory activity  was
significantly decreased, and  shock  sensitivity  was
increased after 20 days of exposure. At the end of
exposure, exploratory and motor activity, time on the
inclined bar, and shock sensitivity were significantly
decreased. The latency to start eating when food was
presented was increased for 30 days after exposure
ended. Time on the treadmill was reduced for 15 days
after the termination exposure. Exploratory activity
was increased throughout the 3-month postexposure
period.  Sensitivity to electric  shock was  reduced
significantly on the 30th and 60th day after exposure
ended and was still below control levels on the 90th
day.
5.5.2 Learned Behavior
5.5.2.1 Respondent Conditioning

Currently, there appear to be no  reports  by U.S.
investigators of microwave exposure altering
respondent behavior, although  the description of an
experiment that used RF radiation as an unconditional
stimulus has been published.  Some studies using
respondent conditioning techniques  that report
changes in behavior as a consequence of microwave
exposure appear in the Soviet literature (e.g.,
Dumansky and  Shandala 1974; Lobanova 1974).
Unfortunately, details regarding exposure conditions
or  behavioral methodology and results  are too
sketchy to permit inclusion in this review. An attempt
was made to use microwave exposure as  an
unconditional stimulus in one experiment (Bermant
et al. 1979). A 30-s presentation of a 525-Hz tone
that preceded 2450-MHz sinusoidally modulated
microwaves that were either presented for 10 s (SAR
= 420 W/kg)  or 30  s (SAR = 220  W/kg), or that
preceded an electric shock to the tail, produced  rises
in  rectal temperature  of 0.7, 0.5,  or 0.37°C,
respectively, during a base-line period over the
course of 10 conditioning sessions at an unspecified
Ta.  Control female Sprague-Dawley rats presented
with the tone alone showed a decrease of 0.47°C in
rectal temperature during this period. The microwave
exposure itself was designed to produce an increase
in rectal temperature of 1.5°C. Aside from this study
there  are  no reports of microwaves  altering
respondently conditioned behavior in which exposure
parameters allow clear determination of the SAR.

5.5.2.2 Operant Conditioning

There is a large body of literature  that examines
alterations of operant behavior produced by micro-
waves for which SAR values have been, or can be,
determined.
Reduction in response rates on  an operantly
conditioned task has been observed during the course
of microwave irradiation. Rats were trained to lever-
press for food pellets on a  random-interval, 30-s
schedule. After behavior on this task stabilized, the
rats showed a response  rate that was in  general
linearly  uniform  and typical of random- or variable-
interval  schedule-controlled performance. The rats
were then  exposed to  microwaves  under each of
several conditions in sessions lasting 25 min, or until
the rat's response rate fell below one third of its base-
line control rate. In an initial experiment (D'Andrea er
                                                                     5-55

-------
a/. 1976), six male Long-Evans rats were exposed to
360-, 480-, or 500-MHz (CW) fields (Ta from 21.1 to
22.2°C, RH from 12 to 40 percent) in a parallel-plate
apparatus  at an incident  power  density of 25
mW/cm2 with the long axis of the rat parallel either to
the electric-field vector or  to the vector of wave
propagation. Responding was reduced  only during
exposures to 500-MHz fields when the .long axis of
the body was parallel to the electric-field vector.
Behavior stopped abruptly after ~11 min of exposure,
and upon  removal from the apparatus the animals
appeared flaccid, wet, and, according to the authors,
heat stressed.  At this  exposure,  the SAR was
comptued  from  the  measured 0.16°C/min rise in
rectal temperature to be ~ 10 W/kg. Exposures that
produced no change in behavior had SARs ranging
from 5  to 6 W/kg.
In a second experiment (D'Andrea etal. 1977), these
results were  confirmed  and extended. Exposures
were conducted in a monopole-above-ground
radiation chamber and lasted for up to 55 min, or until
responding on  the  random-interval schedule fell
below one third of the base-line rate. Exposures of
five male Long-Evans rats for periods up to 55 min to
400-, 500-, 600-, and 700-MHz (CW) fields at 20
mW/cm2 power density (Ta = 22± 1 °C; RH = 50 ±1.5
perent) yielded a U-shaped  function of time to the
criterion reduction in rate, with the minimum of ~ 23
min at  600 MHz when rectal temperature increased
0.09°C/min (SAR estimated at 16.4 W/kg). At the
time they stopped responding, all rats appeared heat
stressed and were engaged in spreading saliva on
their fur. Six additional rats, exposed to 600-MHz
(CW) fields at power densities of 5, 7.5, 10, and 20
mW/cm2,  showed decreased times to stop  their
responding at power densities above 7.5 mW/cm2
when rectal temperature increased- 0.04°C/min or
more. At 10 mW/cm2 (SAR  estimated at 7.5 W/kg)
the rats stopped responding after—45 min, and rectal
temperature  increased  0.04°C/min.  Three rats
exposed to PW microwaves with 170 mW/cm2 peak
and 5.1 mW/cm2 average power density showed no
change in performance. As in the earlier experiment
of D'Andrea et at. (1976), response reduction was
abrupt and was correlated with the rectal temperature
increase.
A series of experiments by 'te Lorge (1976,1979) also
examined alterations in operant performance during
exposure to  microwaves. In most 'of these experi-
ments, the schedule of reinforcement was used to
measure observing and detection responses in an
operant task. Two levers were present in a testing
chamber or in front of a primate restraint chair made
of Styrofoam and sheet plastic  for optimal trans-
parency to microwaves. Responses on the right lever,
called observing responses, produced either a low- or
a high-frequency acoustic signal. The high-frequency
signal was scheduled to occur on a variable interval of
either 30 or 60 s. If the animal made a response on the
left lever when the high-frequency signal occurred—
a detection response, it received a food pellet as a
reinforcer. Observing responses occurred at fairly
linear response rates, a pattern similar to that seen on
variable interval schedules of reinforcement.
In the first experiment (de  Lorge  1976) five male
rhesus monkeys (Macaca mulatta) showed no change
in behavior  on this schedule during  1  or 2 h  of
exposure to 2450 MHz in the far field of an anechoic
chamber (Ta from 21 to 24°C, RH = 70 ± 15 percent)
when power densities ranged to 16 mW/cm2 (SAR
estimated  at 1.2 W/kg). The field was 100-percent
amplitude modulated  at  120 Hz. Three of  these
monkeys  showed  reduced rates of  observing
response during 1-h exposure to this field at 72
mW/cm2 (SAR estimated at 5.0 W/kg) but not at 32,
42, 52, or (in two  monkeys) 62 mW/cm2. At 72
mW/cm2,  rectal temperatures rose ~ 2°C and were
still increasing at the end of the hour-long session.
The animals moved more in their chairs after 20 min
and were observed to take short naps after — 30 min
at this power density. When observing responses
decreased, reaction time to respond on the left lever
increased.

In the second experiment (de Lorge 1979) four male
squirrel monkeys (Saimeri sciureus) were tested and
exposed to 2450-MHz 120-Hz modulated microwaves
under conditions similar to those described above at
10, 20, 30, 40, 50, 60, and 70 mW/cm2 (T. = 22.5 to
23.2°C, RH = 57 percent during 0.5-h exposures and
74 percent during 1 -h exposures). No reduction in the
mean rate of right lever observing responses > 1
standard deviation belqwthe mean of the control rate
was seen at any power density  during the  30-min
exposures. But brief pauses in responding were seen
at microwave onset and offset at 50 mW/cm2 and
above (SAR estimated at 2.75 W/kg). At this power
density rectal temperature rose > 1°C. Three of the
monkeys were also given 1-h exposures. Graphic
representation of data  from one monkey showed
observing  response  rate decreased more, and this
decrease began earlier,  in sessions at the higher
power densities (de Lorge 1979, Figure 6). One of the
three monkeys exhibited an increase of response rate
as power density  increased.  After microwave
exposure, all  monkeys also showed  decreased
responding that was directly related to  the  power
density during the exposure. These decreases were
apparent for a 30-min period following exposure but
gradually returned to base-line values. All monkeys
showed increases in frequency of incorrect respond-
ing on the left lever for food, which  was a direct
function of power density. Changes in behavior
consistently occurred only at power densities of 40 to
50 mW/cm2 at a time when rectal temperatures had
risen by > 1 °C.
                       5-56

-------
In a third experiment (de Lorge and Ezell 1980), eight
male Long-Evans rats were exposed in an anechoic
chamber (Ta varied from 23 to 26.5°C, RH from about
50 to 52 percent) to 5620-MHz (PW) fields at 662
pulses/s, and then to 1280-MHz (PW) fields at 370
pulses/s. They were tested on the behavioral task of
vigilance during exposures (SARs were reported at
0.19 W/kg per mW/cm2 at 5620 MHz, and 0.25
W/kg per mW/cm2 at 1280 MHz). Rates of observing
responses  during  exposures to 1280-MHz fields
decreased somewhat at 10 mW/cm2 (average power
density) and decreased markedly after 15 to 20 min
during 15 mW/cm2. At 5620 MHz, rates of observing
responses decreased only during exposures at  26
mW/cm2 and above. Behavior decrements occurred
at SAR = 2.5 W/kg at 1280 MHz but required SAR =
4.9 W/kg at 5620 MHz.
In a related experiment (Sanza and de Lorge 1977),
four male  Sprague-Dawley rats were trained to
respond  on a fixed-interval  50-s schedule for food
pellets as reinforcers. With  this schedule, the first
response emitted 50 or more seconds after arrival of
the last  food pellet produced another food pellet.
Exposures for 60 min to 2450-MHz, 120-Hz
modulated fields at 37.5 mW/cm2 in the far field of an
anechoic chamber  (Ta = 24  ± 0.6°C, RH = 70 ± 5
percent) produced decreases in  response rates that
had a fairly abrupt  onset. The response decrements
were seen only in two rats that had high base-line
rates and not in two rats with low response rates.
Exposures at 8.8 and 18.4  mW/cm2  produced  no
decrements in performance. All rats spent more time
at the wall opposite the food cup during 18.4- and
37.5-mW/cm2 exposures than during sham or 8.8-
mW/cm2 exposures. The SAR was not given but is
estimated at 3.7 W/kg at 18.4 mW/cm2 and 7.5
W/kg at 37.5 mW/cm2.
Responding of three male rhesus monkeys trained to
a high level of proficiency on a  visual tracking task
was not disrupted by exposures at 10 or 20 mW/cm2
to 1200-MHz (CW)  fields (reported SAR estimated at
0.8 and 1.6  W/kg; Ta = 27 ± 3°C, RH = 50 ± 3 percent)
during behavior sessions  lasting ~ 2 h (Scholl and
Allen  1979).

Some  other studies have  looked at changes  in
previously learned operant behavior at the termination
of single or  multiple exposure periods. Thomas et a/.
(1975) trained four male Sprague-Dawley rats  to
respond on a multiple schedule. One component was
fixed ratio 20 (FR20): Every 20th response on the right
lever was reinforced by a  food pellet. The other
component  was a differential-reinforcement-of-low-
rate of 18 s with a limited hold of 6 s (DRL 18 LH  6):
Responses on the left lever separated by at least 18 s,
but by  no  more  than  24 s,  were  reinforced.
Components alternated on an irregular basis. Only
one of the schedules was in effect during a given time
period. Exposures to microwaves lasted for 30 min.
and  behavioral testing began 5 to 10  min after
exposure. Exposure parameters were as follows: 5,7,
15, and 20 mW/cm2 to 2450-MHz (CW) fields; 5, 10,
15, and 20 mW/cm2 to 2860-MHz (PW) fields at 500
pulses/s and 1 -fjs pulse width; and 1, 5, 10, and 15
mW/cm2 to 9600-MHz (PW) fields at 500 pulses/s
and  1-yt/s pulse width. All rats were exposed to all
parameters while restrained in the far field of an
anechoic chamber. In general, response rates
increased on the DRL schedule and decreased on the
FR schedule. Increased rates on the DRL schedule
were seen following exposures to 9600-MHz fields at
5 mW/cm2 and above (SAR  estimated at 1.5 W/kg),
to 2450-MHz (CW) fields at 7.5 mW/cm2 and above
(SAR estimated at 2.0 W/kg), and to 2860-MHzf ields
at 10  mW/cm2 and above  (SAR estimated  at 2.7
W/kg). Decreased response rates on the FR schedule
were observed following exposures to all frequencies
at 5 mW/cm2 and above (SAR estimated at 1.5 W/kg
for 9600 MHz, and 1.4 WAg for 2450 and 2860
MHz).  Increased response  rates during time-out
periods between components  were seen following
exposures to all three  frequencies at 5 mW/cm2.
Time-out responses peaked  and then dropped after
exposures at higher power densities.

In a second study, Thomas et al. (1976) trained four
male Sprague-Dawley food-deprived rats on a fixed-
consecutive-number-eight (FCN 8) schedule. With
this schedule, at least eight  presses had to be made
on the right  lever before depression of the left lever
would yield a food pellet reinforcer. If the rat made
fewer than eight consecutive responses on the right
lever before switching, the  count was  restarted at
zero. Well-trained  rats were  tested after 30-min
exposures in the near field at 5,10, and 15 mW/cm2
to 2450-MHz (PW) fields at 500 pulses/s, and 1 -//s
pulse width. Because exposures were  in the near
field, SAR cannot be precisely estimated but may be
assumed at 0.4 W/kg for each 1 mW/cm2 (Durney et
al. 1980). The percentage of eight or more consecutive
responses  on the  right  lever  (reinforced runs)
decreased, and the length of these runs decreased
after all  exposures. Those  decreases were direct
functions of power density. However, the overall rate
of responding and the running rate (i.e., the rate after
the first  response following reinforcement)  hardly
decreased during these  postexposure sessions.
Gage (1979a) reported decreases  in response rates
of eight adult male Sprague-Dawley rats trained to
alternate between levers either 11 or 33 times for a
food-pellet reinforcer. The decrements occurred in
sessions following overnight, 15-h exposures (Ta =
22°C, RH = 50 percent) to 2450-MHz(CW)f ields at 10,
15, and 20 mW/cm2, but not after exposures at 0.5
and 1.0 mW/cm2. Only very small decrements were
seen after 55-min exposures at power densities up to
30 mW/cm2. (The SAR measured in rats  under
similar exposure conditions was0.3 W/kg for each 1
mW/cm2.)
                                                                    5-57

-------
Discrimination was tested immediately after 30 min
of irradiation in 10 well-trained, young adult male
Wistar rats (Hunt et al. 1975). Exposures produced
SARs of 0.0,  6.5, and  11.0 W/kg at  2450 MHz
(modulated in a quasi-sinusoidal fashion at 120 Hz in
a multimodal cavity at Ta = 24°C, RH from 20 to 40
percent). All rats were exposed at both SAR values,
but the sequence of exposures was varied. The task
required the rats to obtain saccharin-flavored water
reinforcers by pressing a bar when a light flashed, and
by not pressing a bar when a sonic stimulus was
presented.  One of the two stimuli was presented
every 5 s, and the light was presented 12.5 percent of
the time. Rats exposed at both SARs had an increased
number of omission errors (i.e., failures to respond to
the light after exposure), but there were no increases
in errors of comission (i.e., responding wrongly when
the noise was presented). The failure  to respond
correctly was more  frequent at the start of the
session, as well as after exposure at the higher SAR.
Schrot et al. (1980) experimented  with  three male
albino  rats trained to  learn a  new sequence  of
pressing three levers for food reinforcers daily in a
repeated-acquisition procedure. The rats increased
their number of errors and decreased their rate of
sequence completions when tested immediately after
30-min  exposures to 2800-MHz (PW)  fields (500
pulses/s, 2-fis pulse width) at 5 and 10 mW/cm2
average power density (Ta = 21 ± 1.5°C). No effects
were seen at lower average power densities of 0.25,
0.5, and 1 mW/cm2. Peak powers were 0.25,0.5,1,5,
and 10  W/cm2 in these exposures. The rats were
exposed in a sleeve holder with the electric-field
vector perpendicular to the  long axis of the animal's
body. The  reported SARs based  on temperature
measurements were 0.7 and 1.7 WAg at 5 and 10
mW/cm2, respectively.
Operant behavior has also been examined during or
following the course of chronic microwave exposures
in some reports also described  in Sec.  5.5.3,
Naturalistic Behavior. Mitchell era/. (1977)measured
performance on two schedules in separate groups of
rats pretrained before the start of 22  weeks of 5-h
exposures, 5 days/week, to 2450-MHz fields (aver-
age SAR at 2.3 W/kg) in a multimodal cavity (T. = 24 ±
1.5°C, RH = 50 ± 10 percent). Five  exposed and five
control female Sprague-Dawley rats were tested for
30-min  sessions  on a schedule of  multiple  FR5
extinction for 15 s (MULT  FR5 EXT 15  s). When a
white  lamp  was  on,  every  fifth response was
reinforced by a food pellet,  but during periods when
the lamp was off, no responses were  reinforced.
Although the control rats had higher response rates
during the FR component, this difference was not
significant. Exposed rats had higher response rates
than controls during the extinction component, and
this difference was statistically significant. The ratio
of response  rate during the FR5 component  to
response rate during the  EXT15  component was
significantly different in exposed as compared with
control  rats:  Exposed rats exhibited a higher ratio,
indicating poorer discrimination over the course of
exposure, although they had a value like that of the
control rats before irradiation  began. Another
schedule, Sidman avoidance, was used to test escape
and avoidance response. Five exposed and four
control rats were trained to postpone an unsignaled
2.0-mA foot shock for 15 s (response-shock interval)
by  pressing a lever. During 30-min sessions, the
animals received a shock  every 0.5 s (shock-shock
interval) until they pressed the lever. No significant
effects of microwave exposures were seen with this
schedule, although improvement in avoidance was
seen in all rats both within and over the course of
testing sessions.
Conditioned taste aversion was studied in experiments
where rats were exposed chronically to 918-MHz fields in
circular waveguides at 10 mW/cm2 (Moe et al. 1976)
and 2.5 mW/cm2 (Lovely era/. 1977). Rats were given
a saccharin solution to drink in place of water during
the microwave exposure period. Presumably, if the
saccharin were drunk in conjunction with an agent,
such as microwave radiation, which made the rat sick
or produced a yet unspecifiable effect, a connection
would be learned between the drinking of the agent
and the consequences, and the solution  would  be
avoided  in the future. The investigators measured
preference after  exposure by allowing  a water-
deprived  rat to  choose  between  drinking the
saccharin solution and water for  20 min. In the
experiment at 2.5 mW/cm2 (Lovely et al.  1977),
saccharin preference was tested only from the 9th to
the 13th week of exposure. No difference between
exposed and control  rats  was seen  in amount of
saccharin solution consumed either at 2.5 mW/cm2
(average measured  SAR at 1.0 W/kg)  or  at  10
mW/cm2 (average measured SAR at 3.9 W/kg).
Several studies have  investigated the  ability of
animals to detect or to take behavioral  action to
minimize, avoid, or escape from microwaves. In  an
early paper, King et al. (1971) showed that three
irradiated and three control male  albino rats could
respond to 2450-MHz microwaves doubly modulated
at 60 and 12 Hz as the conditioned"stimulus in a
measure of conditional  suppression. The microwave
radiation was presented in a multimodal exposure
cavity (T, = 24 ± 2°C, RH from 20 to 40 percent). In this
experiment, rats were reinforced with sugar water on
a random interval schedule for licking at a water tube.
At various times during each 2-h session either a
525-Hz tone or the microwaves were presented for 1
min, and an unavoidable 0.5-s electrical foot-shock
followed. Conditioned suppression to the tone was
reliably indicated by no licks being emitted during the
tone. Microwave dose rates of 0.6,1.2, 2.4,4.8, and
6.4 W/kg were substituted for the tone in some
sessions. During irradiation at SAR = 0.6 W/kg one of
three rats suppressed responding, at SAR = 1.2 W/kg
                       5-58

-------
two rats suppressed, and at SAR = 2.4 W/kg all three
suppressed reliably.
Johnson et al. (1976) trained  two male Wistar-
derived rats to nose poke in a restraint for food pellets
on an FR5 schedule  in the presence of an acoustic-
pulse stimulus of 7.5 kHz, 10 pulses/s, 3-//s pulse
duration for 3-min periods, which alternated with 3-
min  periods of no stimulation during  which nose
pokes were  not reinforced (extinction). When
microwaves at 918 MHz (10 pulses/s,  10-//S pulse
durations) were presented for 30-s intervals during
an extinction  period,  or  when microwaves were
substituted  for the  auditory stimulus during
reinforced periods, response  rates  were observed
that were similar to those seen in periods when the
acoustic stimulus was present. The energy density
per pulse of microwaves was  150 fjj/cm2, and the
average power density was  15 mW/cm2. (The SAR
would be near 7.5 W/kg  if the rats were in the far
field.)
Detection of microwaves does not imply that there are
affective properties of this stimulus, i.e., that they
hurt  or feel good. In fact, such detection may be
further evidence of the RF hearing phenomenon(Frey
and Messenger 1973). Indeed, in several experiments
in which PW microwaves are presented during
exposure, the alteration in behavior of the exposed
animal might be due to effects of acoustic stimulation
by the microwave pulses. Such experiments should
have as a control the presentation of pulsed auditory
stimuli.

Frey et al. (1975) experimented with female Sprague-
Dawley rats exposed to 1200-MHz (PW) fields at 1000
pulses/s (0.5-s pulse duration) at an average power
density of 0.2 mW/cm2 (SAR estimated at 0.2 W/kg)
and a peak power density of 2.1 mW/cm2. Six rats
spent only 30 percent of a 30-min period of exposure
in an unshielded half of a Styrofoam shuttle box
during the last 2 of 4 successive daily exposures. Six
other rats exposed to 1200-MHz (CW) fields at 2.4
mW/cm2  (SAR estimated at  2.2 W/kg) spent 52
percent of the time in the unshielded half of the box
on these days. Six other control rats spent 64 percent
of the 30-min period in the unshielded half. Only the
rats  exposed  to  PW microwaves could be  said to
escape  from  the stimulus or  exhibit a modest
preference that would decrease their exposure. A
similar finding was also reported by Frey and  Feld
(1975) to occur in male rats during exposures  to a
1200-MHz (PW) field at 100 pulses/s at 0.4 or 0.9
mW/cm2 average power densities (SAR estimated at
0.4 and 0.81 W/kg, respectively) and at 133 or  300
mW/cm2 peak power densities (Ta=22°C). These rats
spent an average of only 29 percent of their time in the
unshielded half during seven 90-min sessions, and
this side preference was maintained throughout  all 7
days. Sham-irradiated  rats spent an average of 57
percent of their time in the unshielded half of the box.
Two groups of eight male Wistar rats spent more than
half of each of nine weekly hour periods in the side of
a shuttlebox when occupancy of that side kept a PW
microwave field turned off (Hjeresen et al. 1979). The
2880-MHz field was pulsed at 100 pulses/s (3.0-//S
pulse width) with a 9.5-mW/cm2 average and 33-
mW/cm2 peak power density (SAR calculated at 2.1
W/kg). The exposure was in  the  far field of an
anechoic chamber (Ta from 22 to 24°C, RH from 20 to
50  percent). A group of eight rats that could not
extinguish the field and another group that received
no  microwaves  showed no side preferences.
Preference for occupancy of the side that extinguished
the field increased across each weekly session as
judged from the data in Table 1 of Hjeresen et al.
(1979). Statistically significant preferences for
the unexposed side occurred during the hour periods
of weeks  2, 4, 6, and 8,  during which times the
original positions of the exposed and unexposed sides
of the shuttlebox  were reversed. Substitution of a
37.5-MHz (PW) acoustic stimulus for the microwaves
in one session resulted in the rats spending most of
the session in the  side  that kept the acoustic
stimulus off. A continuously occurring broadband
"pink" noise in the anechoic chamber prevented
appearance of side preference in two other groups. In
addition to confirming that rats &void or escape from
pulsed microwaves, this experiment suggests that
pulsed microwaves may be detected as an auditory
stimulus.
Monahan and Ho (1976) showed that male CF1 mice
irradiated for 10 or 15 min at 2450-MHz(CW)fields in
a waveguide (TB = 24 ± 0.5°C, RH = 50 ± 1.5 percent)
when forward-power levels were stabilized at 0.4,
0.8,1.6, 2.4, 3.2, 4.0, or 4.8 W exhibited a decline in
energy absorption  rate at  2.4 W and above.  This
decline usually occurred within the first 5 min of
exposure and stabilized toward the latter part of the
period. Monahan and Ho interpreted the results to
show that the rats changed their behavior to minimize
exposure to the microwaves. However, they did not
directly measure any animal  behavior.  During the
first  5 min of exposure the SARs measured in this
apparatus ranged from 7.7 to 65.7 W/kg. The lowest
SAR at which the mice clearly altered their rates of
energy absorption was 28 W/kg. In a second study
Monahan and Ho (1977) showed that the SAR
associated with reduction of absorption during 20min
of irradiation (RH = 50  ± 1.5 percent) decreased
reliably from 43.6 W/kg when ambient temperature
was  20°C, to 0.6 W/kg when ambient temperature
was  35°C. The reduction in absorption was greater
with higher forward-power  levels at any given
temperature. Forward powers used were 1.62 to 3.84
at20°C, 1.12 to 3.11 at 24°C, 0.42 to 2.45 at 30°C.
and 0.004 to 0.40 at 35°C. The mice did not altertheir
energy absorption at SARs of 28 W/kg or below until
temperature was 30°C or higher.
                                                                     5-59

-------
Videotape observations of rats and mice exposed to
2450-MHz (CW) fields at 15 mW/cm2 for 1 h in the far
field below a radiating horn in an anechoic chamber
did not  reveal  any preferential behavior that
minimized whole-body absorption rates through
parallel orientation to the magnetic- as opposed to the
electric-field vector (Gage et al. 1979). These
observations occurred at ambient temperatures of 22
or 28°C (RH = 50 percent) while the animal was held
in a cuboid or cylindrical enclosure. Usually, the
animals assumed a curled, sleep-like posture in the
hour before  the  microwaves were turned  on  and
maintained that posture throughout most of the expo-
sure period.  However.mice  exposed at 28°C more
often assumed positions that were oriented parallel to
either of the two field vectors than positions oriented
in other directions. The SAR of the rats without any
enclosure was 3.3 W/kg at 15 mW/cm2, independent
of their orientation. The  SAR of the mice when
parallel to the  electric-field  vector was 12.3 W/kg,
and the SAR when the mice were parallel to the
magnetic-field  vector was 6.2 W/kg, without any
enclosure. The animals changed positions more
frequently when some difference in  SAR  existed
relative to the position assumed.

Rats cannot easily learn to escape from or avoid high-
intensity electromagnetic radiation.  Carroll et al.
(1980) reported that 20 female Long-Evans  rats did
not go to a  marked-off floor area in  a multimodal
cavity to reduce the intensity of 918-MHz microwaves
(modulated at  60 Hz with 3-Hz mode  stirrer
modulation) from 60 W/kg to either 40, 30, 20, or 2
W/kg (Ta = 21.1 ± 2°C;  RH = 53 ± 10 percent). The
mean number of entries into the marked-off "safe"
area and the percentage of time spent there during
22-min  sessions were not  different when  the
microwaves  were on (for five 2-min periods) or off.
SARs > 60 W/kg  were associated with lethality
within 8 min of  continuous exposure in  tests  of a
separate group  of rats. However,  in the same
apparatus 10 rats quJckly learned to escape from an
800-//A electric foot shock by going to this marked-off
area within  a 22-min session. Shocked rats would
remain in this area over 90 percent of the  session
time. A later study (Levinson et al. 1982),  extending
these findings, showed  that rats spent significantly
more  time in the marked "safe" area  (which when
entered extinguished radiation and light) on the fifth
day of repeated light and microwave pairings when
microwave radiation was accompanied with the
photic stimulus than  when  the radiation was
presented without the light. Groups of four female
Long-Evans rats were exposed to a 350-lux stimulus
light alone,  microwave radiation alone (918 MHz
modulated  at 60  Hz with  a 3-Hz  mode  stirrer
modulation at 60 W/kg in the same multimodal cavity
used by Carrol et al. [1980]), the microwave radiation
in combination with the light, or faradic shock (800
//Arms) to the feet  and  tail  during alternate 2-min
periods of five successive 22-min daily sessions. The
average time spent in the marked area on the last day
(of  1320 s maximum)  was  about  100 s for rats
receiving light, 300  s for  rats receiving microwave
radiation alone, about 700 s for rats receiving
radiation plus light, and about 1200 s for rats
receiving  shock alone. Additionally,  acquisition
curves indicated rats were learning to escape from
both the microwave radiation presented alone and in
combination  with  light  over  the  course of the
sessions, but escape learning occurred sooner in the
rats given the multiple stimulus.

Sanza and de Lorge (1977) noted that their rats first
tried to jump out of a testing chamber and then
assumed a stationary position after failing to escape
during exposures to 2450 MHz at 37.5 mW/cm2 (SAR
= 7.5 W/kg).

5.5.3 Interactions with Other Stimuli
Interactions of microwaves with two other types of
stimuli have been reported to affect behavior. One
type of  stimulus is  chemical,  specifically,  several
commonly used  psychoactive drugs; the other is
physical, e.g., ambient temperature during exposure.
Response rates of rats performing on a fixed-interval
1 -min schedule of reinforcement beginning  0.5 h
after 2.5- to 20-mg/kg dosages of chlordiazepoxide
(Librium) (Thomas et al. 1979) were increased over
control values.  These increases were further
augmented if an exposure at an  average  power
density of 1 mW/cm2 to a  2450-MHz (PW)field (500
pulses/s, 2-/us pulse width, 100-mW/cm2 peak
power density; Ta = 23 ± 2°C) occurred in the half hour
between the injection and behavioral testing. In a
second study (Thomas and Maitland  1979), the dose-
response function  of rats given d-amphetamine
sulfate was shifted to a lower range after exposure to
microwaves with the same parameters and conditions
of exposure as those mentioned above. The shift to a
lower range of the dose-response function indicated
an  increased potency of the drug. After microwave
exposure, as opposed to sham exposure,  given
changes in response rate were produced by lower
drug doses. Rats in this second study were reinforced
for responses separated by more than 18s(DRL18 s).
Although the rat was in the near field in these studies,
and the exposure field may not have  been as uniform
as  if  it were in  the  far field, the average SAR
measured in a Styrofoam-insulated water model was
0.2 W/kg. Thomas and Maitland (1979) also showed
that 0.5-h exposure to microwaves, at the parameters
indicated above,  4 days/week when amphetamine
was not administered, shifted the dose-response
function of this drug to a  lower range when it was
administered  on  the fifth day of the week when a
microwave exposure did not occur.  In both of these
studies, exposure to microwaves alone or after saline
injection had no effect on behavior. The shift to a
lower range of a dose-response function indicated
                       5-60

-------
that an exposure to microwaves, which by itself did
not affect behavior, acted synergistically with
chlordiazepoxide or amphetamine to increase the
sensitivity of the organsm to the drug. Similar results
have  not  been  seen with chlorpromazine  or with
diazepam (Valium), an analog of chlordiazepoxide, in
experiments in  the same laboratory (Thomas et al.
1980).

In a related experiment, Monahan and Henton (1979)
showed  that chlordiazepoxide and,  with less
consistency,  chlorpromazine and d-amphetamine
altered response rates of mice trained in an operant-
conditioning procedure  to escape from or to  avoid
2450-MHz (CW) fields at an average dose rate of 46
W/kg (Ta  = 24  ± 0.5°C; RH = 50 ± 1.5 percent).
Although this experiment showed that drug effects
can interact to alter microwave exposure effects, the
design of the experiment does not allow any conclu-
sion to be drawn about interactions between drugs
and microwaves on performance.
Interactions with ambient temperature and humidity
were predicted by Mumford( 1969). Monahan and Ho
(1977) showed  that reduction in the  rate of energy
absorption by mice exposed in a waveguide to 2450-
MHz (CW) fields occurred at a low dose rate of 0.6
W/kg when  ambient temperature was  35°C but
required higher doses at lower ambient temperatures.
Although the mice were not directly observed, the
authors presumed the mice reoriented in the
microwave field to reduce the amount of absorbed
energy.


Gage (1979b) showed that overnight exposures at 5,
10, or 15 mW/cm2 to 2450-MHz (CW) fields (SAR =
0.2 mW/g per mW/cm2) at an ambient temperature
of 28°C (RH = 50 percent) reduced operant response
rates of male Long-Evans rats measured the morning
after exposure  was terminated. Similar exposures
when ambient temperature was22°C did not result in
reduced response  rates except after exposure at the
highest power density. The rats were reinforced with
food  on  a random  interval  1.33-min schedule.
Response rates in control sessions that did not follow
exposures ranged from  about 0.25  to  over  2
responses/s for the 1 2 rats but were consistent over
sessions for each  rat.

Two reports  have indicated that mammals  alter
thermoregulatory behavior in the presence of as little
as 5 mW/cm2 of 2450-MHz (CW) microwaves. In one
(Stern et al. 1979), six male rats were trained to press
a lever to switch on an infrared (IR) heat lamp for 2 sin
a cold chamber (Ta from 3.9 to 5.3°C).  Microwaves at
power densities as low as 5 mW/cm2 reduced the
rate of responding for the heat lamp. The reduced
response  rate occurring shortly after microwave
onset was a direct function of power density in the
range of 5 to  20 mW/cm2 and returned to base-line
values when microwaves were switched off. The rats
were  exposed in a far field (SAR = 0.2 W/kg per
mW/cm2) in which the distribution of power densities
varied within  11 percent of the mean value.
Adair and Adams (1980b) showed that thermoregula-
tory behavior of the squirrel monkey (a New World
primate  found in  equatorial jungle areas) was
significantly altered at power densities as low as 6
mW/cm2 in the far zone of a 2450-MHz (CW) field
within 10 min of onset of exposure. In this study, the
monkeys were trained to select a preferred ambient
air temperature by making an operant response to
obtain 15 s of 55°C air when the air was otherwise
15°C  or to obtain 15 s of 15°C air when the air was
otherwise 55°C. Preferred air temperatures ranged
between 35  and 36°C without  microwaves and
decreased as a direct function  of the microwave
power density during exposure. (The SAR in this
experiment  was determined calorimetrically on
saline models to be 0.2 W/kg per 1 mW/cm2.) As in
the experiment by Stern et al. (1979) with rats,
preferred temperatures returned to base-line levels
when the microwave field was extinguished.
In conclusion, there is ample evidence to suggest that
microwaves alter a variety of unlearned and learned
behaviors occurring  during and after exposures. In
most cases the behavior change can be described as a
reduction in the level of ongoing activity. However,
there  are some situations in which increased activity
has been seen. When measured, the magnitude of
behavioral change seems to be related to the power
density or SAR of the exposure. Behavioral changes
usually revert to base-line levels after removal of the
microwave  field.  The  above  studies of behavioral
effects caused by microwave  exposure are summa-
rized in Table 5-13.

5.5.4 Unresolved Issues
Most  unresolved issues regarding behavioral effects
of microwaves arise because observed findings have
not been verified within the  same laboratory or in
other laboratories. Repeating a study involves high
cost and the risk of failure to confirm a finding due to
small unnoticed differences, e.g., between standard
procedures in two  laboratories.  Possibly  for these
reasons verification has not been often attempted.
However, verification and systematic replication
would allow determination of the limits of conditions
within which  a behavioral effect may be expected, as
well as definition of the range of conditions adequate
to observe a threshold of effect.
There is no unifying hypothesis to explain all the
observed behavioral  changes.  The research on
interaction between microwaves and chemicals has
not been verified in independent laboratories, and it
has not been extended to determine whether the
interaction is limited to particular classes of
compounds.  Delineation of  differences between
effects of single and multiple exposures would help
                                                                     5-61

-------
Table 5-13.    Summary of Studies Concerning RF-Radiation Effects on Behavior
                                                               Exposure Conditions
Effects
Decreased exploratory activity
an swimming speed. AT increase
2.5°C
No effect on spontaneous activity
or activity and forced running


Increased locomotor activity

Decreased spontaneous activity and
food intake
No effect on spontaneous activity
or food intake
Decreased activity on stabilimetric
platform, no significant increase
in wheel running
Increased activity on stabilimetric
platform and in wheel running
Decreased time on treadmill and
inclined rod, decreased exploratory
activity, increased then decreased
shock sensitivity. Decreased
activity and shock sensitivity
persisted 90 days after exposure
Rectal temperature rise = 0.37°C
before start of test, AT = 1 .5°C
with microwaves
Response decreased during exposure
on random interval schedule (lowest
intensity for effect. AT = 1 .8°C)
Response decreased during exposure
(maximum effect) on random
interval schedule, AT = 1 .8°C
Decreased observing responses
on vigilance task, AT = 2°C

No effect on observing responses
Decreased observing responses
on vigilance task

No effect on observing responses
Decreased observing responses
on vigilance task
Response rate decreased on fixed
interval schedule in rats with
high base-line rates, spending time
away from lever
No effect on response rate

No effect on visual tracking task


Response rate decreased on FR and
increased on DRL schedules

Decreased length of runs and fewer
remforcers on FCN schedule
Decreased response rate on FR
operant schedule
Increased rate of missed observing
responses on vigilance task

Decreased rate of responding on
repeated acquisition task
Increased response rates in
extinction, decreased stimulus
control, no effect on Sidman
avoidance
Species
(Weight, g)
Young male
rat

Male rat
(160-180)


Female
rat (307)
Male rat
(360-410)
Male rat
(316-388)
Male rat
(350-375)

Male rat
(350-375)
Male rat





Female
rat

Male rat
(350-380)

Male rat
(357-382)

Male rhesus
monkey
(4kg)

Male squirrel
monkeys
(850-950)

Male rat
(362-400)
Male rat
(290-340)




Male rhesus
monkey
(6.2-7.9 kg)
Male rat
(120 days.
150?)
Male rat
(250-300)
Male rat
(284-439)
Young male
rat

Male rat
(275)
Female
rat


Frequency
(MHz)
2450 (PW.
multimodal cavity
120 Hz. AM)
10700(CW)
3000 (CW)
3000 (PW)
3000 (PW)
2450 (CW, multi-
modal cavity)
918ICW)

918(CW)

2450 (CW)


915(CW)

2375 (CW)





2450 (PW, multi-
modal cavity.
60 and 1 2 Hz AM)
500 (CW)


600 (CW)


2450
(120 Hz, AM)


2450
(120H,. AM)


1280(PW)
5620 (PW)
2450
(120 Hz. AM)




1200ICW)


2450 (CW)
2860 (PW)
9600 (PW)
2450 (PW)

2450 (CW)

2450 (PW,
AM, multimodal
cavity)
2800 (PW)

2450 (CW,
multimodal
cavity)

Intensity
(mW/cm'l
?


0.6-0.9
0.5-1.0
1 .5-2.0
24-26 (av)
10

10

2.5

5


5

0.5








25


10


72


16
50



10
26
37.5



8.8-18.4

20


5
5
5
5

10




5

10



Duration
(days x min)
1 x30


7.7 x 1440
7.7 x 1440
7.7 x 1440
17x 1440
110x300

21 x 600

91 x 600

80x480


80x480

30 x 420





10x0.17
10x0.5

1 x 11


1 x55


1 x60


1 x20
1 x30
1 x60

1 x60
1 x40
1 x40
1 x60



1 x60

1 x 120


1 x30
1 x30
1 x30
1 x30

1 x900

1 x30


1 x30

110x300



SAR*
(W/kg)
6.3


0.2
0.3
0.6
8.3
2.3

3.6

1.0

1.2


2.5

0.1





420
220

10


7.5


5.0


1.1
2.8


0.6-1.7
2.5
4.9
7.5



1.8-3.7

1.6


1.4
1.4
1.5
?

2.7

6.5


0.7

2.3



Reference
Hunt eta/ (1975)


Robert, era/ (1975)



Mitchell et at. (1977)

Moeetal (1976)

Lovely era/. (1977)

D'Andrea et al. (1979)


D'Andrea « a/. (1980)

Rudnevef a/. (1978)





Bermant et al. (1979)


D'Andrea et al. (1976)


D'Andrea et al. (1977)


deLorge(1976)


deLorge(1976)
deLorge(1979)


deLorge(1979)
de Lorge and Ezell
(1980)
Sanza and de Lorge
(1977)


Sanza and de Lorge
(1977)
Scholl and Allen
(1979)

Thomas et al. (1975)


Thomas et al. (1976)

Gage (1979a)

Hunt era/. (1975)


Schrot era/. (1980)

Mitchell er al. (1977)



                            5-62

-------
Table 5-1 3. (Continued)
Effects
No effect on flavor aversion test
No effect on flavor aversion test
Microwaves detected as stimulus
Microwaves detected as stimulus
Spending more time in shielded vs.
unshielded compartment
Spending equal time in shielded vs.
unshielded compartment
Spending more time in shielded vs.
unshielded compartment (occurrred
in first of 7 sessions)
Spending more time in unirradiated
compartment
Decrease in SAR at 24°C
Decrease in SAR when ambient
temperature increased from 20°C
to 35°C
No preferential orientation of rats
or mice in far field of plane wave
Cannot take specific action to
reduce intensity of irradiation
Augmentation of increased
response rates produced by
chlordiazepoxide
Shift to lower doses of
d-amphetamine on dose-response
curve for DRL schedule
No effect on dose response curve
for chlorpromazine or diazepam
Chlordiazepoxide reduced responses,
decreased avoidance responses, and
increased escape responses to
microwaves
Response rate decreases were
augmented after exposures at
higher ambient temperatures
Reduced responding for heat lamp
in a cold room
Selection of a lower ambient
air temperature
Species
(Weight, g)
Male rat
(360-410)
Male rat
(316-388)
Male rat
(409-427)
Male rat
(300-350)
Female
rat
Female
rat
Male rat
(250)
Male rat
Male mouse
(30-34)
Male mouse
(30-34)
Male rat
(200-360)
Male mouse
(25-33)
Female
rat
(290)
Male rat
(325-375)
Male rat
(250-300)
Male rat
(360-380)
Male mouse
(35-44)
Male rat
(315-365)
Male rat
(325-450)
Squirrel
monkey
(750-1100)

Frequency
(MHz)
918ICW)
918(CW)
2450 (PW,
120 Hz, AM,
multimodal
cavity)
918IPW)
1200(PW)
1200ICW)
1200IPW)
2880 (PW)
2450 (CW)
2450 (CW)
2450 (CW)
2450 (CW)
918(PW,
60 Hz, AM,
multimodal
cavity)
2450 (PW)
2450 (PW)
3800 (PW)
2450 (CW)
2450 (CW)
2450 (CW)
2450 (CW)
Exposure Conditions
Intensity Duration
(mW/cm2) (days x mm)
10 21x600
2.5 91 x 600
1 x 1
15 1 xO.5
0.2 4 x 30
2.4 4 x 30
0.4 1 x 90
9.5 9 x 60
1 x 15
1 x20
15 1 x60
15 1x60
5x2
1 1 x30
1 1 x30
1 4x30
1 1 x30
1 x30
10 1 x 930
5 1 x 15
6 1x10

SAR'
(W/kg)
3.9
1.0
0.6-2.4
7.5
0.2
2.2
0.4
2.1
28
43.6-0.6
3.3
6.2-12.3
(depending on
orientation)
60
0.2
0.2
0.2
0.2
46
2.0
1.0
1.0

Reference
Moeera/. (1976)
Lovely era/. (1977)
King era/. (1971)
Johnson era/. (1976)
Frey et al. (1975)
Frey era/. (1975)
Frey and Fold ( 1975)
Hjeresen er al. (1979)
Monahan and Ho (1976)
Monahan and Ho (1977)
Gage era/. (1979)
Carroll et al (1980)
Thomas era/. (1979)
Thomas and Maitland
(1979)
Thomas era/. (1980)
Monahan and Henton
(1979)
Gage (1979b)
Sterner al. (1979)
Adair and Adams (1980b)
•If measured SAR was not reponed, SAR was estimated when possible.

determine whether effects of chronic exposure are
qualitatively different from repeated measurement of
effects of each single exposure.

Behavioral  experiments have  used only  a limited
sample of microwave exposure conditions. Exploration
of frequency spectrum, modulations, waveforms, and
interactions between waves of different parameters
has hardly begun. Most behavioral work has used rats
as subjects. Animals more like humans in physical
size and shape have been studied only infrequently to
help extrapolate  findings  to humans. Specific
conditions of exposure in addition to the microwave
stimulus, such as ambient temperature and humidity,
have not been consistently controlled and reported
so that their influence  on behavior  may be
determined. There is  no information on exposure
effects of specific  or  limited areas of the body in
comparison to total-body exposure to evaluate the
effects of localized  energy absorption.
                                                                      5-63

-------
5.6 Special Senses
Joe A. Elder
5.6.1 Cataractogenic Effects

A cataract is an opacity in the crystalline lens of the
eye. These lens defects  may be clinically insignificant
or may cause partial or total blindness. The following
conclusions may be drawn from animal experiments
on the cataractogenic potential of RF radiation (Tables
5-14 and 5-15):

 1. RF radiation  is cataractogenic if exposure is of
   sufficient intensity and duration.
 2. For  single  acute  exposures, the threshold
   intensity for  cataract production exceeds 100
   mW/cm2. Multiple exposures  at intensities near
   threshold values for single  acute exposures
   results in lens opacities.
 3. The cataractogenic potential of RF radiation varies
   with frequency; the most effective frequencies
   appear to be microwave frequencies in the 1 - to
   10-GHz range.

 4. Similar ocular effects are produed by CW and PW
   radiation of the same average intensity.

 5. In contrast with the above conclusions, which are
   based on acute, near-field exposures to the eye or
   head,  no cataracts have been  reported in
   unrestrained  animals after  far-field exposure at
   power densities near lethal values or in rabbits
   exposed at 10 mW/cm2 (2450  MHz) for 180 days.
   Although the  RF  power density for cataract
   induction for long-term exposure  of animals
   (including humans) has  not  been  defined, the
   threshold  is,  most probably, significantly higher
   than that required to induce many  other
   physiological  changes.
The review of the literature on human cataracts is in
 Sec. 5.10, Human Studies.

5.6.1.1 Microwave Radiation in  Experimental
Cataractogenesis

The rabbit eye has been the experimental model most
 often used in animal studies because of its similarity
 in size and anatomy to the human  eye (Figure 5-3). Its
 diameter is ~ 75 percent that of the human eye; the
 cornea is as large, and although the lens is thicker, its
 diameter is the same as that of the human eye. In a
 typical study (Carpenter et al. 1960b), one eye of an
^anesthetized rabbit was exposed to  microwave
'radiation in the near field, and the nonirradiated eye
served as the control. The eyes were then examined
 at various intervals by an ophthalmoscope, slit-lamp
 biomicroscope,  or both instruments. The earliest
 positive reaction in t!~iis type  of study, occurring
within 24 to 48 h after a cataractogenic exposure, is
the appearance of one  or two narrow translucent or
 milky bands in the posterior cortex of the lens, just
under the capsule, which extends no further than the
lens equator. These bands can be seen only by slit-
lamp examination with an angled beam. If the ocular
injury is minimal, no further change occurs, and the
cortical banding disappears within a few days.
Otherwise, in  2  to 4  days after exposure small
granules appear in  the region of the suture of the
posterior lens. If a  more  intense reaction occurs,
larger numbers of granules appear over a larger area
within the next  few days, and small vesicles may
develop. These early changes  may develop  further
and  become either well-defined circumscribed or
diffuse cataracts. These changes in the lens remain
as permanent ocular defects. In general, it has been
found that microwave cataracts in rabbits  involve
only the posterior  cortex of the  lens, unless the
exposure  is so  intense that  the opacity extends
throughout the lens.

At exposure levels that cause cataracts, other ocular
reactions also occur, but they are transient and differ
in severity with  the intensity and duration of the
exposure.  Examples include  swelling and chemosis
of bulbar and palpebral  conjunctivae,  pupillary
constriction, hyperemia of iris and limbal vessels, and
vitreous floaters and filaments (Carpenter 1979).

5.6.T.2 Exposure Threshold Values for Cataracts

Following  the publication of reports demonstrating
that microwaves can cause cataracts in experimental
animals (Richardson et al. 1948; Daily era/. 1950a,b),
three laboratories (Williams et al. 1955; Carpenter et
al. 1960b; Carpenter and Van Ummersen 1968; and
Guy et al.  1975a) published  time vs. power-density
threshold curves for cataract induction in rabbits by a
single exposure to near-field 2.45-GHz radiation. The
time vs. power-density threshold  curves  originally
published  by the three  laboratories are similar in
shape  but are quantitatively  different.  The more
recent studies found lower  threshold values than
those reported by Williams et al. (1955). This finding
probably  reflects differences in  the irradiation
method and in techniques used to measure power
density. In fact. Carpenter (1979) has determined that
his power densities were 50  percent higher than
originally published. His corrected data are plotted in
Figure 5-4 along with the results of Williams et al.
(1955) for comparison. Guy et al. (1975a) replicated
Carpenter's work for single acute exposures with
essentially the same results, and also quantified the
threshold of cataractogenesis in terms of SAR (Figure
5-4). For  example,  these workers found  the
cataractogenic threshold for a 100-min exposure, the
longest period of irradiation, to be 150 mW/cm2 (138
W/kg peak absorption) (Figure 5-5).
The  cumulative  effect  of microwave  radiation  on
cataractogenesis in the rabbit has been examined by
repeated irradiation of the eye at power densities
below the threshold  for  single acute exposures
(Carpenter 1979). For example, daily 1 -h exposures at
                       5-64

-------
Table 6-14.    Summary of Studies Concerning Ocular Effects of Near-Field Exposures

                                                        Exposure Conditions
Effects
Cataract
Cataract






Cataract and other
ocular effects
No cataract
Cataract
Cataract

Cataract and other
Species
Rabbit
Rabbit






Rabbit

Rabbit
rabbit
Rabbit

Rabbit
Frequency
(MHz)
5,500 (CW and PW)
800 (CW)
4,200 (PW)
4.600 (PW)
5,200 (PW)
5,400 (CW and PW)
5,500 (CW and PW)
6,300 (PW)
2,450 (CW)

2,450 (CW)
2,450 (CW)
2,450 (CW)
10,000 (CW)
2,450 (CW)
Intensity Duration
(mW/cm2) (days x min)
470-785* 1 x 2-100
785*
785*
785*
500-785*
500-785*
500-785*
785*
x25
x17
x15
x5-12
x3-4
x2-3
x5
180* 1 x240
120-180 20x60
75 20 x 60
150 1 x100
295 1 x 30
375 1 x30
180 1 x140
SAR
(W/kg)
300-500t
500t
500t
500t
350-500t
300-500t
300-500t
500t



138tt


100tt
Reference
Birenbaum et al. (1969a)
Birenbaum era/. (1969b)






Carpenter (1979)

Carpenter (1979)
Guy era/. (1975a)
Hagan and Carpenter ( 1 976)

Kramarera/. (1978)
ocular effects
Second- to third-degree    Rhesus    2,450 (CW)
nasal burns; no ocular     Monkey
effects
No cataract; keratitis       Rabbit     35,000
(inflammation of cornea)              107,000
                       300
                       40
                       40
                        x22
                        x60
                        x60
                                    115tt
>175#
>238#
             Kramarera/. (1978)
Rosenthal era/. (1976)
'Estimate of average power density calculated by dividing the microwave power by the irradiated area (d = 1.27 cm) of the eye.
tEstimate based on the assumption that ail the incident power was absorbed by the eye (2 g).
ttMaximum SAR in the eye.                                        «
//Estimated SAR values for the cornea. (See text for discussion of Rosenthal et al. studies of frequency specificity.)
Table 5-15.    Summary of Studies Concerning Ocular Effects of Far-Field Exposures

                                                               Exposure Conditions
Effects
Species
Frequency
(MHz)
Intensity
(mW/cm2)
Duration
(days x min)
SAR
(W/kg)
Reference
No ocular effects, including no
lenticular changes
Acute ocular changes, e.g.,
hyperemia of lids and conjunctiva,
meiosis, anterior chamber flare,
engorgement of iris vessels, and
periorbital cutaneous burns; no
lenticular changes
Rabbit
3000 (CW)   100, 200    1 x 15, 30
                          300, 400,
                          500
                          x15
       14, 28*   Appleton et al. (1975)


       42, 56*
       70*
Death

No catarats


No cataracts

No ocular effects
No ocular effects



Rabbit


Rabbit

Rabbit
Monkey
(M. mulattaj


385 (CW)
385 (CW)
468 (CW)
2450 (CW)

2450 (CW)
9310(PW)

300
500
60
30
60t
10

10
1 50 (av)

1 x30
1 x15
10x 15
10x90
10x20
5 x 480
(x 8- 17 weeks)
180 x 1380
30-40 x 294-665ft

42*
70*
48*
24*
8.1
1.5*

17#




Cogan et al. (1958)


Ferri and Hagan (1976)

Guy era/. (1980b)
McAfee era/. (1979)

'Estimated average whole-body SAR values (Durney et al. 1978, Figure 31).
tWaveguide average whole-body exposure.
ttTotal exposure time in minutes for the entire 30- to 40-day experimental period.
//Maximal SAR in head.
                                                                                     5-65

-------
Figure 5-3.  Cross-sectional sketch of the human (left) and the rabbit (right) eye (from Birenbaum ft al. 1969a, Figure 1).

                                    f Ciliary Body

      Retinas
     Choroid

     Sclera
  Fovea
Optic
Nerve
      Central
     Artery and
    Vein of Retina
Suspensory
 Ligament
  of Lens

    Iris
   Lens

  Cornea

   Anterior
   Chamber
   Posterior
   Chamber
Conjunctiva

 Ciliary Body

      Cornea


      Lens

     Anterior
     Chamber
                Optic
                Nerve
                           Vitreous Body    Conjunctiva
                       Vitreous
                        Body
                                       Posterior
                                       Chamber
180 mW/cm2 for 13 to 20 days were found to be
cataractogenic in 8 of  10 animals, whereas single
exposures at this power density were not effective. At
150 mW/cm2, 4 of 10  rabbits gave a positive
response after 18 to 32 daily exposures. No cataracts
were observed after 20 daily 1 -h exposures at 75
mW/cm2.

The power densities used in the cumulative-effect
experiments were only slightly lower than those that
cause cataracts after a  single exposure (Figure 5-4).
For example, the lowest effective power density (120
mW/cm2) given for 1  h  daily for 20 to 24 days
produced only one positive response in nine animals
(See Table I in Carpenter 1979). But that would be a
"thermalizing" level if extended to the whole body. If
the  entire body of an unanesthesized rabbit is
exposed at  power  densities  similar to those that
cause cataracts  under near-field conditions,  the
animal exhibits signs of acute stress; Appleton et al.
(1975) reported that rabbits became heat stressed
and struggled out  of  the field during  a  15-min
exposure at 100 mW/cm2. This report is described in
more detail in  Sec.  5.6.1.4, along with other
experiments in which no evidence of cataracts was
observed in animals exposed in the far field. In one of
these studies, Guy et al. (1980b) found no change in
the eyes of rabbits exposed at 10  mW/cm2 (2450
MHz) for 23 h/day for 180 days. Although a
cumulative effect on the  lens of rabbits has been
described by Carpenter and his colleagues at high
exposure  levels, the report by Guy et al. (1980b)
provides evidence against such an effect at power
densities <10 mW/cm2.

5.6.1.3 Frequency Specificity
As indicated above, most studies of experimental
cataractogenesis were conducted at 2.45 GHz, but

                        5-66
         opacities in rabbit eyes have been reported after near-
         field exposures at 0.8,4.2,4.6,5.2,5.4,5.5,6.3, and
         10  GHz (Birenbaum et al.  1969a,b; Hagan and
         Carpenter 1976). In several studies, the cataracto-
         genic potential of different frequencies  was
         addressed. For example, after Hagan and Carpenter
         (1976) determined the relative effects of 2.45- and
         10-GHz  (CW) radiation on  the  rabbit eye,  they
         concluded that the cataractogenic potential for single
         acute exposures is greater at the lower frequency. At
         both frequencies, the opacities were characteristically
         located in  the posterior subcapsular cortex of the
         lens, although the initial appearance and subsequent
         development  differed.  These  differences  probably
         reflect differences in the pattern of absorbed
         microwave energy in the eye due to the different
         depths of penetration of the radiation at these two
         frequencies.

         Guy et al. (1974,1975a) measured the distribution of
         absorbed energy in rabbit eyes exposed to 918- and
         2450-MHz radiation and found the  patterns to be
         significantly different. At 2450  MHz,  energy
         absorption was  maximal in the vitreous body at  a
         point midway between the posterior surface of the
         lens and the retinal surface (Figure 5-5). The locus of
         peak absorption thus correlates  well with the
         observation of irreversible changes in the posterior
         cortical  lens. Exposure to 918-MHz  fields  in  a
         specially devised cavity resulted in relatively uniform
         absorption in the eye, but maximal absorption was
         only ~ 25 percent of the peak absorption at 2450 MHz
         (Figure  5-6). Therefore, one would expect the
         threshold for cataractogenesis in rabbits exposed to
         918 MHz to be considerably higher than the threshold
         at 2450 MHz. But more important, at 918 MHz, peak
         absorption in the rabbit brain was 36 percent higher
         than in the eye. It is possible that lens effects or ocular

-------
Figure 6-4.  Time and power-density threshold for cataracto-
           genesis in rabbits exposed to near-field 2450-MHz
           radiation; values of maximal SAR are also given
           (from Guy et al. 1975a, Figure 7).
  700'
  600-
-500H

o
x
Jsoo-l
l_
0>

£200-1
  100-
             Carpenter et al. (1960b, 1979)
                      Williams et al. (1955)
       Guy et al. (1975a)
           20     40     60    80
                 Exposure Time (min)
                                            600
                                           -500
                                           ^400
                                                C

                                             300 «
                                                a
                                                E
                                            200
                                           -100
                                    100
Figure 6-5.
  1.25-1
          Distribution of energy absorption rate (W/kg) per
          mW/cm2 incident power density in the rabbit's eye
          and head exposed to 2450-MHz radiation (from
          Guy et al. 1974, Figure 3).
E
o
$ 1 00-|
E

I
5*0.75-

5
  0.50-
  0.25-
  0.00
        Posterior Pole
                       Retina
       — Cornea
              -i	1	1	=»>-
               1234
               Depth from Cornea! Surface (cm)
changes may not occur before more severe damage
occurs in other sensitive tissues, such as the brain,
during exposure to 918-MHz fields.

Both  Hagan and Carpenter (1976)  and Guy et al.
(1975a) used exposure  systems that  applied
microwave energy across an air space to the eye, and
both reported  lenticular  cataracts in the  posterior
                                                    Figure 5-6.
                                                      0.30-,
           Distribution of energy absorption rate (W/kg) per
           mW/cm* incident power density in rabbit's head
           and eye exposed to 918-MHz radiation (from Guy
           atal. 1974. Figure 27).
                                                    E
                                                    i
                                                      0.25-
                                                §  S
                                                ^_  m
5 0.20 H
EC
C
o
                                                      0.15-
                                                         -L
                                                          0
                       234
                Depth from Cornea! Surface (cm)
subcapsular  cortex.  Birenbaum  et al. (1969b)
produced cataracts in  the anterior cortex of rabbit
lens with an exposure system that applied PW
microwaves to the cornea! surface. Furthermore, as
the frequency decreased from 6.3 to 5.2 to 4.6 to4.2
GHz,  longer exposure times at a  constant field
strength were required to produce lens  defects.
Under similar experimental conditions, even longer
exposure times were required to induce cataracts at a
lower frequency,  i.e., 0.8-GHz  (CW) radiation.
Furthermore,  Birenbaum  et al.  (1969a) found no
substantial difference in the cataractogenic threshold
values  for  CW and  PW  5.5-GHz  radiation  and
concluded that the average, not the peak,  rate of
energy  absorption determines whether lens injury
will occur. The peak power density of the pulses was
1000 times greater than the estimated average
power densities shown  in Table 5-14.
Rosenthal et al. (1976)  examined the effects of 35-
and 107-GHz radiation on the rabbit eye.  At both
frequencies keratitis (inflammation  of the cornea)
occurred at lower intensities than required to produce
any other demonstrable ocular effect, such as  lens
injury (cataract) or iritis. Irradiation at 107 GHz  was
more effective  in  producing  immediate  cornea!
damage, but this change was generally gone by the
next day. Effects at 35 GHz were persistent, were
almost  always present the next day, and were
associated with marked injury to the cornea!
epithelium. Effects on the cornea correlate well with
the pattern of microwave-energy absorption because
most  of the energy at these high  frequencies  is
absorbed in the outer regions of the eye. The earliest
stage of keratitis  or minimal cornea! stromal injury
occurred  after a  30-min exposure  at an incident
                                                                        5-67

-------
power of 50 mW or after a 60-min exposure at 25
mW. Estimates of the rate of energy absorption in the
eye at the lower incident power (25 mW) are 35 mW
at 35 GHz and 47.5 mWat 107GHz(Rosenthalera/.
1976); therefore, the average SARs of a rabbit eye
weighing 2 g are 17.5 and 23.8 W/kg, respectively.
Since maximal  absorption occurs  in  the  outer
structures of the eye, the SAR in  the cornea  is
estimated to exceed the average SAR of the eye by
more than one order of magnitude.

Although the above data cannot be compared directly
because of widely varying experimental procedures,
these results indicate that the potential for cataract
induction in rabbits is higher in the frequency range
between 1 and 10 GHz than at either lower or higher
microwave frequencies. Power densities that would
cause  cataracts  at frequencies below 1 GHz and
above 10 GHz give rise to insult in other ocular sites
and  extra-ocular tissues.  Rosenthal  et  al. (1976)
found  that 35- and  107-GHz radiation primarily
affected the outer structure of the eye, the cornea
for  example;  and at  918  MHz, Guy  et  al. (1974)
showed that maximal energy absorption occurred in the
brain, not in the eye.

The effects of near-field microwave exposures on the
rabbit eye have been summarized by Cleary (1980) as
follows:
  The induction of cataracts in experimental animals,
  principally New Zealand white rabbits, has been
  described by a number of  investigators using a
  variety of microwave exposure modalities. Gene-
  rally microwave field intensities necessary for
  cataract induction in the rabbit are such that acute
  whole  body exposures  would  be  lethal due to
  hyperthermia. Most cataract studies have, there-
  fore, employed focused or near-zone fields which
  limit exposures to the  head  or eyes. Localized
  thermal trauma is  still  of such a  magnitude to
  necessitate the use of general or local anesthesia.
  Cornea!  irrigation with physiological  saline
  solutions have also been used to prevent corneal
  damage due to tissue dehydration during microwave
  exposure.  Anesthesia and  corneal  irrigation, as
  well as  air temperature  and humidity, may
  significantly affect the temperature of  ocular
  structures.
The experimental results strongly  suggest that
radiation-induced temperature elevation may be
essential for the cataractogenic effect of microwaves.
Additional  evidence for  this position  has been
provided  by Kramar et a I. (1975), who reported that
rabbits kept  under general  hypothermia during
irradiation at known cataractogenic levels of 2.45-
GHz (near-field) radiation did not develop cataracts.
This study is described more fully in Sec. 5.6.2.
Kramar et al. (1978) did a comparative study of the
cataractogenic threshold and SAR patterns in the
eyes of rabbits and rhesus monkeys exposed in the
near field at 2450 MHz. In this experiment, the
radiator was designed to simulate a leaky microwave
oven  door.  In rabbits, a time  and power-density
threshold curve for cataractogenesis was found that
was similar to earlier results (see Figure 5-4), and the
peak SAR occurred within the lens. These authors
reported that the microwave exposure caused
immediate effects similar to those described  by
Carpenter era/. (1960b), i.e., constricted pupil, dilated
conjunctiva!  and  iris  vessels, a  turbid anterior
chamber, and a milky band in the posterior cortex of
the lens.

The monkey eye was irradiated at 400 mW/cm2  for
30 or 60 min or at 500 mW/cm2 for 60 min with the
applicator centered, over one eye. Following these
exposures at time and power density values that
exceeded those  required to  produce cataracts in
rabbits, no lens  opacities were observed in three
monkeys for a period of 13 months. At the highest
exposure, lid edema, contracted pupil, and changes in
the anterior chamber were noted. It is interesting that
the SAR pattern in the monkey eye revealed that the
peak SAR occurred in the anterior chamber and not in
the lens. Additional evidence for a marked difference
in the pattern of energy absorption in the monkey was
the observation that the exposures caused varying
degrees of nasal burns in the three animals. A fourth
monkey  was exposed  to an applicator centered
between the eyes. An exposure of 300 mW/cm2 for
22 min caused second- to third-degree burns on the
nasal bridge, but the eyes were not  affected.
These  results show that exposure  conditions that
cause cataracts in rabbits do not induce cataracts in
monkeys. Monkey eyes are better shielded by facial
bones and tissues than rabbit eyes. Because of the
different anatomical features, the peak SAR, i.e., peak
temperature  increase, was not observed  at the
posterior surface of the lens. For example, the peak
retrolental temperature following an exposure of 300
mW/cm2 was only40.2°C in the monkey but 45.1°C
in the  rabbit. Kramar et al.  (1978)  stated that  an
additional factor that helps explain the differential
effects on the eyes of rabbits and monkeys might be
the relative as well as the absolute size of the lens.
The anterio-posterior diameter of the adult rabbit lens
is ~ 7 mm, which represents almost one-third of the
eye's diameter. In the rhesus monkey and man, the
eye is larger but the lens is about 4 mm thick (Figure
5-3).- The smaller lens is probably more effectively
cooled by the larger pool of surrounding fluid. Since
the facial features and eye structure of human beings
are more like those of the monkey than those of the
rabbit,  extrapolation of the results with monkeys is
more useful. Based on studies at 2450 MHz, one may
conclude that under conditions of acute exposure, the
threshold for cataracts in monkeys, and presumably in
human beings,  is  greater than the  threshold for
rabbits. The  reader should note that the report by
                       5-65

-------
 Kramar et al. (1978) vividly describes a number of
 other clearly undesirable effects, e.g., facial burns,
 that occur at subthreshold conditions for cataracts in
 monkeys.
 5.6.1.4 Far-field Exposure Studies
 In contrast with the acute, near-field exposures that
 can cause  cataracts and other ocular effects,
 cataracts have not been produced in rabbits whose
 entire bodies were exposed to radiation in the far field
 (Table 5-15). Appleton et al. (1975) exposd anesthe-
 tized rabbits to far-field radiation at 3000 MHzfor 15 or
 30 min at 100 or 200 mW/cm2. No ocular changes
 were observed during or  immediately after exposure.
 Fourteen daily examinations and four weekly
 examinations, followed by monthly examinations for
 1  year,  revealed  no lenticular changes. During
 exposure at  higher levels (300,400, or 500 mW/cm2
 for 15 min)  animals exhibited acute ocular changes
 consisting of  hyperemia of lids and  conjunctiva,
 meiosis, anterior chamber flare, engorgement of iris
 vessels, and periorbital cutaneous burns. Subsequent
 examinations revealed  no  morphologic lenticular
 abnormalities. The authors  concluded  that  "It is
 noteworthy  that one year after a single microwave
 exposure, sufficient in intensity to cause both thermal
 cutaneous and acute gross ocular effects, no lens
 changes or  cataracts were observed."  It is also
 noteworthly that these power levels and durations
 were well above  the cutaneous sensation  level,
 because unanesthetized animals became heat
 stressed and struggled out of the field during a 15-
 min exposure at 100 mW/cm2. Exposures at 300
 mW/cm2 for 30 min or 500 mW/cm2 for 15 min were
 lethal to some of the rabbits.
Ferri  and Hagan (1976) exposed unanesthetized
rabbits to 2450-MHz (CW) radiation in the far field at
10 mW/cm2,8 h/day, 5 days/week for 8 to 17 weeks.
Weekly examinations of the eye showed no abnormal
changes  during the study, and no post-irradiation
changes were  observed during the following 3
months.  Guy et al.  (1980b) exposed four rabbits
individually (2450 MHz, 10 mW/cm2)for 23 h/day for
 180 days; the peak SAP in the head was 17 W/kg. An
equal number of animals were sham-exposed. The
authors  stated that "Aside from  normal  aging
changes  found  in the lenses of the  8 animals,  no
differences  were  noted in the eyes of  the two
groups."
Cogan et al.  (1958) found no cataracts in rabbits 4
weeks after  exposure at 385 MHz; the rabbits were
irradiated twice weekly for 5 weeks at 60 mW/cm2 for
15 min or 30 mW/cm2 for 90 min. Six weeks after
exposure in a waveguide,  no cataracts were observed
in rabbits irradiated at 468 MHz, 60 mW/cm2 (SAR =
8.1  W/kg), for 10 days (20 min  daily). Although the
authors reported that the exposure  levels at both
frequencies  were near lethal values, the 8.1-W/kg
SAR is considerably lower than the value  other
 investigators have found to be sublethal to rabbits
 (Table 5-15). The environmental conditions (tempera-
 ture,  airflow, etc.) within the waveguide were not
 given; therefore, one must assume that the ambient
 conditions  were significantly different from  normal
 values or that the measured SAR is in error.
 McAfee  et al. (1979) trained unfettered monkeys
 (Macaca mulatta) to expose their faces and  eyes to
 9.31-GHz (PW) radiation at 150 mW/cm2 average
 power density (peak power density = 300 W/cm2).
 Over  a period of  about three months, the animals
 were irradiated for a total of 294 to 665 min during 30
 to 40 daily sessions. No cataracts or cornea!  lesions
 were observed in these  monkeys during a 1-year
 period following irradiation.

 5.6.2 Unresolved Issues
 Exposure of the rabbit eye to microwave radiation at
 sufficient power densities and durations causes an
 immediate increase in intra-ocular temperature, and
 after a latent period of a few days, opacities develop in
 the posterior subcapsular cortex of the lens (Kramar
 et al.  1975). This sequence of events has led to the
 assumption that microwave-induced lens opacities
 are thermally caused. Several experiments have been
 designed to test directly the  cause-and-effect
 relationship between temperature  increase  and
 cataract formation. Kramar et al. (1975) exposed
 rabbit eyes to cataractogenic levels of microwave
 radiation while the animal's body was submerged in
 cold water. By this means, microwave-induced intra-
 ocular temperature was limited to < 41 °C, and no
 lens opacities developed. In a  later experiment,
 Kramar et al. (1976) used heated water to produce
 ocular and rectal temperatures characteristic of those
 in rabbits  exposed to a cataractogenic level of
 microwaves. Although the vicinity of the lens was
 heated to temperatures above those  known to be
 associated with  microwave cataracts, no lens
 opacities were observed; however, the rate of ocular
 temperature rise was about  one-tenth the rate of
 increase  with microwaves. Kramar et al.  (1976)
 concluded that a combination of a sharp temperature
 gradient  and rapid rise in  temperature  following
 irradiation may be more traumatic to the lens than a
 critical temperature per se.

 Carpenter et al. (1977) reported no cataracts of the
 posterior cortex of the lens in rabbits after 6 weeks of
treatment in which "the eye was heated at the same
 rate, to the same extent, and for an equal period of
time as it would experience during a cataractogenic
microwave  exposure, the difference being that the
equal  heating was provided by other means, namely,
direct application of heat to the surface of the eye." In
addition, elevating retrolental and rectal temperatures
to values characteristic of a cataractogenic microwave
exposure, through a combination of restricted body
 heat loss and irradiation of one eye to power densities
                                                                     5-69

-------
slightly below the cataractogenic threshold, produced
cataracts in only 3 of 10 rabbits. According to their
hypothesis, if cataracts were solely of thermal origin,
all animals given the  two treatments should have
developed cataracts. Carpenter et al. (1977), therefore,
concluded that the increase in intra-ocular tempera-
ture occurring during microwave irradiation is not the
sole causative factor in microwave cataractogenesis.
The reason for the apparent disagreement between
the conclusions of Carpenter et al. (1977) and Kramar
et al. (1976) probably rests with the difficulty  of
duplicating by nonmicrowave heating techniques the
temporal and spatial temperature profiles induced by
microwave irradiation of  the  eye. Note that the
temperature at a single site in the eye was the basis
for evidence of duplicating microwave heating of the
eye  at the  same rate and to the same extent.
Furthermore, even though retrolental and rectal
temperatures characteristic of cataractogenic
microwave exposures were  produced, the overall
effects on the  rabbits were more traumatic than a
near-field microwave exposure to one eye. Within 24
h after heated water was applied directly to the eye,
almost all the  eyes exhibited hemorrhage into the
anterior cavity, so that observation of the lens became
impossible. Only 21 of 32  rabbits survived the heat
treatment. Of these,  five developed lenticular
cataracts  of the  anterior cortex, and only  three
developed cataracts of the posterior cortex of the lens
that are  characteristic  of  microwave-induced
opacities (Carpenter et al.  1977).
As mentioned above, the difficulty of duplicating  by
nonmicrowave techniques the time-temperature
profile of microwave energy absorption in the eye is
probably responsible for the unsuccessful attemptsto
prove that an elevation of temperature is responsible
for  microwave cataracts (cf.  Kramar et al.  1976;
Carpenter et al.  1977). On the other hand,  strong
evidence for thermalization being the causative factor
in microwave cataracts is provided by the experiment
of Kramar et al. (1975), which showed that cataracts
were not produced in hypothermia rabbits receiving a
cataractogenic microwave exposure. At present, it is
generally understood that intense localized exposure
of the eye for substantial durations (i.e., 150 mW/cm2
for  100 min) is necessary to induce cataracts  in
laboratory animals,  and that such acute  exposures
cause death by hyperthermia  if the entire animal is
irradiated.

5.6.3 Auditory Effects
When the human head is exposed to PW RF radiation,
an audible sound described as a click, buzz, chirp, or a
knocking sensation is perceived by some individuals;
the sound appears to originate from within or behind
the head. This auditory phenomenon is called "RF
sound" or "RF hearing." Our present knowledge of
RF hearing is summarized  below.
1. The RF sound occurs only upon exposure to PW
   sources;  it  appears to originate from within or
   near the back of the head regardless of orientation
   of the head in the RF field, and the sound varies
   with pulse width and pulse repetition rate.
2. RF hearing has been reported at frequencies of
   216 to 7500 MHz. The effect has been found to
   occur at average power densities as low as 0.001
   mW/cm2 with peak power densities in the range
   of 100 to 300 mW/cm2.  Effective pulse widths
   vary from 1 to 1000//s; however, it is the energy in
   the first 30 fjs or so of the pulse that determines
   the threshold and loudness levels regardless of
   pulse widths greater than about 30 /us (Table 5-
   16).
3. The ability to perceive RF pulses has been shown
   to be related to bone-conduction hearing and to
   the ability to hear high-frequency acoustic waves
   above 5 to 8 kHz.
4. The available data support the conclusion that the
   RF auditory effect is evoked by a  mechanism
   similar to that for conventional acoustic stimuli
   and that the  primary site of interaction is
   peripheral to the cochlea.
5. The most generally accepted mechanism respon-
   sible for the RF auditory sensation isthermoelastic
   expansion. That is, the absorption of the energy in
   a brief RF pulse causes a  small temperature rise
   (~ 10~*°C) in a short time(~ 10/us), which results in
   thermoelastic expansion  of  matter  within the
   head, which then launches a pressure wave that
   is detected by the hair cells in the cochlea via bone
   conduction. After stimulation of the auditory
   nerve in the high-frequency region of the cochlea,
   transmission of  the PW-induced response to the
   auditory center of the brain follows the same
   auditory pathways as do all acoustically induced
   responses.
5.6.3.1 Human Perception of Pulsed RF
Radiation
In their recent review article on the  RF hearing
phenomenon, Chou et al. (1982) wrote:
  The earliest report we have found on the auditory
  perception of pulsed microwaves appeared in 1956
  as an advertisement of the Airborne Instruments
  Laboratory in Vol. 44 of the Proceedings of the IRE.
 .The advertisement described observations made in
  1947 on the hearing of sounds that occurred at the
  repetition rate of a radar while the listener stood
  close to a horn antenna. When the observers first
  told their coworkers in the Laboratory of  their
  hearing experiences, they encountered skepticism
  and rather pointed questions  about their mental
  health.
Frey (1961) was the first to study systematically the
human auditory response to pulse-modulated
                       5-70

-------
radiation. The subjects, who were more than 30 m
from  an enclosed  antenna,  reported hearing a
transient  buzzing sound upon  exposure to the
intermittent rotating beam. The apparent location of
the sound, which was described as a short distance
behind the head, was the same no matter how the
people were oriented  in the RF field. When an RF
shield (aluminum flyscreen) was placed between the
subject and the RF source, the subject did not perceive
RF sounds (Frey 1973). When earplugs were used, a
reduction in the ambient noise level and an increase
in the RF sound level were reported. The sensitive
area for detecting RF sounds was later described as a
region over the temporal lobe of the brain, because
the placement of a small piece of metal  screen (5x5
cm) over this area completely stopped the RF sound
(Frey  1962).

Guy et al. (1975b) described the effect of PW radiation
on two of the coinvestigators. Three pulses, 100 ms
apart, were presented each  second to keep the
average power density below 1  mW/cm2.  Each
individual pulse was heard as a distinct and separate
click, and short pulse trains were heard as chirps with
the tone pitch corresponding to the pulse repetition
rate (PRR). The RF.sound appeared to originate from
within or near the back of the head. This report also
included the note that transmitted digital codes could
be accurately interpreted by the subject when the
pulse generator was  keyed manually. Guy  et al.
(1975b) also  reported that the threshold for RF
hearing was lower when earplugs were used.
In a study  by Constant (1967), the RF sound was
described as being in the area of the ear on the side
opposite to the one that was irradiated. All three of his
subjects readily detected 2-//S pulses, whereas0.5-//s
pulses were not perceived. All three experienced a
buzzing sensation  at  PRRs greater than 100/s,
whereas individual pulses were heard when subjects
were exposed to PRRs below 100/s.
Five of eight human  subjects reported hearing
distinct clicks either inside the head or behind the
head when exposed  to  15-yus pulses (Cain and
Rissmann 1978). The remaining three people heard
faint clicks when the pulse width was increased to 20
/us.

5.6.3.2 Effective Radiation Parameters
In the initial report by Frey (1961), human subjects
perceived  PW radiation at frequencies of 1310 and
2982 MHz.  Although  the peak of power-density
thresholds for RF hearing was 266 mW/cm2 for
1310-MHz and 5000 mW/cm2 for 2982-MHz fields,
the average power density  thresholds were 0.4
mW/cm2  and  2 mW/cm2, respectively. When
earplugs were used to attenuate the ambient noise
level of 70 to 80 dB, the subjects reported an increase
in the RF sound levels.
In the following  year, Frey (1962) reported  that
humans could perceive PW radiation at 425 MHz with
an average power density threshold of 1 mW/cm2;
the peak of power  density was 263 mW/cm2. A
frequency of 8900 MHz was not effective even at an
average power density  of 25  mW/cm2; the peak of
power density was 25,000 mW/cm2. At 216 MHz, the
lowest effective frequency reported in the literature,
the average power density threshold was 4 mW/cm2;
the peak of power density was 670 mW/cm2 (Frey
1963).
Table 5-16.    Summary of Studies Concerning Auditory Effects of RF-Radiation in Humans

                                                Exposure Conditions
Effect
RF hearing
"distinct'
clicks"
RF hearing:
buzz heard
at PRR > 100:
individual
pulses heard
at PRR < 100
No auditory
response
RF hearing
"buzzing sound"
RF hearing
"buzz.
clicking, hiss.
or knocking"

No auditory
response
RF hearing
"buzzing sound"
RF hearing
"clicks, chirps"
RF hearing

Comment
Threshold
values











Threshold
values





Threshold
values
Threshold
values
Polytonal
sound
Number
of
Subjects
8


3







4

Not
given





8
7
1

18

Frequency
(MHz)
3,000


3.000

6.5OO



9.500

1.245
1.245
216
425
425
425
425
8900

1.310
2982
2.450

BOO

Pulse
Repetition
Rates Is")
05


<1OO-1.0OO

0 00-1.000





50
50
_
27
27
27
27
400

244
400
3

1.000-1.200

Pulse
Width
(us]
5
10
15
1-2

1-2



05

10
70
_
125
250
500
1.000
25

6
1
1-32

10-30

Peak
Intensity
(mW/cm'l
2500
225-2.000
300-1.000
2,500-50.000

2.500-50.000





370
90
670
263
271
229
254
25.000

267
5.000
1.250-40.000

>SOO

Average
Intensity
(mW/cm'l
0.006
0 001 -0.01
OOO2 -0.007
5

5





019
032
40
1.0
1.9
3.2
7.1
25

04
2
0.1

_

Energy/
Pulse Noise Level
(jjj/cm1) (dB)
125 45 (* plastic
2.3-20.0 foam earmuffs)
4.5-15.0
40









70-90 1+ ear
stopples)



70-90 (+ ear
stopples)
70-80
(+ earplugs)
40- 45

— 40 (» ear
stopples)
References
Cam and Rissmann ( 1 978)


Constant (1967)







Frey and Messenger
(1973)
Frey (1962. 1963)




Frey (1962)

Frey (1961)

Guy era/. (1975b)

Tyazhelov « «/. (1979b|

•Calculated peak-absorbed-energy density per pulse is 16 mJ/kg
                                                                     5-71

-------
In the study by Constant (1967), three  human
observers were exposed to PW radiation at 3,6.5, and
9.5 GHz at an average power density of 5 mW/cm2
(pulse width was 0.5 to 2.0 //s;  PRR was up to
1000/s). Only two of the three observers perceived 3-
and 6.5-GHz radiation; none experienced a response
to the highest frequency. Cain and Rissmann (1978)
reported that all eight of their subjects  heard RF
pulses at 3 GHz. In this study, plastic foam earmuffs
were worn to attenuate the ambient noise, which
was 45 dB. The average threshold energy density per
pulse was 10.6 yuJ/cm2 (range of values was 3.4 to
17.5 /yJ/cm2). Expression of the peak power
threshold (225  mW/cm2)  for the most sensitive
individual in units of average incident power density
yields 0.001 mW/cm2, the lowest threshold value for
human beings found in the literature. For a given peak
power, average power density depends on the pulse
repetition rate;  the low threshold average power
density was a result of a pulse repetition rate of only
0.5/s(Table5-16).
The range of microwave pulse widths varied from 1 to
32 fjs in the study by Guy era/. (1975b) on one human
subject. The results indicate that regardless of the
peak power of  the pulse or the pulse width, the
threshold for RF hearing of 2450-MHz radiation was
related to an energy density of 40 //J/cm2 per pulse,
or energy  absorption per pulse  of 16  fiJ/g. as
calculated with the  aid of a spherical model. The
background noise of the exposure chamber was 45
dB. When earplugs were used, the threshold level
decreased to 28 /yJ/cm2. The threshold for a second
subject, who had a hearing deficit, was approximately
135 /uJ/cm2. Guy et al. (1975b) stated that two pulses
which occurred within several hundred microseconds
of each other were perceived as a single pulse with
energy equal to the sum of the two pulses.
The  human studies  cited  above  indicate that  the
highest effective frequency of RF hearing is between
6.5 and 8.9 GHz  and that the lowest effective
frequency is 216 MHz. Also, the results describe
other radiation parameters (peak power density,
energy density per  pulse, and pulse width) that are
important in determining the threshold for RF hearing
in humans. Again, the RF  hearing  phenomenon
depends on the energy in a single pulse and not on
average power density.
5.6.3.3 Dependence of RF Hearing on Acoustic
Hearing
Standard audiograms measure hearing thresholds
for air conduction at  acoustic frequencies of 250 to
8000 Hz and for bone conduction to 4000 Hz.  Cain
and Rissmann (1978) measured the hearing ability of
eight subjects over the frequency range of 1 to 20 kHz
in addition to determining their standard audiograms.
They  found that, although there was no apparent
correlation between  the ability to perceive pulsed
 microwaves at 3000 MHz and hearing ability as
measured by standard audiograms,  there was a
strong correlation between microwave-hearing
threshold and hearing thresholds to air-conducted
acoustic signals above 8 kHz. For example, three of
the subjects who had normal hearing below 4 kHz
could not hear microwave pulses of less than 20-fJS
duration under conditions in which the other subjects
could perceive RF sounds. All three had a hearing
deficit at frequencies above 8 kHz.
Frey (1961) compared human acoustical hearing and
RF hearing and reported that a  necessary condition
for perceiving the RF sound was the ability to hear
audiofrequencies above approximately  5 kHz,
although not necessarily  by air conduction. This
conclusion was based on one subject who had normal
air-conduction hearing below 5 kHz but failed to hear
the microwave pulses. The person was subsequently
found to have a substantial loss in bone-conduction
hearing. On the other hand, a subject with good bone-
conduction  hearing but with poor air-conduction
hearing perceived the RF sound at approximately the
same power density that induced threshold perception
in subjects with normal hearing. The studies by Cain
and  Rissmann (1978) and  by Frey (1961) show RF
hearing to depend on high-frequency hearing above 8
kHz and bone-conduction hearing at lower acoustic
frequencies. It is interesting that the 1956 report from
Airborne Instruments Laboratory stated that two
persons with hearing loss above  5  kHz did not
perceive RF  sounds as well as did observers with
normal hearing up to 15 kHz.
5.6.3.4 Similarity of Auditory  Response to
Microwave and Conventional Acoustic Stimuli
Taylor and Ashleman (1974) measured the electrical
response in three successive levels of the cat auditory
nervous system (eighth  cranial  nerve,  medial
geniculate nucleus, and primary auditory cortex) to
both  acoustic  and pulsed-microwave (2450-MHz)
stimuli. They concluded that the microwave-induced
auditory effect on the animal is exerted similarly to
that of conventional acoustic stimuli.  Furthermore,
these authors reported that inactivation (perforation
of the round window and aspiration of perilymph) of
the cochlea, the known first stage of transduction for
acoustic stimuli, affected the central nervous system
(CNS) response to acoustic  and microwave energy in
the same way; i.e., the evoked electrical activities of
all three sites were abolished by cochlear destrution.
These results indicated that the locus of the initial
interaction of pulse-modulated  microwave energy
with the auditory system might reside peripheral to
the cochlea.
In an experiment in which the thresholds of evoked
electrical responses from the medial-geniculate body
in cats were determined as a function of background
noise, Guy et at. (1975b) found that as the noise level
(50- to 15,000-Hz bandwidth) increased from 60 to 80
dB,  there was only a negligible increase  in the
                       5-72

-------
threshold for the 2450-MHz microwave stimuli, a
moderate increase in the threshold for a piezoelectric
bone-conduction source, and a large increase in the
threshold for  loudspeaker-produced stimuli. The
finding  that the evoked response to microwave
stimuli did  not increase in relation to background
noise, which included acoustic frequencies to 15,000
Hz, indicates that microwaves may interact with the
high-frequency portion of the auditory system.
Guy et al. (1975b) also demonstrated that potentials
evoked by microwave stimuli could be recorded  at
CNS sites other  than those that  correspond to the
auditory nervous system. This finding indicates that
elicited potentials recorded from  any CNS location
could be  misinterpreted  as indicating a  direct
microwave  interaction with the particular system in
which the recording is made.
Prior to 1970, Frey (1962) had suggested that RF
hearing might be a result of direct cortical or neural
stimulation. He  based  this suggestion on (1) his
observations that the perception  of RF pulses was
instantaneous and occurred at low average-incident-
power  densities and on (2) the failure to  record
cochlear microphonics at power densities much
higher than those required to elicit auditory nerve
responses. Cochlear microphonics are electrical
potentials that mimic the sonic waveforms of acoustic
stimuli; they are the signature of mechanical distor-
tion of cochlear  hair  cells, the first stage of sound
transduction. The failure  to observe microwave-
induced  cochlear microphonics had led to the
suggestion  that pulsed microwaves, unlike conven-
tional acoustic stimuli, may not act on any sensor
prior to  acting directly on  the inner-ear  apparatus
(Frey 1967).
In 1975, Chou  et al.  reported  their success  in
overcoming the technical problems  that  had
prevented  investigators from recording cochlear
microphonics from microwave-irradiated animals.
The cochlear microphonics of guinea pigs exposed to
918-MHz (PW) radiation were found to be similar to
those evoked by acoustic stimuli.
The results of the above studies of evoked electrical
potentials in the auditory system,  including the
demonstration of pulsed-microwave-evoked cochlear
microphonics, strongly indicate that the microwave-
induced auditory sensation is detected similarly  to
conventional sound detection and that the site  of
conversion from microwave to  acoustic energy
resides peripheral to the cochlea.  However, it is not
known what structure(s) in the head transduce(s) the
microwave energy to acoustic energy.
5.6.3.5 Mechanism of RF Hearing

As mentioned above, Frey (1967) had suggested that
RF hearing might be a result of direct cortical  or
neural stimulation because of the failure to  record
cochlear microphonics and because the perception of
 RF pulses was instantaneous and occurred at low
 average-incident-power densities. The latter points
 were evidence against a radiation-pressure/bone-
 conduction hypothesis. (See also Guy et al. 1975b.)
 Sommer and von Gierke (1964) had suggested that
 radiation pressure exerted by the RF pulse impinging
 on the surface of the head could launch an acoustic
 signal of sufficient amplitude to be detected by the
 inner ear  via bone  conduction. Other types of
 pressure much greater than radiation pressure can
 be produced in  tissue exposed to RF pulses.  They
 include thermal-expansion  forces,  which are
 proportional to the square of the electric field in the
 material. For example, Gournay (1966) has shown
 that pressures greatly exceeding radiation pressure
 result when visible light from a laser is converted to
 acoustic  energy by  thermal  expansion due to
 absorbed energy in various  liquids.
 Foster and Finch (1974) extended Gournay'sanalysis
 to  a  physiological solution exposed to microwave
 pulses similar to those that produce RF hearing in
 humans. They showed both theoretically and
 experimentally that  radiation-induced pressure
 changes would  result from the absorption of RF
 pulses and could produce significant acoustic energy
 in the solution. In fact,  audible sounds were produced
 by rapid thermal expansion, resulting from only a 5 x
 10~6°C temperature rise in the physiological solution,
 because of the absorption of the energy in the RF
 pulse.

The following  experimental results led  Foster and
Finch (1974) to propose thermoelastic expansion as
the mechanism for RF  hearing.

(1) Acoustic transients were recorded with a
    hydrophone immersed in a  0.15 N KCI solution
    exposed to 2450-MHz pulses that would elicit RF
    sounds in  a human subject. In  addition,  they
    reported the measurement of acoustic transients
    in blood, muscle, and brain exposed in vitro to
    microwave pulses.
(2) The radiation-induced pressure wave generated
    in distilled water  inverted  in phase when the
    water was cooled below 4°C, and the response
    vanished at 4°C, in  agreement with the tempera-
    ture dependence of the thermoelastic properties
    of water.

(3) The thermoelastic theory predicts that the
    maximal pressure in the medium  is proportional
    to the total energy of the pulse for short pulses
    and is proportional to the peak power for  long
    pulses. Foster and  Finch found the relationship
    between pulse width  and the microwave-
    generated acoustic  transient in the KCI solution to
    be consistent with  the theory.

Based on these findings, they concluded that the RF
sounds involve perception, via bone conduction, of
                                                                     5-73

-------
the thermally generated sound transients caused by
the absorption of microwave pulses. The pulses must
be  moderately intense  (typically  500 to 5000
mW/cm2 at the surface of the head). However, they
can be  sufficiently brief  (<  50 //s)  such  that  the
maximum increase in tissue temperature after each
pulse is very small «10"S°C).
A year earlier. Prey and Messenger (1973) published
the results of a study of RF hearing  in four human
subjects. The data in this report are in agreement with
the mechanism proposed by Foster and Finch (1974).
That is,  the loudness of the RF hearing sensation in
the human subjects depended upon the incident-
peak-power density for pulse widths < 30 //s;  for
shorter  pulses, their data show that loudness is a
function of the total energy per pulse.
More recently, Chou and Guy (1979) reported that the
threshold for RF hearing in guinea pigs, as measured
by auditory brainstem-evoked electrical responses, is
related  to the incident energy per pulse for pulse
widths > 30 fjs and is related to the peak power for
longer pulses. The threshold dependence on pulse
width is in agreement with  the predictions  of  the
thermoelastic mechanism as  stated above.
The results on threshold and  loudness  may be
summarized as follows. The energy in the first 30 (js
or so of the pulse determines  the  threshold and
loudness levels regardless of pulse widths greater
than  about 30 /us. For   wider  pulses (> 90 fjs),
loudness is related to peak power rather than energy
because the energy associated with the first 30ps of
the pulse increases directly with peak power. Thus, if
sufficient energy is deposited within a 30-fjs period, an
RF sound will result without  regard to pulse  width.
And, for pulses > 30 fjs, loudness increases with an
increase in peak power. Thus, the auditory response
undergoes  a  gradual  transition  from an energy-
related  effect at pulse widths < 30 /js to an effect
dependent on peak power at pulse widths > 90 //s
(Frey and Messenger 1973; Chou and Guy  1979).

The hypothesis of Foster and Finch (1974) predicts
also  that the RF hearing  effect  is  related to
thermoelastically induced mechanical vibrations in
the head. Vibrations of this type can be produced by
other means,  such as by a laser pulse or by a pulsed
piezoelectric crystal in contact with the skull (Chou et
al. 1976). Frey and Coren (1979) used a holographic
technique to test whether the skull and the tissues of
the head of an animal have the predicted vibrations
when exposed to a pulsed RF field. No displacements
were recorded, but subsequent to this report, Chou et
al. (1980a) demonstrated  that the sensitivity  of  the
holographic technique used by Frey and Coren (1979)
was 3 to 4 orders of magnitude too low to  detect
displacements related to vibrations from microwave-
induced thermoelastic expansion in biological
tissues.
Tyazhelov et al.  (1979b) conducted a series of
psychophysical experiments  with  18 subjects to
evaluate the adequacy of the thermoelastic hypothesis
and to study the perceptual qualities of RF-induced
sounds. Audiofrequency signals were presented
alternately to or concurrently with microwave pulses
(see  Table  5-16)  under  conditions in which  the
subject could adjust the amplitude, frequency, and
phase of the audio signal. The authors concluded that
the thermoelastic hypothesis  adequately explained
some of their findings for microwave pulses of high
peak power  and short  width  (< 50 fjs), but other
results were interpreted as  inconsistent with a
thermoelastic  mechanism for RF  hearing.  For
example, pulse widths  greater than  50 /us, which
increased the mean power level, produced increases
in loudness that rose more rapidly than predicted by
the thermoacoustic model. In addition, suppression of
RF sounds by the audio signal was reported to be
inconsistent with the model.
Lebovitz and Seaman (1977) compared the response
to pulsed microwaves (915 MHz) with the response to
acoustic clicks by monitoring single auditory neurons
in the  cat. A response  of these neurons to pulsed
microwaves was  predicted  by earlier studies that
demonstrated subjective  auditory perception (Frey
1962), auditory  evoked potentials (Taylor and
Ashleman 1974), and cochlear microphonics (Chou
et al.  1975). Furthermore, the thermoelastic model
predicts that a  mechanical wave of pressure
stimulates the inner ear via bone conduction. Thus, the
response of the neurons in the auditory pathway to
pulsed microwaves should  be similar to their
response to  transient mechanical stimuli such as
acoustic clicks. The results indicated that mechanical
factors within the cochlea are similarly involved in
determining  both the  acoustic and the  pulsed
microwave response (Lebovitz  and Seaman  1977).
Other data in this report (Lebovitz and Seaman 1977)
appeared to be  inconsistent with the thermoelastic
model  that predicts  a  high-frequency component
such as the microwave-induced cochlear microphonic
recorded by Chou  et al.  (1975). That is, Lebovitz  and
Seaman (1977)  observed a decrease in sensitivity of
high-frequency auditory units  to microwave pulses.
However, they used long pulses of 250 to 300 //s in
duration to obtain maximal energy per pulse. More
recently, Tyazhelov et al. (1979b) reported that long
pulses (~ 100 //s) result in a  lower pitch of the RF
sound in humans. Two of their observers who had a
high-frequency auditory limit of 10 kHz could not hear
short RF pulses but could hear  long pulses. Thus, the
results of single unit recordings in cats are consistent
with human perception  of RF pulses when the pulse
widths are long.

Lin (1977) developed  a theoretical model that
estimates the characteristics of acoustic  signals
induced in laboratory animals and  humans by
                       5-74

-------
microwave pulses; his  model is based on thermal
expansion in spherical  heads irradiated by pulsed
microwave energy. The frequency  of the induced
sound was found to be a function of head size and of
acoustic  properties  for  brain tissues; hence, the
acoustic pitch perceived by a given subject is the
same regardless of the RF frequency. The calculations
of Lin show that thefundamental frequency predicted
by the  model varies inversely with the radius of the
head;  i.e., the larger the radius,  the lower the
frequency of the perceived RF sound. He estimated
the fundamental frequency of vibration in guinea
pigs, cats, and adult humans to be 45,38, and 13 kHz,
respectively. The frequency for an infant head was
estimated  to  be  about  18 kHz. These calculations
provide further evidence that a necessary condition
for auditory perception by adult humans is the ability
to hear sound above 5 to 8 kHz (Frey 1961; Rissmann
and Cain 1975).
The  results of Lin  (1977)  appear  to  be  in  good
agreement with the measurements of Chou et al.
(1975), who found cochlear microphonics in guinea
pigs to  be 50 kHz. In a later report, Chou et al. (1977)
found the  frequency of the cochlear microphonic in
guinea  pigs and cats to correlate well with the longest
dimension of the brain cavity. Extrapolation of these
results  indicates that the frequency of the microwave-
induced cochlear microphonic  in  human beings
should  be  between 7 and 10 kHz. Chou et al. (1977)
concluded also that the frequency of the cochlear
microphonic was independent of microwave fre-
quency (915  or 2450 MHz) and exposure method
(horn applicator or cylindrical waveguide).
In summary,  evidence  from many studies—the
measurement of acoustic transients in water, KCI
solution, and tissues (Foster and Finch 1974) and in
muscle-simulating materials  (Olsen and Hammer
1980); the  relationship of pulse duration and
threshold (Foster  and  Finch 1974;  Frey and
Messenger  1973; Chou and Guy 1979); the
characteristics of the field-induced  cochear micro-
phonics in laboratory animals (Chou et al.  1975,
1977) and the theoretical calculations (Lin 1978)—
indicates  that thermoelastic expansion is the
mechanism that explains most of the characteristics
of the RF hearing phenomenon.

5.6.4 Unresolved Issues
Several investigators have tried to determine the
thresholds for the RF-induced auditory sensation in
human beings and in laboratory animals (Table 5-17).
Human studies and animal experiments, in general,
are few and have used small sample sizes, different
frequencies, and different experimental procedures.
However, the  radiation parameters that are  most
important in determining the threshold for the
auditory response  in humans (discussed in Sec.
5.6.3.2, Effective Radiation  Parameters) and  in
laboratory animals (discussed below) are being
characterized.

The  threshold for an  auditory response in cats
exposed to 918- and 2450-MHz (PW) radiation was
studied by recording of electric potentials  from the
medial geniculate body (Guy et al.  1975b). As the
pulse width increased from 0.5 to 32/ys, the threshold
value for the peak of power density decreased
proportionately for pulse widths below 10 /us. The
thresholds of average  power density and energy
density per pulse also increased with pulse width, but
these parameters did not show the strong proportiona I
relationship with pulse width as did the peak of power
density.  Guy et al.  (1975b) concluded  that the
threshold for the  evoked  auditory response was
related to the incident energy density per  pulse,  at
least for pulse widths less than 10/us. In the cat, the
threshold energy density per pulse was found to be
about one-half of that  which  had produced a
sensation in one human subject.

In a  similar experiment conducted at frequencies
between  8670 and 9160 MHz, Guy et al. (1975b)
found that the threshold values of power density and
of energy density per pulse, which  include the
auditory response of cats, were an order of magnitude
higher than those  required at 918  and 2450 MHz
(Table 5-17). Furthermore, no auditory response was
obtained  at 8670 to 9160 MHz until the brain was
exposed by enlarging the hole in the skull that served
as the electrode access port.

Cain and Rissmann (19.78) determined the threshold
for auditory responses in animals exposed  to 3000-
MHz (PW) radiation. Although their results are  in
general agreement with Guy et al. (1975b) in that the
threshold energy density per pulse was relatively
constant  for pulse widths of 5, 10, and 15 /us, their
results are confounded by the use of scalp electrodes
and only few animals of three species (two  cats, two
chinchillas, and a dog).
In the following year,  Chou and  Guy (1979)
determined the threshold for RF hearing in guinea
pigs by  measuring  auditory brainstem-evoked
electrical responses  with carbon-loaded Teflon
electrodes attached to the skin of the head. In this
experiment, the head of the guinea pig was inserted
into a 918-MHz circular waveguide exposure system.
The  threshold  dependence on pulse width was
found to be in agreement with the predictions of the
thermoelastic expansion  mechanism;  that is, the
threshold was related to the incident energy per pulse
for short pu Ise widths « 30 fjs) and was related to the
peak power for longer pulses. At the shortest pulse
width (10/us), the threshold absorbed energy density
of RF hearing in the guinea pig was about  6 /uJ/g.
Wilson et al. (1980) described an autoradiographic
technique in which [14C]2-deoxy-D-glucose was used
                                                                    5-75

-------
Table 5-17.   Summary of Studies Concerning Threshold Values for Auditory-Evoked Potentials in Laboratory Animals
                                                  Exposure Conditions
Effect
Response obtained
with scalp
electrodes
Response obtained
from round window
with carbon lead
Response obtained
with carbon-loaded
Teflon electrodes
Response obtained
from medial
geniculate with
glass electrode
Response obtained
from individual
auditory neurons
with glass electrode
Species
(n)
Cat (2)


Guinea
Pig (5)

Guinea
pig (not
given)
Cat (2)



Cat
(not
given)

Frequency
(MHz)
3000


918


918


918
2450
8.670-9.160

915



Pulse
Repetition
Rate
(s-)
0.5


100


30


1
1
1

<10



Pulse Width
Uis)
5
10
15
•1-10


10-500


3-32
0.5-32
32

25-250



Pulse
Intensity
(mW/cmJ)
2.200, 2.800
1.300
580
•


62-156


800-5.800
600-35.600
14.800-38.800

—



Average
Intensity
(mW/cm'l



*


0.05-1.4


0.017-0.028
0.015-0.047
0.472-1.24

S1.0



Incident
Energy Density
Per Pulse
11. 14
13
87



1.56-468


17.4-28.3
15.2-47.0
472-1.240

	



Peak Absorbed
Energy Density
Per Pulse



20


6-180


12.3-20.0
8.7-26,7


4-40



References
Cam and Rissmann ( 1 978)


Chou era/ (1975)


Chou and Guy (1979)


Guv era/ (1975b)



Lebovitz and Seaman
(1977)


•Direct comparison of power density in the waveguide exposure system to free-field power density is improper because the efficiency of energy coupling is 10 times higher than for free-
 field exposure. (See Chou el »l. 1975. p. 362.)
to map auditory activity in the brain of rats exposed to
acoustic  stimuli and to PW and CW microwave
radiation. With this technique, in vivo determination
of metabolic activity (i.e.,  glucose utilization and
associated  functional activity in the brain) can be
visualized. Prior to exposure to the acoustic stimuli
or to microwaves, one middle ear was ablated. First,
the authors showed the expected bilateral asymmetry
of radioactive tracer uptake in the auditory system of
rats exposed to acoustic clicks or weak background
noise.  Second, in contrast, a symmetrical uptake of
tracer was  found  in  animals  exposed to PW
radiation.  Thus,  the  autoradiographic  results
confirmed  the finding  that RF  hearing does not
involve the  middle ear (Frey  1961;  Chou and
Galambos 1979). Unexpectedly, Wilson et al. (1980)
found similar patterns of radioactive tracer uptake in
the auditory system of rats exposed to CW radiation
(918 MHz; 2.5 and 10 mW/cm2) and to PW radiation
(2450 MHz).

This result, indicating an auditory response to CW
microwaves,  was unexpected, because no report of
a direct hearing sensation due to exposure to CW
microwaves had appeared in the literature. Since the
thermoelastic hypothesis of RF hearing is based on
the properties of PW radiation, this observation
suggests that another mechanism may be involved
in the interaction of RF radiation with the auditory
system of the brain.

It is well documented that some human beings can
hear pulsed RF radiation as a buzz, click, or knock.
Furthermore,  the thermoelastic  expansion mecha-
nism  explains how the pulse of RF  energy can  be
transformed  to an acoustic  impulse in the  head.
Since  a single pulse of RF radiation can be heard,
calculation of the threshold energy in terms of the
average-incident-power density results in a very
small number.  Because a very low average-power
density can cause an acoustic response in the head
and there is the potential for exposure of the public
to pulsed fields that induce the effect, an unresolved
issup is the need to assess the psychological effects
of RF hearing, particularly in populations that may
have no knowledge as to the origin of the RF sounds
in the head.
5.6.5 Human Cutaneous Percept/on
Exposure of the human body to microwave radiation
can  cause heating  that  is detectable by  the
temperature-sensitive receptors in  the skin. As
shown in Table 5-18, several  investigators have
experimentally determined the microwave intensities
that cause sensations of warmth and thermal pain in
human subjects.
5.6.5.1 Frequency Specificity
Hendler  and colleagues (1963, 1968) exposed  a
circular area (37 cm2) of the forehead to 3- and 10-
GHz (PW) radiation and to infrared (IR) radiation. The
forehead was  selected for a study of warmth
sensations because the temperature receptors in the
skin of the forehead are relatively numerous and are
evenly distributed, so that the area constitutes a low-
threshold region of uniform  temperature sensitivity.
The  lower-frequency microwaves (3  GHz)  had  a
higher-intensity threshold for warmth sensation. The
higher-frequency  IR radiation  was more effective
than  either microwave frequency, because  the IR
energy was absorbed more effectively in the outer
skin layers containing the thermal sensors.

The study by Justesen et al. (1982) confirmed the
earlier work on  perception of microwave radiation.
Four human adults individually exposed the ventral
surface of the  forearm for  10  s to 2450-MHz or
infrared radiation; two additional subjects experienced
                        5-76

-------
Table 5-18.   Summary of Studies Concerning Human Cutaneous Perception of RF-Radiation
Effect
Sensation of
warmth on
forehead
Sensation of
warmth on
forehead
Sensation of
warmth on
inner forearm
Sensation of
warmth on
inner forearm
Sensation of
pain on inner
forearm
Frequency
(MHz)
3,000
(PW)
10,000
(PW)
2,880
2450 (CW)
3,000
(PW)
3,000
Intensity
(mW/cm*)
—
74
56
26.7
300-2,500
2,500
1,000
Durtion
(s)
1-5
1-5
15-73
50-180
10
<6
30
130
SAP*
(W/kg)
20-40
25-35
—
^~
"
2,000
Reference
Handler (1968)
Hendler et al.
(1963)
Schwan et al. (1966)
Justesen et al. (1982)
Vendrick and Vos
(1958)
Cook (1952)
*SAR value is estimated.
microwave radiation only. Thresholds of detection of
just-noticeable warming were determined to be 26.7
mW/cm2 for microwave radiation and 1.7 mW/cm2
for  infrared waves. The 15-fold difference between
thresholds was attributed to (1) differential scatter
(nearly two-thirds of the incident microwave energy
is scattered and not absorbed), and (2) the differing
depths of penetration of the two forms of energy (a
relatively  small  proportion of the total absorbed
microwave energy is absorbed at the skin's surface).
The total energy absorbed at threshold values was
estimated  to be  10.2 joules (9.5 mJ/cm2) and  1.8
joules  (1.7 mJ/cm2) for microwave and infrared
radiation, respectively. Although absorption of
microwave energy was more than five times greater
at threshold,  none  of  the  four  observers who
experienced both types of radiation could distinguish
a difference  in  sensory quality. The authors
concluded that the same set of superficial  thermore-
ceptors was being stimulated, only less efficiently so,
by the more deeply penetrating,  more diffusely
absorbed microwave energy.

5.6.5.2 Temperature Thresholds
For both microwave and IR radiation at intensities
producing warmth sensations, a threshold of warmth
was experienced when the temperature  of a more
superficial layer of subcutaneous tissue ~ 0.2 mm
below  the skin's surface was increased  ~ 0.01  to
0.02°C over the temperature of a deeper layer in the
skin lying ~ 1 mm below the surface. In this study it
was also noted that there was a persistent
sensation of warmth for ~ 7 s after cessation of the
exposure, which indicated the continued existence of
an  effective temperature difference  between the
subcutaneous tissue layers (Hendler 1968).
Schwan et al.  (1966) exposed a small area of the
forehead (7-cm diameter) equivalent to the area
exposed in Hendler's studies to 2.88-GHz radiation
and measured the length of time that elapsed before
the person was aware of a sensation of warmth. The
times for four subjects varied from 15 to 73 s at 74
mW/cm2 and from 50 to 180 s at 56 mW/cm2. The
authors found that the  reaction  times were not
linearly proportional to the reciprocal of the incident
power density  and  concluded  that  subjective
awareness of warmth was not a reliable indication of
microwave hazard.
Vendrik and Vos (1958) exposed a 13-cm2 area of the
inner forearm to  3-GHz (PW) radiation (300 to 2500
mW/cm2) and found the threshold for temperature
changes to be  0.4 to  1.0°C. Skin temperature
increases that were kept below 1 °C were linear with
microwave intensity for six exposure durations.  In
contrast to the regularity of skin temperature changes
induced by microwaves, the reports of temperature
sensations were variable. Sensations  of warmth
occurred < 0.5  to 3.5 s after rapid rises in skin
temperature. The sensations did not cease when the
skin  temperature  began  to  drop. In this  study,
microwave radiation (3 GHz) was found to be a factor of
10 less effective in producing a temperature elevation
than was IR radiation at a similar intensity.


Cook  (1952) determined the pain threshold in six
subjects who were exposed to 3-GHz radiation at five
different sites on the body's surface. The initial skin
temperature  ranged  from  31.5  to 33.5°C.  Pain
resulted when a  critical skin temperature  (~ 46°C)
was reached rather than from  a critical temperature
rise (AT). For an  inner forearm area of 9.5 cm2, the
power density pain thresholds varied  from  2500
mW/cm2 for a 30-s exposure to 1000 mW/cm2 for an
exposure of 130 s. The pain threshold was lower for a
larger exposed area of 53 cm2.  The skin temperature
corresponding  to burning pain was found to be
                                                                     5-77

-------
independent of the area of exposure, radiation
intensity, exposure time, and anatomical site. At high
intensities, the exposure time needed to produce pain
was an inverse function of radiation  intensity. The
author reported that the sensations of warmth and
pain with microwave heating differed little from those
resulting from IR. radiation.
5.5.5 Unresolved Issues
The few  studies on  thermal  pain  and  warmth
sensations in human beings exposed to frequencies
in the range of 3  to 10 GHz  provide useful data on
exposure levels that are clearly  undesirable for the
general population. Cutaneous perception, however,
may be a reliable indicator of an  unsafe exposure
level only at RF frequencies with wavelengths small
in comparison to the length of the exposed body, i.e.,
wavelengths  comparable to or smaller than the
thickness of skin. Under these conditions, most of the
energy  is absorbed in the outer tissue layers
containing thermal  sensors.  At  lower frequencies,
which have wavelengths approximately equal to  or
longer than the human body, modeling studies have
shown that much of the energy is absorbed within the
body belowthe superficial skin layers. In all mammals
tested, the threshold temperature (~42°C) of cellular
injury for  sustained elevations (seconds to tens  of
seconds) is below the  threshold (~ 45°C) of  pain
(Hardy et al.  1967). These results strongly indicate
that cutaneous perception of RF  energy is not a
reliable sensory  response that protects  against
potentially harmful  levels of RF radiation over the
broad frequency range of 0.5 MHz to 100 GHz.
                       5-78

-------
5.7 Endocrine, Physiological and
Biochemical Effects
Charles G. Liddle
Carl F. Blackman

5.7.1 Endocrine Effects
The endocrine system  in coordination with the
nervous system is a major regulatory system in the
body. Alterations in the function of the neuroendocrine
system often reflect the efforts on the part of the
animal to maintain homeostasis when  subjected to
stressful  internal and external stimuli.  Detection of
changes in the endocrine system is a sensitive means
of analyzing the animal's  responses either to direct
stimulation of the endocrine organs themselves or to
stimulation of the  CNS.

Animals exposed to a wide variety of stimuli generally
respond with a rather specific pattern of physiologic
changes,  usually referred to as the general
adaptation syndrome, and  the stimuli that can
provoke the syndrome are usually called stressors.
An increase in the concentration of corticosteriods in
the blood above that which would normally occur at
that time of day  in the absence of a stimulus is
considered by many to be an indicator of an animal's
response to stress. Such an increase results when an
internal  or external stimulus, either chemical,
physical,  or  emotional,  excites neurons of the
hypothalamus to release corticotropin-releasing
hormone, which  drives  the pituitary to release
adrenocorticotropic hormone  (ACTH). This hormone
then  stimulates  the adrenal cortex to secrete
corticosteroids. Among  the strongest stressful
stimuli are surgery,  anesthesia, cold, narcosis,
burning, high environmental temperature, and
rough handling or  restraint.

The thyroid gland plays a principal role in regulating
basal metabolism,  as well as in generating metabolic
heat in the tissues. Changes in thyroid activity can
result from changes in thyroid-stimulating hormone
from the hypothalamic-hypophyseal system and/or
increased metabolic activity of the thyroid gland due
to heating. Direct interaction with the CNS could also
produce changes  in  thyroid  activity.  Moderate or
gradual  heating results  in a reduction of  thyroid
hormone;  rapid  or  marked  elevation of  body
temperature results in an increase in thyroid activity.
The effects of RF radiation on the endocrine system
are discussed below and summarized in Table 5-19.

Thyroid function of rats following exposure to 2450-
MHz (CW) microwaves at 10, 20, and 25 mW/cm2
(SARs estimated at 0.25 W/kg per mW/cm2)for 4 or
16 h was studied by Parker (1973). No effects on
thyroid gland function were found at these exposures.
However, at 15 mW/cm2 (SAR estimated at 3.75
W/kg), exposure  for  a longer period  (60 h) was
reported to produce a significant decrease in serum-
protein-bound iodine, thyroxine, and thyroid/serum
iodine ratio. A significant rectal temperature increase
(1.7°C) was reported at 25 mW/cm2, but not at lower
power densities.
Magin et al.  (1977a,b) irradiated the surgically
exposed  thyroid gland  of  anesthetized dogs with
2450-MHz (CW) microwaves  using a waveguide
applicator at  power densities of 72, 162,  and 236
mW/cm2 (SARs from 58 to 190 W/kg) for 2 h. One
thyroid was irradiated while the other was used as a
control. Tissue temperatures of 39, 41, and 45°C
were maintained in the thyroid at the three power
densities. They reported an increased release  of
thyroxine (TH) and triiodothyronine (T3) at all power
densities, which showed that the thyroid gland in the
dog can be stimulated directly by microwave heating.

Milroy and Michaelson (1972) exposed rats to 2450-
MHz microwaves at 1, 10, and 100 mW/cm2 (SAR =
0.25 W/kg per mW/cm2) for single exposures of 10,
20,30, and 45 min and at 1 and 10 mW/cm2,8 h/day,
for 8 weeks,  and reported no  effect on T3  levels,
thyroxine  levels, or on the  uptake of radioactive
iodine. No rectal temperature increase was observed
at 10 mW/cm2 or less. At 100 mW/cm2 there was a
constant  rise in  rectal temperature  throughout
exposure,  up  to 42°C at the end  of the exposure
period. Histopathological examination of the thyroid
glands also showed no effect from  the exposure.

An  increased production  of thyroid hormone  in
rabbits as measured by increased incorporation of1311
and  increased radioactivity per gram of thyroid
(verified by  autoradiography) was reported by
Barahski et al. (1972). The animals were exposed for
3 h/day  for  4 months to 10-cm (3-GHz, PW)
microwaves at an average incident power density of 5
mW/cm2  (SAR estimated  at 0.25 to 0.75 W/kg).
They  reported no increase in body temperature or
thyroid temperature. (The pulse parameters were not
given, so peak power density cannot be calculated.)
Lu et al.  (1977) reported serum thyroxine  levels  in
rats exposed to 2450-MHz microwaves at 1, 5, 10,
and 20 mW/cm2 (0.25, 1.25, 2.5, and 5 W/kg) for 1,
2, 4, or 8 h. Decreased thyroxine levels were observed
at 20 mW/cm2 following 4- and 8-h exposures. The
thyroxine  values at shorter  exposures  and lower
power densities were not significantly different from
the sham values, except for an increase after 4 h at 1
mW/cm2, which appears to be a chance variation,
since at both higher and lower power densities and at
longer and shorter periods of exposure no effect was
detected. There were small but statistically significant
rectal temperature increases at  1 mW/cm2 after 4 h,
at 5 mW/cm2  after 1 and2h,andat 10 mW/cm2 after
2 and 4 h of  exposure. The increases were in the
range of 0.2  to 0.56°C.  The  lack  of correlation
between  power density, exposure  time, and rectal
temperature increase, along with  the small rectal
                                                                     5-75

-------
temperature change, suggests that the effects may
not  have  been due  entirely  to the  microwave
exposure but possibly from the stress of confinement.
At 20 mW/cm2, however, the rectal temperature was
significantly elevated at all four exposure times and
tended to increase with longer exposures. At 1  and 2
h the  changes were small (0.64 and 0.54°C), but
larger (1.01 and 1.35°C) increases occurred after 4
and 8 h. Serum corticosteroid levels were decreased
at 20  mW/cm2 after 8  h only, which the authors
report as a shift in the circadian periodicity. Serum
growth hormone measurements showed no change
at any of the power densities or times reported. They
also measured no change in the mass  of the thyroid,
pituitary, and  adrenal glands following irradiation.

In another study on the effects of microwaves on
acute endocrine responses in rats, Lu et al. (1981)
exposed animals to 2450-MHz  microwaves (AM  at
120 Hz) for  1  and 4 h, and measured colonic
           temperature, thyrotropin and corticosterone levels
           immediately after exposure. The 1 -h exposures were
           conducted at power densities of 1, 5,  10, 20, 40, 50,
           60, or 70 mW/cm2, and the 4-h exposures at 0.1,1,5.
           10, 20, 25, and 40 mW/cm2 (measured SAP = 0.21
           W/kg per mW/cm2). The results for the 1 -h exposure
           showed an increase in colonic temperature with
           increasing power density with a significant increase
           at 20 mW/cm2 and above, but not at 10 mW/cm2 or
           below. Corticosterone also showed an increase with
           increased power density.with evidence of a threshold
           between  20 and 40  mW/cm2. Thyrotropin values
           showed a decrease with increasing power densities,
           with significant results at  40 mW/cm2 and above,
           and  equivocal  results in the  10- to 20-mW/cm2
           range. (The 10-mW/cm2 values were significantly
           lower; those at 20 mW/cm2 were not.) Similartrends
           were observed following  4-h  exposures,  with  a
           significant increase in colonic temperatures at 10
           mW/cm2  and above, and no increases at 5 mW/cm2
Table 5-19.    Summary of Studies Concerning RF-Radiation Effect* on Endocrinology
                                               Exposure Conditions
Effects
                       Species
                               Frequency
                               (MHz)
Intensity
(mW/cm*)
Duration
(days x min)
                                                                SAR (W/kg)
                                                                            AT" (°C)
                                                                                           Reference
Increased thyroxine and         Dog      2450 (CW)
triiodothyronine

No effect on thyroid gland or      Rat      2450 (CW)
thyroid hormone

No effect on thyroid function      Rat      2450 (CW)
Decrease in serum-protein-bound   Rat      2450 (CW)
iodrne. thyroxine and thyroxine/
serum ratio

Increase in thyroid hormone      Rabbit     3000 (PW)
Decrease in serum thyroxine      Rat      2450 (CW)
levels

Increase in corticosterone        Rat      2450 (120 Hz
levels                            AM)

Decrease in thyrotropin levels


Increase in corticosterone levels


Decrease in thyrotropin levels


No effect on thyroid, pituitary.     Rat      2450 (CW)
or adrenal glands weight or
growth hormone levels
                                            72-236
1. 10. 100
1. 10

10. 20. 25
                                                      1 x 120
1 x 10-45
56 x 480

1 x 240-960
                                                      25 x 1440
                                                                58-190
0.25-25
0.25-2.5

2.5, 5.
6.5 lest)

3.8 lest)
2-8
(thyroid temp)

£ 10 mW/cm3 = 0
lOOmW/cm'fiS

§ 20 mW/cm' = 0
25 mW/cm! = 17
5
20
Negt 1-10
40-70
Neg 1-20
10. 4O-70
Neg 1-5. 20
10-40
Neg 1-5
25. 40
Neg 1-20
1-20
48 x 180
1 x 240-480
1 x 60-480
1 x60

1 x240
5.2, 84
002-42
1 x 60-480
0.25-0.75
lest)
5
025-25
84-14
Neg 021 -4.2
2.1-14.7
Neg 0.21 -4.2
2.1-8.4
Neg 0.2- 1.0
06-2.1
0-1
0.25-25
Not reported
0-0.6
0.6-1 4
1.3-3.0
0-0.6
0-3.0
0-0.6
03-2.1
0

0-1-4
Maginer*/. (1977a.b)


Milroy and Michaelson
(1972)

Parker (1973)


Parker (1973)



Baranski et al
(1972)

Luetal. (1977)


Luet al (1981)
                                                Luetal (1977)
No effect on thyroid, anterior
pituitary gland, adrenal, prostate
or testes weights; no change in
follicle-stimulating hormone or
gonadotropic hormone levels
Increase in leutmizing hormone
Increased adrenal weights and
significant adrenal response
Increased plasma corticosterone
levels




Increase in corticosterone

No effect on serum corticosterone
levels
Rat




Rat
Infant
Rat
Rat





Rat

Rat

2860-2880 ICW)




2860-2880 (CW)
2450 (CW)

2450 (CW)





2450 (AM.
120 Hz)
918 (CW)

10




10
40

50. 60
Neg 13-40

20-40
Neg 13

50. 60

25

36




36
6x

1 x


1 x


1 x

91

x360




x360
5

30-60


120


60

X600

1 -2 lest)




1 -2 (est)
20-60 lest)

11 5-13.8 lest)
Neg 3.0-9.2

4.6-9.2 es
Neg 13 lest)

8.3. 9.6

1.0

Not reported




Not reported
1 5-2.5

13 mW/cmJ
20 mW/cm!
30 mW/cm'
40 mW/cmJ
50 mW/cm'
60 mW/cm'
Not reported

0









= 0.5
= 0.7
= 09-1.4
= 1.3-1.4
= 1 6-2.4
= 2.5-2.9




Mikolajczyk<1976)




Mikolajczyk (1976)
Guillen and
Michaelson (1977)
Lotz and Michaelson
(1978)




Lou and Michaelson
(1979)
Lovely •(»/. (1977)

•AT = Rectal temperature mcreii
tNeg = Effect not found at value indicated
                         5-50

-------
or below with the exception of one small group at 1
mW/cm2. The trends for the increase in corticosterone
levels and the decrease in thyrotropin levels were not
significant; however, the corticosterone level at 40
mW/cm2 was significantly increased, but not at 25
mW/cm2 or  lower values. Also, the thyrotropin
values  at  25 mW/cm2 and 40 mW/cm2 were
significantly decreased,  whereas those  at  20
mW/cm2 and below were not. Comparing the effects
of 1 - and 4-h exposures on the three parameters, the
change in colonic temperature was similar for both
exposures; however, the stimulatory effect on serum
corticosterone levels was less for the 4-h exposures
than for the 1-h exposures, but the depression of
serum thyrotropin was more pronounced after 4 h
than  after  1  h. The  authors concluded  that the
hormonal changes probably represented a general
nonspecific stress reaction that  was related  to the
intensity and duration of the stressing agent, rather
than to the nature of the agent itself.

No change in  thyroid weight was seen in a study by
Mikolajczyk (1976), where  rats were  exposed to
2860-  to 2880-MHz microwaves at 10  mW/cm2, 6
h/day,  6 days/week for  6 weeks. The SAR is
estimated to be  1  to  2  W/kg for a  single animal
exposure, but the animals were exposed close
together in a box with dividers every 10 cm,  which
should give a somewhat higher SAR  value. The
author did find a significant increase  in leuteinizing
hormone from the anterior pituitary gland but  no
chance in follicle-stimulating or  gonadotropic
hormone levels. The weights of the anterior pituitary,
adrenal, prostate, or testes were  not affected  by the
exposure.

Lovely  et al.  (1977) exposed  rats to 918-MHz
microwaves at 2.5 mW/cm2 (SAR ~ 1.04 W/kg), 10
h/day  for  13 weeks, and  no changes in serum
corticosterone levels were  observed. In a  study by
Guillet and Michaelson (1977),  rat  pups were
exposed to 2450-MHz microwaves at 40 mW/cm2
(SAR estimated at 20 to 60 W/kg), 5  min/day for 6
days beginning at 1 day postpartum; no change in
basal corticosterone levels was found. There was a
significant adrenal response to the microwaves, but
this response was the same as seen following ACTH
administration,  which  would  suggest a stress
reaction. Rectal temperatures of the exposed animals
averaged 1.5 to 2.5°C higher than those of the sham-
exposed animals. Adrenal  mass of  the irradiated
animals was significantly greater than those  of the
control animals.

In another study, Lotz and  Michaelson (1978)
irradiated rats with 2450-MHz microwaves at power
densities of 13, 20, 30, and 40 mW/cm2 for 30,60, or
120 min and at 50 and 60 mW/cm2 for 30 or 60 min
(SAR estimated at 0.21 W/kg per mW/cm2). Plasma
corticosterone levels were  increased at power
densities at or above  50 mW/cm2, but not  at 40
 mW/cm2 or less, for the 30- and 60-min exposures.
 At the longest exposure time (120 min), increased
 levels were seen at or above 20 mW/cm2 but not at
 13 mW/cm2.  Graphs were presented of the rectal
 temperature taken at the completion of exposures;
 estimates of the rectal temperature increases taken
 from the graphs for 50 mW/cm2 were 1.6 and 2.4°C
 and for 60 mW/cm2 were 2.5 and 2.9°C for 30 and 60
 min,  respectively. The temperature increases for 40
 mW/cm2 were 1.3 and 1.4°C for 30 and 60 min. For
 the 120-min exposures, rectal temperature increased
 0.5°C at 13 mW/cm2 and 0.7°C at 20 mW/cm2.

 A study by Lotz and Michaelson (1979) suggests that
 the change  in corticosterone levels is  due to
 stimulation of the pituitary gland, probably due to
 hyperthermia. They exposed hypophysectomized rats
 to  2.45-GHz  microwaves  (AM at  120 Hz) at 60
 mW/cm2 (9.6 W/kg) and found significantly lower
 levels of corticosterone compared to normal and
 sham-hypophysectomized  rats. In another study,
 dexamethasone effectively  suppressed  the cortico-
 sterone response in rats exposed to 50 mW/cm2 (8.3
 W/kg), but only partial suppression was observed at
 70 mW/cm2  (11.2  W/kg). These results  provide
 further evidence for a stimulatory effect of microwave
 radjation on  the pituitary  gland  rather than the
 adrenal gland.
 In summary,  microwave effects on thyroid  function
 have been reported at SARs as low as 2.1 W/kg, and
 negative results have been reported at values as high
 as 25 W/kg. The duration of exposure, as well as the
 exposure rate, appears to be important, as demon-
 strated  by the  study  by  Parker (1973) where
 exposures  at  6.25 W/kg  for 16 h produced no
 changes, whereas 3.75 W/kg exposures  for 60 h
 resulted in a  decrease  in  serum  thyroxine levels.
 Changes in corticosterone levels have been  reported
 from microwave exposure at SARs as low as 10 W/kg
 but not at 6.25 W/kg, and  adrenal responses  have
 been reported at levels as low as 4.6 W/kg but not at 3
 W/kg.

 5.7.2 Clinical Chemistry and Metabolism
 An individual's response to many stresses manifests
 itself through  changes in  some of the  clinical
 chemistry indices. Serum calcium and phosphate
 levels normally increase initially and then decrease
 below normal in response to stress; whereas serum
 glucose,  blood  urea  nitrogen, and uric acid levels
 increase following stress.  These blood chemical
 responses are consistent with induction of release of
 adrenocortical hormones  in response to stress.
 Physiological responses, including thermal responses,
 to RF-radiation exposure may occur at exposure
 levels too low to produce large changes in thermo-
tregulatory behavior or colonic temperature. In such
 instances there may be changes in various metabolic
 parameters. There are few clinical chemistry and
                                                                     5-81

-------
metabolism reports where exposures were sufficiently
defined to relate results to the power density or SAR.
These studies are described below, and are outlined
in Table 5-20.
A study  of the  effects of microwaves on serum
chemistry values was reported by Lovely et al. (1977),
who exposed rats in circularly polarized waveguides
to 918-MHz (CW)  radiation. The animals  were
irradiated 10 h/day for 13 weeks at a power density of
2.5 mW/cm2 (SAR ~ 1.04 W/kg). They reported no
change in Na+, K*, ion gap,  Cf, blood urea nitrogen, or
glucose values compared to the sham-irradiated
animals. They did report a significant difference in the
serum  calcium values at the end of the 12-week
exposure period. However, this is most probably a
spurious  result, because the calcium  levels in the
sham-irradiated animals were  decreased  from
previous values, whereas the levels in the irradiated
animals remained unchanged from earlier values.
These results, therefore, are interpreted to mean that
918-MHz microwaves at 2.5 mW/cm2 do not alter the
measured clinical chemistry values. Colonic temper-
ature measurements were made on the animals, and
no detectable changes were observed.

Somewhat  differing from these results, Wangemann
and  Cleary (1976) reported an increase in serum
glucose levels in rabbits exposed for 2 h to 2450-MHz
microwaves (CW and PW) at power densities of 5,10,
and 25 mW/cm2 (SARs estimated at 0.8,1.6, and4.0
W/kg, respectively), and  an increased blood  urea
nitrogen  value in animals exposed to 25-mW/cm2
microwaves only. Increased uric acid  levels  were
found at  10 and 25 mW/cm2, but not at 5 mW/cm2
(both CW and PW). Levels of calcium, phosphorus,
cholesterol, total protein, alkaline phosphatase, lactic
dehydrogenase, and serum glutamic  oxalic trans-
aminase were unaffected at the three power densities.
The authors stated that the PW and CW results could
not be compared directly because the exposure
conditions were different. The results were those that
would be expected from heat stress. Animals exposed
at 25 mW/cm2 for 2 h showed a significant rectal
temperature increase of 1.7 and 2.9°C for PW and
CW  exposures,  respectively. Those exposed at  10
mW/cm2 showed evidence of mild heat stress,  such
as peripheral vasodilation, but no significant increase
in rectal temperature. One possible explanation for
the difference  between these results and those of
Lovely  et al. is the  difference in absorbed energy
patterns in  rats and rabbits at 2450 MHz.

Brains from rats exposed at 1600 MHz for 10 min at a
power density of 80 mW/cm2 (SAR estimated at 48
W/kg)  were analyzed for  selected heavy metals by
Chamness et al. (1976). Iron levels were increased in
all the areas of the brain (hypothalamus, corpus
striatum, midbrain, hippocampus, cerebellum,
medulla,  cortex).  Manganese was increased in the
cortex and medulla, and copper was increased in the
cortex; while calcium, zinc, sodium, and potassium
levels were unchanged. The observed changes were
probably a result of  hyperthermia,  since  most
alterations seen were also observed in rats subjected
to a hyperthermal environment, and the irradiated
animals showed a rectal temperature increase  of
4.5°C.

Platelet-rich human plasma was irradiated in vitro
with 2450-MHz microwaves at power densities from
10 to 280 mW/cm2 (SARs 1.3 to 38 W/kg) for 0.5 to
24 h by  Boggs et al.  (1972),  and effects on blood
coagulation were  analyzed. They  reported  no
significant changes in platelet  count, coagulation
time, or clot strength at these power densities. The
plasma temperature  was maintained  below  37°C
during the exposures. They also conducted studies on
the effect of heating on coagulation time and clot
strength. Samples were heated either by microwaves
or by radiant heating, and the results were compared
to unheated control samples. The relative coagulation
time for  the microwave-heated samples remained
unchanged throughout the temperature range studied
(34, 37, 39, and 42°C), and the samples heated  by
radiant energy showed increases  in  the  relative
coagulation time (1.58 and 2.03 times the nonheated
samples) at 39 and 42°C, respectively. A similar but
reverse effect was seen in the relative clot strength.
The samples heated to 42°C by microwaves showed
a decrease to 0.74 times the control samples, and the
radiant-heated samples showed a decrease to 0.31
times the control values. A possible explanation for
these differences is that it is difficult to produce the
same heating rate and pattern within a sample with
different sources of heat.

Ho and Edwards (1977b) measured the rate of oxygen
consumption of mice exposed to 2450-MHz micro-
waves in a waveguide for 30 min at dose rates from
1.6 to 44.3 W/kg at an environmental temperature of
24°C. They found a  significant decrease in the
specific metabolic rate (SMR) at 10.4 W/kg or higher,
but not at 1.5 or 5.5 W/kg. In measurements taken
immediately after exposure, there was no detectable
rectal temperature increase at or below 10.4 W/kg,
but there was a 0.5°C increase at 23.6 W/kg and a
1.0°C increase at 44.3 W/kg. These results  indicate
that the mouse compensates for large dose rates of
microwave  energy by adjusting its metabolic rate
downward to compensate for the thermal load.

In'a more recent report,  Ho  and Edwards (1979)
reported  on a  continuation  of the previous study.
Exposure conditions were identical to those of the
previous study except environmental temperatures of
20, 30, and 35°C were also used. At 20°C, they found
a significant decrease during exposure in the SMR at
12.1 W/kg and above, but not at 6.0 W/kg or below. A
significant decrease was found in the SMR during the
30-min postexposure period at 45.1 W/kg, but not at
27.0 W/kg or less. For the exposures at 30°C there
                       5-52

-------
 Table 6-20.   Summary of Studies Concerning RF-Radiation Effects on Clinical Chemistry and Metabolism

                                             Exposure Conditions
 Effects
                                Frequency
                        Species  (MHz)
Intensity  Duration
(mW/cm2) (days x min) SAR (W/kg)  AT* (°C)
                                                                                   Reference
 No effect on serum chemistry Rat     918(CW)
 values
 Increase in serum glucose    Rabbit   2450 (CW
                                and PW)
 Increase in blood urea       Rabbit   2450 (CW)
 nitrogen
 No increase in blood urea    Rabbit   2450 (CW)
 nitrogen                 Rabbit   2450 (PW)
 Increase in uric acid values   Rabbit   2450 (CW
                                and PW)
                                                              1.0
2.5      91 x 600

5,  10,25  1 x120

25       1 x 120    4.0 (est)
                      Lovely era/. (1977)
                                                              0.8-4.0 (est) 0, 0, 1.7 (PW) Wangemann and Cleary
                                                                        0, 0.2.9 (CW) (1976)

                                                                        2.9
No effect on other serum     Rabbit
chemistry values
Increased iron and manganese Rat
levels in brain
                                2450 (CW
                                and PW)
                                1600(CW)
5 and 10   1 x 120
5,10,25   1x120
10.25     1x120

Negt5
5, 10,      1 x 120
and 25
80        1x10
                        Mouse   2450 (CW)   —
Decrease in specific
metabolic rate
(Ambient T = 24°C)
Increase in specific
metabolic rate
(Ambient T = 35°C)
Increased NADH fluorescence Rat      591 (CW)
Decreased ATP             (exposed
                        Mouse   2450 (CW)   —
                                           13.8
           x30
           x30
         1 xO.5
0.8, 1.6 (est) 0
0.8-4.0 (est) 0,0, 1.7
1.6, 4.0 (est) 0, 1.7(PW)
           0, 2.9 (CW)
Neg 0.8     0
0.8-4.0 (est) 0,0, 1.7(PW)
           0, 0, 2.9 (CW)
48 (est)     4.5

10.4
(Neg 5.5)

8.6        0
(Neg 3.6)

0.36-2.2    0
Chamness et al. (1976)

Ho and Edwards (1977b)


Ho and Edwards (1979)


Sanders et al. (1980)
Decreased CP
Decreased ATP
Decreased CP
Increase in oxygen
consumption
Decreased in metabolic
heat production
No effect on blood
coagulation
brain)
Rat
(exposed
brain)
Rat
Monkey
Human
plasma


591 (CW) 5
2450 —
(120HzPW)
2450 (CW) 6
2450 (CW) 10-280


1 xO.5
1 x30
1 x 10
1 x30


0.13-0.8
6.5, 11.1
(Neg 4.5)
0.9
1 .3-38 (est)


0
0.9, 1.8
0.4
0
Not reported


Sanders et al. (1980)
Phillips et al.
(1975b)
Adair and Adams
(1982a)
Boggsera/. (1972)
*AT = Rectal temperature increase.
fNeg = Effect not found at value indicated.

was no change in the SMR at any of the dose rates
used (1.4 to 23.7 W/kg) either during exposure or in
the  postexposure period.  At 35°C  there was a
significant increase in the  SMR at 8.6 W/kg and
higher, but not at 3.6 W/kg or lower during exposure,
and  no changes were observed during the postexpo-
sure period. The authors reevaluated their first report
(Ho and Edwards 1977b) by comparing the results
during exposure  with the pre-exposure values, not
with those of sham-exposed animals. With  this
method  of evaluation,   the SMR  values  during
exposure at 24°C were  not significantly different
from the pre-exposure values. They reported that this
method  of  comparison accounted for the  lack  of
uniformity among animals at the beginning of each
experiment  and therefore was a better method  of
comparison.

The general trends indicated by their studies are that
microwave exposure at the lowest ambient tempera-
ture  resulted in a reduction  in the SMR, exposure at
                                                     the  highest ambient temperature resulted in an
                                                     increase in the SMR, with no significant changes at
                                                     the intermediate temperatures. A possible interpreta-
                                                     tion of these  trends is that at the lower ambient
                                                     temperatures  the  animals  are producing  heat to
                                                     maintain thermal neutrality; addition of the microwave
                                                     heating reduces the  animals' demand for additional
                                                     endogenous heat. At the higher ambient temperatures
                                                     the animals may be  actively trying to dissipate the
                                                     additional heat from the microwaves as evidenced by
                                                     spreading saliva, thereby increasing their SMR.
                                                     Adult rats were  exposed to 2450-MHz  microwaves
                                                     pulsed at 120 Hz in a multimodal cavity for 30 min at
                                                     SARs of 4.5, 6.5, and 11.1 W/kg; and their metabolic
                                                     rate was measured beginning within 10 min after the
                                                     completion  of exposure  by determining  oxygen
                                                     consumption and carbon dioxide production (Phillips
                                                     et al. 1975b). Room temperature was maintained at
                                                     24.2°C. They found no change in metabolic rate at 4.5
                                                     W/kg,  a decrease  at  6.5 W/kg, and a  greater
                                                                         5-83

-------
decrease at 11.1 W/kg. Colonic temperature measure-
ments were made at the end of exposure, and the
temperatures  were elevated 0.4, 0.9, and 1.8°C for
the 4.5-, 6.5-, and 11.1 -W/kg exposures, respectively.
These results  tend to confirm those initially reported
at 24°C by Ho and Edwards (1977b), where they
compared  the exposed  animals to sham-exposed
animals,  but  not  the reanalyzed  results (Ho  and
Edwards 1979), where they compared the exposure
values with  pre-exposure values on  the same
animals.
Squirrel monkeys  were exposed to 2450-MHz CW
microwaves for 10-min periods at 2.5,4,6,8, and 10
mW/cm2 (SARs from 0.4 to 1.5 W/kg) at ambient
temperatures  of  15,  20, and 25°C  by  Adair  and
Adams (1982a), and the metabolic heat  production
was calculated from oxygen consumption measure-
ments. They found thatfor monkeys restrained at cool
temperatures and exposed at power densities of 4 to 6
mW/cm2 and above, the metabolic heat  production
was reduced in direct relationship to the microwave
energy absorbed. They also exposed  monkeys at 8
mW/cm2 (1.2 W/kg) for 90-min (ambient temperature
= 20°C) and found that the metabolic heat production
initially decreased and then  leveled off at ~ 1.2 W/kg
below pre-exposure values; i.e., the reduction in
metabolic  rate was equal to the rate of microwave
energy deposition.

A study on the effects  of microwaves  on  energy
metabolism of the rat brain  was reported by Sanders
et al.  (1980). First, a small area  of the brain of
anesthetized animals was surgically exposed. Then a
horn antenna  was positioned so that exposures were
in the far field and only the exposed  surface of the
brain was  irradiated with the electric field parallel to
the body axis. Animals were exposed at 591-MHz
(CW) radiation for  0.5, 1,  2, 3, or 5 min at 13.8
mW/cm2  or  for  0.5 or  1 min  at  5  mW/cm2.
(Calculated SAP for 5 mW/cm2 = 0.13 W/kg  using a
2-cm sphere model or 0.8 W/kg using a semi-infinite
plane  model.) During exposures at 13.8 mW/cm2
they found an  increase in nicotinamide adenine
dinucleotide (NADH) fluorescence to a maximum of
4.0 to 12.5 percent above pre-exposure control levels.
In addition, adenosine triphosphate (ATP) levels were
significantly decreased at all exposure times, as were
creatine phosphate (CP) levels. At 5 mW/cm2, ATP
and CP levels were  also  significantly  decreased
following 0.5- and 1 -min exposures. The ATP and CP
changes  at 5 mW/cm2  were not significantly
different from those seen  at 13.8  mW/cm2. There
were no changes in rectal temperature at any of the
exposures and  no significant difference in brain
temperature between  exposed  and sham-exposed
animals.  The authors concluded  that the  results
(increased NADH fluorescence,  decreased ATP and
CP levels) support the hypothesis that RF radiation
inhibits mitochondria!  electron transport chain
function and that the changes cannot be attributed to
general tissue hyperthermia.
In summary, changes in clinical chemistry values
have been reported at dose rates as low as 0.8 W/kg
in rabbits, and negative results have been reported at
exposures as high as 1  W/kg  in rats. The clinical
chemistry changes that have been reported are those
that would be expected  from heat  stress. In other
studies, effects on the rate of oxygen consumption of
mice have been reported  at 10.4 W/kg, in rats at 6.5
W/kg, and in squirrel monkeys at  0.9 W/kg; and
changes in brain energy metabolism have been found
at an SAR estimated to range from 0.13 to 0.8 W/kg
(Table 5-20).
5.7.3 Growth and Development
Few investigators  have  reported the  effects of a
combination  of pre- and postnatal  exposure or
postnatal exposure only to RF radiation on the growth
of laboratory animals (Table 5-21). Smialowicz et al.
(1979a) exposed rats for 4 h per day, beginning on day
6 of gestation through 40 to 41 days postpartum to
2450-MHz (CW) microwaves at  5 mW/cm2 (SARs =
0.7 to 4.7 W/kg), and reported no difference between
the weight gains of the exposed and sham-exposed
animals.
In another study, Smialowicz et al. (1981 a) reported
the growth and development of rats exposed to 100-
MHz (CW) microwaves at an incident power density of
46 mW/cm2 (average SAR = 2.8 W/kg) for 4 h/day
from day 6 of gestation through 97 days of age. The
ambient temperature was maintained at 22°C except
for days 1 to 14, when it was maintained at 27°C (RH =
50 percent).  There was no  consistent difference
between the body weights of the exposed and sham-
exposed animals, though the exposed animals tended
to be larger than the sham-exposed animals. Some of
the animals were tested for neurological development,
and no differences were observed in the development
of a startle response or righting reflex. There was a
significant difference in the age of eye opening, with
the mean age of eye opening in the sham-irradiated
animals occurring  almost one day later than  the
irradiated animals. The authors stated that  the
change probably did not represent an acceleration of
eye opening, as  the age of eye opening  in  the
irradiated animals was similar to that normally seen
in control animals  in other experiments, but that eye
opening was delayed in the sham-irradiated animals.
Tests  for  development of  motor activity at 35 and 84
days of age showed no difference between exposed
and sham-exposed animals. No  difference was
observed between  the exposed and sham-exposed
animals  for complete  blood  counts, mitogen-
stimulated response of lymphocytes, frequency of T-
and B-lymphocytes, or antibody response to Strepto-
coccus pneumoniae capsular polysaccharide.  No
mutagenic effect was observed on the sperm cells
after 20 days using the dominant lethal assay. Seven
                       5-84

-------
regions of the brain were weighed at 22, 40, and 97
days of age, and there was a significant increase in
the weight of the medulla in the irradiated animals at
40 days of age but not at 22 or 97 days of age. There
were no differences in the weights of any other brain
region. There were also no differences in the brain
protein  concentrations for the seven regions. Brain
acetylcholinesterase (AChE) activity, however, was
reduced in the striatum and medulla at 22 days of age
and in the midbrain at 40 days of age. No effects on
AChE activity were noted in the other regions at 22
and 40 days of age, and no differences were noted at
97 days of age. The differences were small, and there
appeared to be no pattern to the changes; the sample
size  was  small  (3 to 5 animals),  and out of 21
comparisons made at the 5-percent confidence level,
one significantly different result would be expected.
Consequently,  without replication  with  a larger
sample size, it is difficult to ascribe these changes to
the microwave exposure.

The effect of postnatal  exposure to microwaves on
growth and development was studied by McAfee et
at, (1973). Weanling mice were exposed to 2450-MHz
microwaves at 10 mW/cm2, 2 min each hour for 24
days; no effect on animal growth was found. There
was no elevation of body temperature in the exposed
animals. The authors also  stated that  a  previous
study, in which  a stimulatory effect on growth from
microwaves was claimed,  was probably  in error
because of  inaccuracies in weighing the  animals
(Nieset era/. 1958).

Guillet and Michaelson (1977) exposed neonatal rats
to 2450-MHz microwaves at 40 mW/cm2 (SARs at 20
to 60 W/kg), 5 min/day for 6 days beginning on day 1
postpartum.  No  effect  on body mass was found,
though the authors did report adrenal changes, as
discussed in Sec. 5.7.1, Endocrine Effects.

Stavinoha et al.  (1975) irradiated 4-day-old mice at a
field intensity of 5800 V/m (8.92 W/cm2) for 20 min
with  10.5-, 19.27-,  or 26.6-MHz radiation, and
observed no change in growth rate to 22 days of age
compared to the control animals (SARs estimated at
0.9, 1.8, and 3.6 W/kg). Mice were also exposed  to
19-MHz microwaves for 40 min/day for 5 days from a
near-field synthesizer that delivered an E field of 800
V/m (17 W/cm2) and an H field of 55 A/m (114
W/cm2); no change in body mass through 120 days of
age was found (SAR estimated at  6.3 W/kg).

In a chronic study, Lin et al. (1979a) exposed 4- to 7-
day-old mice to 148-MHz microwaves at 0.5
mW/cm2 (SAR = 0.013 W/kg), 1 h/day, 5 days/week
for 10 weeks and reported no significant effect on
weight gain.

Albert et al. (1981 a) reported a significant decrease in
the number of Purkinje cells in the cerebellum of rats
exposed in utero (days 16 to  21  of gestation) and
postnatally for 97 days (4 h/day) to 100-MHz (CW)
 microwaves at 46 mW/cm2 (SAR = 2.8 W/kg). The
 animals were from the study by Smialowicz et al.
 (1981 a), as reported above. A similar decrease was
 found  in rats exposed 21 h/day in utero (days  17 to
 21) to 2.45-GHz (CW) microwaves at 10 mW/cm2
 (SAR = 2 W/kg). In rats exposed 7 h/day for 5 days
 beginning at 6 days of age, a decrease in the number
 of  Purkinje cells was seen in animals immediately
 after exposure, but after 40 days there was apparent
 recovery.

 In a report  by Kaplan et al. (1982) squirrel monkeys
 were exposed to 2.45-GHz microwaves for 3 h/day in
 utero and through 9.5 months of age at 0.034, 0.34,
 and 3.4 W/kg.  The  results indicated a possible
 increase in mortality at 3.4 W/kg. Because of the
 small number of  animals involved, a follow-up study
 was done  with  larger  number of animals (Kaplan
 1981),  and this study showed no  difference in the
 mortality between the irradiated  (3.4 W/kg)  and
 sham-irradiated  animals.  Other aspects  of these
 papers are discussed in Sec. 5.3.1.2, Reproductive
 Effects in Mammalian Models.
 Albert  et al. (1981b) examined the squirrel monkeys
 described in the  preceding paragraph (Kaplan et  al.
 1982) and found no change in the number of Purkinje
 cells as compared to the sham-irradiated animals.
 These studies indicate that RF exposures at SARs <
 2.8 W/kg have no effect on growth and development
 of animals exposed pre- and postnatally or postnatally
 on ly. Exposures at 20 to 60 W/kg postnata lly for short
 exposure times also did not alter the growth rate of
 animals even though adrenal changes were seen.

 5.7.4 Cardiovascular System
 5.7.4.1 Whole-Body Exposures
 Some of the initial studies of biological effects  of
 microwave  radiation were made on the cardiovascular
 system. Presman and Levitina (1962) reported  that
 whole-body or ventral exposure to CW microwaves at
 2400 MHz, 7 to 12 mW/cm2, promoted a decreased
 heart  rate  (bradycardia) in the rabbit.  However,
exposing only the head to microwaves at the same
power  densities resulted  in an inceased heart  rate
(tachycardia). The SAR could not be estimated as the
animals were exposed in the near field; therefore, the
study  is not  listed with  other  cardiac physiology
studies  in Table 5-22. The report is mentioned here
because it  provided the impetus for other studies
discussed below.
 Kaplan  et  al.  (1971) attempted to replicate the
tachycardia reported  by Presman  and Levitina  by
exposing the head of  rabbits to 2400-MHz (CW)
microwaves at  an incident  power density of 10
mW/cm2 (SAR estimated at 2 W/kg) for 20 min. The
exposure conditions differed in  that Kaplan et al.
exposed the animals  in the far field, whereas
Presman and Levitina had used near-field exposures.
                                                                     5-85

-------
Table 5-21 . Summary of Studies Concerning RF-Radiation Effects on Growth and Development
Exposure Conditions
Effects
No effect on weight gain
No effect on growth, neuro-
logical or immunological
development or mutagenicity
Possible decrease in brain
acetylcholinesterase activity
Decrease in Purkinje cells
Decrease in Purkinje cells
Decrease, then recovery,
in Purkinje cells
No change in Purkinje cells
No change in infant mortality
No effect on growth
No effect on body weights
No effect on growth

No effect on weight gain
Species
Rat
Rat

Rat
Rat
Rat
Monkey
Monkey
Mouse
Infant
rat
Mouse
Mouse
Mouse
Frequency
(MHz)
2450 (CW)
100 MHz (CW)

100 MHz (CW)
2450 MHz (CW)
2450 MHz (CW)
2450 MHz (CW)
2450 MHz (CW)
2450 (CW)
2450 (CW)
10.5, 19.27,
26.6 (CW)
19
148 (CW)
Intensity
(mW/cm2)
5
46

46
10
10


10
40
8900
17,000-
114,000
0.5
Duration
(days x min)
55x240
112x240
37.55
x240
112x240
5x 1260
5x240
285 x 180
285x180
24x48
6x5
1 x20
5x40
50x60
SAR (W/kg)
0.7-4.7
2.8

2.8
2
2
3.4
3.4
6-8 (est)
20-60 (est)
0.9, 1.8
3.6 (est)
6.3 (est)
0.013
AT' (°C)
Not reported
Not reported

Not reported
Not reported
Not reported
Not reported
Not reported
Not reported
1.5-2.5
Not reported

Not reported
Reference
Smialowicz et al. (1979a|
Smialowicz et al. (19818)

Albert era/. (198 la)
Albert « •/. (1981 a)
Albert era/. (1981a)
Albert era/. (1981 b)
Kaplan (1981)
McAfee era/. (1973)
Guillet and Michaelson
(1977)
Stavinoha el al.
(1975)

Lin era/. (1979a)
"AT = Rectal temperature increase.
Kaplan et al. reported no change in heart rate. They
then exposed rabbits at increasing power densities
(20, 40, 60, 80, and 100 mW/cm2) and measured
respiration  rate, body temperature, and heart rate.
Respiration increased at 40 mW/cm2 and greater,
and body temperature was elevated at 80 and 100
mW/cm2 (0.5°C at both power densities). Heart rate
was increased at 100 mW/cm2 only.

Birenbaum  et al.  (1975) exposed  the  heads of
unanesthetized rabbits to 2.4-GHz (CW) microwaves
for 60 min at 20,  40, 60, and 80 mW/cm2 (SAR
estimated at 3 to 12 W/kg), and found increases in
heart rate, respiration rate, and subcutaneous
temperature (lower back) at all four power densities,
with greater increases at higher  exposure  levels.
They  also compared 2.8-GHz (CW and  PW, 1000
pulses/s, 1.3 /us) microwaves at 20 mW/cm2 for the
same three parameters and found no differences in
the responses to CW or PW irradiation.
Phillips er al. (1975b) exposed adult rats to 2450-MHz
microwaves pulsed at 120 Hz for 30 min (SARs = 4.5,
6.5, and  11.1 W/kg) and measured colonic and skin
temperatures and heart rate after exposure. They
reported  increases in colonic and skin temperatures
with  increased  exposure levels immediately after
irradiation. At the  highest exposure, the colonic
temperature dropped  below normal at  1 h after
exposure and remained subnormal for 4 h. There was
no postexposure effect on heart rate at 4.5 W/kg;
however, there was a mild bradycardia at 6.5 W/kg
and the more pronounced decrease in heart rate at
11.1 W/kg. They attributed the effect to the heat
stress induced  by the microwaves. They also
discussed the possibility that heating the regulatory
center in the hypothalamus may have stimulated the
fall in body temperature.

Hamrick and McRee (1980) assessed the effects of
body temperature on the heart rate of embryonic
quail. They  exposed the  embryos to 2450-MHz
(CW) microwaves for 5 to 10 min (SARs = 3, 6, 15,
and 30 W/kg) at incubation temperatures from 35 to
38°C, and to 2450-MHz (PW) microwaves (10-^s
pulses, varied from 10 to 50 pulses/s; SARs at 0.3,
1.5, and 3 W/kg) at incubation temperatures of 35 to
39°C. There were no significant differences between
the heart rates of the exposed and control embryos in
any of the groups at any of the temperatures used. The
authors did observe that the embryonic heart rate
increased ~  23 beats/min for each  1°C  rise in
incubation temperature in the 36 to 39°C range.
Chou et al. (1980b) exposed rabbits both dorsally and
ventrally to 2.45-GHz microwaves 20 min/day for 10
days to both continuous and pulsed waves (100
pulses/s, 1 -fjs  pulse width, and 10-/us  pulses
triggered by the R wave at various delay times). The
incident  power density  for the CW  and  the PW
condition was 5 mW/cm2, with a calculated local
SAR to the heart of 0.093 W/kg for dorsal irradiation
and 0.3 W/kg for ventral  irradiation. No effect on
heart rate was found. Rabbits exposed at 80 mW/cm2
showed an increase in heart rate, presumably from
heat stress.
                       5-86

-------
5.7.4.2 Isolated Heart Preparations

Although the data supporting a microwave-induced
bradycardia in the intact animal are equivocal, some
researchers have exposed isolated heart preparations
and reported an effect of microwaves on heart rate.
Tinney et at. (1976) attempted to determine the locus
of microwave-induced bradycardia  in the  isolated
turtle heart  using 960-MHz (CW) microwaves.
Although the heart was isolated from the CMS, it was
still capable of responding to neurohumoral agents.
They found  that exposing the  isolated  heart to
microwaves at 8 mW/cm2 caused bradycardia. When
the heart was treated with atropine to block the
sympathetic system, tachycardia resulted. However,
when the heart was treated with propranolol to block
the parasympathetic system, the heart rate decrease
was more significant than that following microwave
exposure alone, which would show that microwaves
                          may have  affected the  parasympathetic  reflexes.
                          With  both blocking agents added to the heart,
                          microwave exposure had only slight effects on the
                          heart  rate.  Tinney et at. postulated that microwave
                          exposure of the isolated heart equally enhances the
                          release of acetylcholine and norepinephrine; however,
                          since the activity of the former usually predominates
                          over the latter (Johansen 1963), the net  effect of
                          microwave exposure is a decrease in heart rate.
                          A similar  drug-microwave interaction has been
                          observed in the  isolated rat heart by Olsen  et al.
                          (1977).  They exposed the heart  to 960-MHz (CW)
                          microwaves for 4  min  at 1.3  and 2.1 W/kg  and
                          observed a decrease in heart rate. The decrease was
                          greater  at  1.3 W/kg than at 2.1 W/kg. They also
                          conducted studies at 2.1 W/kg with drugs to block the
                          sympathetic and parasympathetic nervous system
                          and obtained the same results as Tinney et al. (1976).
Table 5-22.    Summary of Studie* Concerning RF-Radiation Effect* on Variou* Aspect* of Cardiac Physiology

                                                    Exposure Conditions
Effects
Species
Frequency
(MHz)
Intensity
(mW/cm2)
Duration
(total min)
SAR
(W/kg)
Reference
Bradycardia develops after
whole-body exposure, along
with hyperthermia

Exposure to head promotes
tachycardia; exposure to
back raises respiratory rate
but not heart rate
Increased respiration

Increased heart rate from
dorsal exposure of the head

Alterations in ECG
(shortening of QT interval,
increased height of T-wave,
appearance of U-wave)
No effect on heart rate that
cannot be attributed to
microwave heating
Pulses synchronized with each
R-wave do not affect heart
rate

Synchronized pulses with QRS
complex causes increase in heart
rate with some arrhythmias
Increased heart rate

No effect on heart rate

Low power levels cause
bradycardia in the isolated
turtle heart

Causes slight decrease in the
isolated heart

Synchronized exposures with ECG
have no effect on heart rate
  Rat    2450 (PW)    28,48
 Rabbit
Rabbit

Rabbit

 72-h
 Chick
 haart


 Quail
embryo

 Frog
 Frog



Rabbit


Turtle



 Rat


 Frog
2400 (CW and  20
PW)
2400

2400
40-100

100
24,000 (PW)   74
2450 (CW and  NAt
PW)
2450 (CW)     80

             5
960 (CW)      NA
960 (CW)
NA
1420-3000    0.0006
(PW)
                        30
                                60
20

20
                      6.5, 11.1
            5 to 10
1420-10,000   32/yW/cm2  100-^s
(PW)                    pulses


1425 (PW)     0.6/uW/cm2  10-jus
                        pulses
20x10d

20 x 10 d

60



5 to 10


2-, 10-,
150-/US
pulses
8-20 (est)
20 (est)

NG"
          0.3-30
                      NG
                       Phillips et al. (1975b)
                                  Birenbaum et al. (1975)
Kaplan et al. (1971)
Kaplan et al. (1971)

Paff era/. (1963)
             Hamrickand McRee(1980)
                                  Liu et al. (1976)
                       Frey and Seifert (1968)
12           Chou et al. (1980b)
0.3 and 0.093
2-10         Tinney ef al. (1976)


1.3 and 2.1    Olsen et al. (1977)

NG           Clapman and Cain (1975)
*NG = Not given.
tNA = Not applicable.
                                                                           5-87

-------
Paff et al. (1963), working with the isolated embryonic
chicken heart, were unable to detect changes in heart
rate during exposure to 24,000-MHz (PW) radar fields.
They did, however, detect effects on the electrocar-
diogram (ECG), including abnormal P andT waves from
3-min exposures at 74 mW/cm2.

Frey and Seifert (1968) showed that 10-//S pulses at a
carrier frequency of 1.425 GHz given at a synchronous
period  with  the ECG (220 ms after the P wave)
resulted in tachycardia or heart arrhythmia in the
isolated frog heart. The peak power density was 60
mW/cm2 (average power density — 0.6/uW/cm2). Liu
et al.  (1976) reported no effect on heart rate with
isolated frog hearts or in hearts irradiated in situ in a
similar study. The in situ hearts were exposed to 100-
fjK pulses of either 1.42 or 10 GHz, and the isolated
frog hearts were exposed to 100-//S pulses of 1.42
GHz. The pulse was delivered on the rising phase of
the R-wave from the  ECG, which as somewhat
similar to, but not exactly the same as, the 200-ms
delay following the P-wave used by Frey and Seifert.
(The R-wave follows  the P-wave by about 200 ms.)
The peak and average power densites of 320mWand
32 (AN were also considerably higher than those used
by Frey and Seifert. These factors, plus differences in
the manner of preparing the isolated hearts (Liu et al.
curarized the frogs, whereas Frey and Seifert
decapitated the frogs), make it difficult to compare the
results of the two studies.
Clapman and Cain (1975), however, tried to replicate
the study of Frey and Seifert  using  similar pulse
widths (10 //s), peak and average power densities (60
mW/cm2 and 0.6 //W/cm2), carrier frequency (1.42
GHz), and method of isolating  the frog heart; they
reported no change in heart rate. Also, no heart rate
changes were found  when they conducted studies
with a different peak power (5.5 W/cm2), frequency (3
GHz), and pulse widths (2 and 150 /us). Clapman and
Cain were able to produce an  increased  heart rate
with 20-mA current pulses synchronized 200 ms after
the P-wave peak.
The results of microwave exposure on the cardiovas-
cular system (Table 5-22)  indicate that whole-body
exposure of sufficient intensity to produce heating
also produces an increase in heart rate similar to that
which would be expected from  heating alone. In the
isolated heart there appears to be a stimulation of the
autonomic nervous system from microwave exposure
at levels where very little heating would be expected
(1  to 2 W/kg).  Low levels of synchronized PW
microwaves (0.6 to 32  mW/kg) apparently are
ineffective in producing detectable alterations in
heart rate.

5.7.5 Biological Effects of Low Frequency
Modulation of RF Radiation
Interest in the  biological effects of low frequency
modulation  of RF radiation stems from  reports of
changes caused by exposure to electric and magnetic
fields in the sub-ELF range (0 to 30 Hz). It has been
reported  that  exposure to low-frequency electric
fields changes the reaction time in humans (Konig
and Ankermuller 1960; Hamer 1968; Konig 1971) and
in monkeys (Gavalas; etal. 1970; Gavalas-Medici and
Day-Magdaleno 1976), and alters circadian activity
in human beings (Wever 1973). Friedman etal. (1967)
observed that magnetic  fields modulated at low
frequencies also change reaction time in human
beings.

Two other studies that provide important background
information are reported by Kaczmarek and Adey
(1973, 1974). In the first report, they described
release of  calcium ions  and  x-aminobutyric acid
(GABA) from the cerebral cortex of cats in response to
small changes in the extracellular concentration of
calcium.  In 1974,  they  demonstrated  release of
calcium ions  and GABA from the  cat cortex in
response to low intensity electric currents, pulsed at
200 Hz, applied directly to the cerebral cortex. Thus,
extracellular  calcium  and electric  current  have
similar effects on the release of GABA and calcium
ions from brain tissue.

The studies of (1) behavioral changes in animals and
human beings induced by low frequency signals and
(2) biochemical changes in the cat brain caused by
electric currents led to a study of the influence of
electric fields  on EEG patterns associated with a
conditioned behavioral response in cats (Bawin et al.
1973). To increase the penetration of the signals into
the tissue, they chose an RF carrier wave of 147 MHz,
which was amplitude modulated at sub-ELF frequen-
cies (e.g., 3 to 14 Hz). Alterations  were observed in
the rate of performance, accuracy  of  reinforced
patterns, and resistance to extinction  in learned
behavior of  the exposed  animals  compared to
controls, indicating that the fields were acting as
reinforcers. In  order to determine whether  these
effects were  mediated via  peripheral receptors or
occurred as a  result of changes induced directly on
the CNS, experiments were designed to examine the
effects of modulated RF carrier waves on brain tissue
in vitro.

5.7.5.1 Calcium Ion Efflux In Vitro: A
Fundamental Finding

The association of calcium ions with brain tissue was
selected as the biochemical marker to examine the
influence of modulated RF fields because calcium ion
efflux has  been shown to  be sensitive to electric
currents applied directly to brain tissue in vitro, and
because calcium ions have a prominent role in many
biochemical and biophysical processes (e.g., cellular
membrane integrity and function, enzyme cofactor,
putative second messenger for the  conduction of
extracellular signa'ls to the nucleus of the cell, neural
tissue excitation  and secretion of transmitter
                       5-88

-------
substances at  synapses). The first  report  of
the influence of modulated RF fields on excised brain
tissue was Bawin et al. (1975), who showed that a 20-
min exposure of chick brain tissue in vitro to a 147-
MHz field at 1 to 2 mW/cm2 (SAR estimated at 0.002
W/kg) caused  enhanced efflux of calcium ions, but
only if the field was sinusoidally amplitude modulated at
frequencies of  6, 9, 11, 16, or 20 Hz. Maximal efflux
was measured at 16 Hz. Modulation frequencies of 0,
0.5,3, 25, and 35 Hz were ineffective. This frequency-
specific  response, which occurred while the 147-
MHz carrier field was maintained at the same power
density,  indicates that  the  field-induced efflux  of
calcium  ions was not due to heating of the samples.

In another  report, Bawin et al. (1978) exposed chick
brain tissue for 20 min to 450-MHz fields, amplitude
modulated at 16 Hz, at 0.75 mW/cm2 (SAR estimated
at  0.0035 W/kg)  under a variety of chemical
conditions. The  results  demonstrated that  (a) the
enhanced efflux of calcium ions is not highly sensitive
to the external  calcium concentration, (b) bicarbonate
appears to be important for enhanced efflux,  (c)
lowering the pH  from 7.6 to 6.8 in the presence of
bicarbonate may enhance the magnitude of efflux,
and (d) lanthanum causes a reversal to field-induced
retardation of calcium ion efflux.

Corroboration of the frequency-specific response
described by Bawin and co-workers was provided by
Blackman  et al. (1979), who  showed that 16-Hz
amplitude  modulation of  147-MHz  carrier waves
caused enhanced efflux in chick brain tissue in vitro,
whereas modulation frequencies  of 3, 9, and 30  Hz
did not. Although the data  had large variances,  an
unusual intensity response was described, i.e., only
0.83  mW/cm2 (SAR estimated  at 0.0014 W/kg)
produced a statistically  significant efflux enhance-
ment (intensity values  are  corrected  based  on
discussion  in Blackman era/. 1980a); power densities
(0.11,0.55,1.11 and 1.38 mW/cm2) below and above
the effective value did  not  cause efflux. In a  later
report,  Blackman et al. (1980a) used a revised
statistical  model and experimental  procedure  to
reduce the  influence of the large sample variance. An
intensity response identical to their earlier result was
found. However, when the distance between samples
was halved, the range of intensities that produced
enhanced efflux increased to include 0.55,0.83,1.11
and 1.38 mW/cm2, whereas lower and higher values
of  0.11 and 1.66 mW/cm2  were ineffective.  In
addition, an intensity  region from 0.55 to 1.11
mW/cm2 caused enhanced efflux when 9 Hz was
used as the modulation frequency.  These data,
obtained with a more rigorous experimental protocol,
provided additional support for the results of Bawin et
al.  (1975) and  Blackman et al. (1979); however, the
explanation for the dependence on sample spacing
awaited further developments.
Joines et al.  (1981) examined the dependence on
sample spacing by calculation of the electrical
coupling  between  the samples; for  similicity the
samples were modeled as spheres. They found that
increased electrical interaction  between the more
closely packed spheres produced a broader range of
internal field strengths within each sphere. Thus, if a
given internal field strength were necessary to cause
enhanced efflux, the chance would be greater for that
internal field strength  to occur in closely  coupled
samples  exposed to a specific  range  of  incident
intensities. Joines et al. (1981) found this result to be
consistent with the  experimental findings  in
Blackman et al. (1980a). Thus a potential  artifact was
shown to be  a logical  result of the experimental
procedures.

The intensity  response  observed by Blackman et al.
(1979) with modulated  147-MHz carrier waves was
confirmed by  Sheppard ef al. (1979) with 450-MHz
carrier waves, modulated at 16 Hz; calcium-ion efflux
was enhanced at 0.1 and 1.0 mW/cm2 but not  at
0.05, 2.0, or 5.0 mW/cm2. (The estimated SAR at 1.0
mW/cm2 is 0.0047 W/kg.) The results of these two
reports show that the intensities producing calcium-
ion efflux from chick brain tissue in vitro are within
the range of 0.1 to 1.38 mW/cm2for modulated 147-
MHz and 450-MHz carrier waves.

The  apparent carrier-frequency independence  of
effective intensities was tested with a 50-MHz carrier
wave, amplitude modulated at 16 Hz. Enhanced efflux
of calcium ions occurred within two intensity regions
(between 1.44 and 1.67, and at 3.64 mW/cm2; SARs
were 0.0013 and 0.0035  W/kg, respectively)
separated by intensities of no effect, including 0.72
mW/cm2 (Blackman et al. 1980b). These effective
intensity values were different from the corresponding
values of 147-MHz radiation; thereby indicating  a
dependence on carrier frequency.  In addition this
result revealed the existence of more than one range
of effective intensities.

The apparent discrepancy in effective  power
densities at the three  different carrier  frequencies
(50,  147, and 450 MHz) has been resolved by the
finding that efflux is dependent on the electric field
strength within the tissue and not on incident
intensity (Joines and Blackman 1980). The calcula-
tion  to transform the  incident intensity to  internal
field strength was based on an empirical model
described by Joines et al.  (1981). With the  data
available at 50 and 147 MHz, the model was used to
predict intensities that would produce both altera-
tions and no alterations in calcium-ion efflux; some
predictions were  tested and found  to  be valid
(Blackman et al. 1981). These reports described two
intensity ranges that appear effective for enhanced
efflux at both 50 and 147 MHz, identified the internal
electric field strength ratherthan incident intensity as
the important exposure parameter,  and  showed the

                   5-89

-------
importance of frequency-dependent complex permit-
tivity values of brain tissue in the conversion of
incident  intensity to internal  field strength.  The
exposures at 50 and 147 MHz caused no generalized
heating of the sample.  The maximum temperature
rise was  calculated to be <0.0004°C,  and SAR
calculated at each  carrier frequency was <0.0014
W/kg (Blackman et a/. 1980b).

Subsequent to the critique by Athey (1981) that the
simple spherical model used by Joines and Blackman
(1980) was too idealized, these authors showed that a
layered sphere model produced relationships between
incident intensities at 50,  147, and 450 MHz and
internal field strengths that were also consistent with
the experimental results  (Joines  and Blackman
1981).  The success  of the  initial, simple models to
predict intensity regions of  both field-induced efflux
enhancement and  no enhancement demonstrates
the utility of the approach. More refinements in the
models are  necessary before the experimental
situation  is realistically described.
Shelton  and Merritt (1981), who  used different
procedures from those described by Bawin et af.
(1975), Blackman  et  a/.  (1979,  1980a,b),  and
Sheppard et a/. (1979) reported no change in calcium-
ion efflux from rat brain. Brain tissue, labeled in vitro
with radioactive calcium, was  irradiated at  1 GHz,
pulse-modulated with square waves at 16 or 32 Hz
(0.5, 1.0,  2.0, and 15 mW/cm2). In a second report,
Merritt et a/. (1982) exposed rat brain tissue labeled in
vivo with  radioactive calcium to microwave radiation,
pulse  modulated at 16 Hz (20-ms pulse width). The
intensities for the 1-GHz carrier frequency were 1
mW/cm2 (SAR = 0.29 W/kg) and 10 mW/cm2 (SAR =
2.9 W/kg); and for the 2.45-GHz carrier frequency, 1
mW/cm2  (SAR = 0.3 W/kg). In addition, animals
labeled with radioactive calcium  were exposed for 20
min to 2.06-GHz radiation  at  one  of 17 different
combinations of intensity and pulse repetition rate: 0,
0.5, 1.0, 5.0,10.0 mW/cm2 (SAR was 0.24 W/kg per
mW/cm2); and 0, 8, 16, 32 Hz (pulse width was 10
ms). After exposure, brain  tissue was analyzed for
radioactivity. No statistically significant field-induced
enhancement  of calcium-ion efflux or change of
calcium content in the brain tissue was found. The
reason for these negative  findings  is not known;
however,  the use of square wave rather than sine
wave modulation, the different biological preparation,
and different medium composition are factors  that
may have influenced the outcome.

5.7.5.2 Additional CNS Studies
The reports of field-induced calcium-ion efflux from
chick  brain tissue in vitro have led to other CNS
studies. Synaptosomes, prepared from rat cerebra
and labelled with radioactive calcium, were exposed
for 10 min  at 0.5  mW/cm2  to 450-MHz fields,
amplitude modulated at 0, 16, or 60 Hz (Lin-Liu and
Adey 1982). Only 16 Hz affected the efflux kinetics of
calcium ions. Although the SAR can be estimated as
low,  an exact  value cannot be unequivocally
established because the exposure chamber may have
been operated in a multimodal condition. (See Weil et
a/.  1981.) Nevertheless,  this result is modulation
dependent, and it is unlikely that heating is  involved
as a causative agent.
Similar field-induced efflux enhancement has been
reported in a live animal. Adey et al. (1982) exposed
awake, immobilized cats to 450-MHz fields, amplitude
modulated  at  16 Hz, at 3.0 mW/cm2  (SAR = 0.29
W/kg). The  release of calcium ions from the cortex
was observed as a function of time. Irradiation for 60
min caused episodes of enhanced efflux lasting 20 to
30 min and extending into the postexposure period.
Although focusing on a different component  of the
efflux kinetics than that studied by Lin-Liu and Adey
(1982), these results  demonstrate  that RF  fields
modulated  at  16 Hz can cause changes in both a
subcellular membrane  system  and in the live
mammal. Thus, the field-induced phenomenon is not
restricted  to an avian  species nor to  in  vitro
preparations.

Recently, Dutta et al. (1984) observed field-induced
enhancement of  calcium ions from cells of human
origin. Monolayer cultures of human neuroblastoma
cells were exposed for 30 min at ten SARs from 0.01
to 5.0 W/kg to  915-MHz  fields,  with or  without
sinusoidal amplitude modulation (80 percent) at
frequencies between  3  and  30 Hz.  Significant
increases in the efflux of calcium ions occurred at two
SARs (0.05 and 1.0 W/kg).  The increased efflux at
0.05 W/kg was dependent on the presence of 16-Hz
modulation  but not at the  higher value. Exposure at
modulation frequencies between 3 and 30 Hz (SAR -
0.05 W/kg) revealed a peak in the response at 16 Hz.
Although the effective SAR (0.05 W/kg) for 16-Hz
modulation  is more than 38 times greater than the
SARs for enhanced efflux of calcium ions from chick
brain  tissue in vitro, the low-frequency response
pattern was  similar to that reported by Bawin et al.
(1975) and  Blackman et al.  (1979). The relation  of
enhanced efflux with unmodulated fields at 1.0 W/kg
with the effects of modulated fields is not known at
this time; however,  it is not due to a  temperature
increase  in the sample because enhancement was
not found at SARs of 2.0 and 5.0 W/kg.

The effect of modulated RF fields on the EEG was
investigated by Takashima et al. (1979). Rabbits were
exposed 2 h daily for 6 weeks to 1.2 MHz, amplitude
modulated at 15 Hz, or 5 MHz amplitude modulated at
14 Hz. Following exposure,  the EEG was recorded
with scalp electrodes and,  when compared to the
pretreatment EEG pattern, was found to be altered
with  enhanced  low-frequency components and
decreased high-frequency  components. The EEG
pattern returned to the pretreatment pattern after
                       5-50

-------
several weeks postexposure. Although the electric
field intensity was given as 500 V/m, with an error
factor as large as 2, the important  aspect of the
results was that  unmodulated fields of similar
intensity  had  no effect on the  EEG pattern. The
absence of metallic electrodes  in the animal during
exposure avoids the major criticism of earlier studies
that reported field-induced changes in EEG patterns
(Gavalas et al.  1970;  Bawin et al. 1973).

Sagan and Medici (1979) studied the influence of
450-MHz fields, sinusoidally amplitude modulated at
either 3 or 16 Hz, on  locomotor activity  in young
chickens. The experiments were performed  in  a
plastic, modified Skinner box with light beams to
monitor activity; the complete apparatus was placed
in an anechoic chamber and exposed in the far field.
The authors found no statistically significant change
in performance during or immediately after a 23-min
exposure at 1 or 5 mW/cm2 (SAR estimated at 0.2
and 1.0 W/kg). They concluded that the lack of a field-
induced  response could be  due  to the use of
modulation frequencies not present in the chicken's
EEG during performance on the particular (fixed-time
schedule) task. An alternative possibility, based on
the multiple-intensity  ranges  observed for field-
induced calcium-ion efflux, is that the two intensities
used in this study may have been outside the effective
ranges.

In summary, four groups (Adey et al.: Blackman et al.;
Dutta et al.; Takashima et al.) have shown that RF
fields, sinusoidally modulated at sub-ELFfrequencies,
especially 16 Hz, cause CNS changes  in different in
vitro preparations and  in the live animal. Many of
these studies have been analyzed in reviews (Adey
1981; Blackman et al. 1981; Greengard et al. 1982;
Myers and Ross 1981). It is generally agreed that both
the mechanism of interaction and the physiological
consequences of  these changes  are yet to be
established.

5.7.5.3 Non-CNS Studies
The effects of  exposure of pancreatic tissue and T-
lymphocytes to  RF fields, sinusoidally amplitude
modulated at low frequencies, have been examined.
An increase of calcium-ion efflux from rat pancreatic
tissue exposed in vivo at 2 mW/cm2 for 1 to 2.5 h at
147 MHz,  modulated at  16  Hz (estimated SAR <
0.075 W/kg),  has been  reported by Albert et al.
(1980). However, the  efflux was not accompanied by
a change in protein secretion, which is  normally
associated with calcium mobilization in the pancreas.
The authors attributed the lack of protein secretion to
a limitation imposed by the exposure conditions, i.e.,
a relatively small volume of medium was available to
the tissue for normal metabolic activity.

In another in  vitro assay, the  cytotoxic activity of
mouse T-lymphocytes  was suppressed  by a  2-h
exposure (1.5 mW/cm2)to450-MHzfields, modulated
at frequencies between 16 and 100 Hz (Lyle et al.
1983). Peak suppression occurred at 60-Hz modula-
tion, with smaller effects at 16, 40, 80, and 100 Hz.
The  exposed  cells recovered full cytotoxic activity
12.5 h after the termination of exposure. This result
demonstrated an inhibitory but reversible effect on a
cell-mediated immune response  by modulation
frequencies.

5.7.5.4 Sinusoidal ELF and Sub-ELF Signals
Most of the studies reviewed above demonstrate an
absolute requirement for low-frequency sinusoidal
modulation of the RF carrier wave in order for the
signal to be effective biologically. For completeness,
several reports are mentioned that describe biological
effects of exposure to low frequencies in the absence
of an RF carrier wave. Bawin and Adey (1976,1977)
exposed chick and cat cerebral tissue for 20 min to 1,
6,16, 32 or 75 Hz at electric field gradients of 5,10,
56, and 100 Vp-p/m  in air. Only two frequencies, 6
and 16 Hz, caused a reduction in calcium-ion efflux at
10 and 56 V/m for the chick tissue, and at 56 V/m for
the  cat tissue.  Because  all other combinations
produced no  field-induced responses, the authors
described "amplitude and frequency windows" for
calcium-ion efflux. Electric field gradients within the
tissue  were estimated to be  10r5 V/m. The field-
induced reduction in efflux is  in  contrast to the
enhancement caused by modulated RF carrier waves.
Nevertheless, the frequency dependence observed in
the two studies was similar, which suggests  an
interaction with a common substrate as the site of
interaction.

Blackman et al. (1982) used chick brain to study the
influence of 16-Hz signals at 15 intensities between 1
and  70 Vp-p/m on the efflux of calcium ions. Two
intensity regions that included 5,6, and 7.5 V/m and
35, 40, 45, and 50 V/m caused enhanced efflux. No
field-induced effects were seen below (1, 2, and 3.5
V/m), between (10, 20, and  30 V/m), or above (60
and  70 V/m)  the two effective intensity  regions.
Moreover, 1 - and 30-Hz signals at 40 V/m caused no
change in efflux. This finding is consistent with the
reports of multiple-intensity  regions of enhanced
efflux caused by modulated RF radiation (Blackman
et al.  1980b,  1981). In  addition to the  intensity
response, the  frequency dependence corroborated
reports by Bawin and Adey (1976) for low-frequency
signals, and by Bawin et al. (1975) and Blackman et
al. (1979) for modulated RF fields.

In these two low-frequency studies,  the cause of the
slight difference in effective intensities is unknown.
The major disagreement in the results of Bawin and
Adey (1976)  and Blackman  et  al. (1982) is the
direction of the change in efflux; the latter authors
state that the  "cause may be found in  the slightly
different preparations and procedures used in the two
laboratories."
                                                                      5-91

-------
Several research groups have reported biological
changes induced by low-frequency,  sinusoidally
oscillating magnetic fields. The myxomycetePhysarum
polycephalum has a longer mitotic cycle and reduced
respiration rate after chronic exposure to 2.0-gauss
magnetic fields at 75 Hz (Goodman et al. 1979).
Human fibroblasts in culture exposed to sinusoidally
varying  magnetic  fields  for a  wide range of
frequencies (15 Hz to 4 kHz) and amplitudes (0.25 to
5.6 gauss) exhibit enhanced DNA synthesis (Liboff et
al.  1984).  Fruit flies (Drosophila melanogaster)
preferred not  to  deposit  eggs  in  a 10-gauss,
sinusoidally  varying  50-Hz  magnetic field;  similar
exposure during development of the egg produced
less viable eggs and  pupae  in the exposed samples
than in controls (Ramirez et al.  1983). These results
suggest that low-frequency, sinosoidally varying
fields may alter fundamental biological processes.
Low-frequency, pulsed  magnetic fields  have also
been  reported  to produce alterations in diverse
biological systems.  These 'systems  include  the
developing chick embryo (Delgado et al. 1982; Ubeda
et al.  1983), Drosophila egg laying and mortality
(Ramirez et al. 1983), the de-differentiating amphibian
red blood cell (Chiabrera et al. 1979), transcription in
the Dipteran chromosome (Goodman et al. 1983),
nerve cells in  culture (Dixey and Rein 1982),  and
mouse bone cells in culture (Luben era/. 1982). Many
of these studies used an intricate pulsed waveform,
which has been used in therapeutic devices for bone
nonunions. All the studies used pulse repetition rates
below 500 Hz, with most below 100 Hz. Recently,
Liboff  et al. (1984)  questioned the  need for  the
particular wave shapes because it appears that the
essential element is the low-frequency field.

5.7.5.5 Summary
Many reports of effects of RF fields that are amplitude
modulated at very low frequencies have not been
independently corroborated. The major exception is
calcium-ion efflux from chick brain tissue in vitro at
intensity levels far below those that cause heating.
This exception, combined with the results of studies
of brain biochemistry and EEGs in animals and with
synaptosomes  and human  neuroblastoma cells in
culture, provides evidence  that CNS tissue from
several species, including human beings, is affected
by  low-intensity RF  fields sinosuoidally amplitude
modulated at specific low frequencies (Table 5-23).
The physiological significance of these field-induced
effects is not established.

5.7.5 Unresolved Issues
In addition to the CNS-related  changes, amplitude-
modulated RF fields have been reported to alter an
immune response and a pancreatic tissue function.
These reports with diverse biological systems are
without apparent connection to each other except for
the physical agent causing the change. The biological
effects of frequency-modulated RF radiation, e.g., FM
radio signals, are not known. The reports cited above
of Merritt and  co-workers  indicate that pulsed
square-wave modulation may not cause calcium-ion
efflux,  whereas data from  the  Bawin et al.  and
Blackman  et al.  studies  show that sine wave
modulation is effective.

No report has yet described a mechanism of action in
sufficient detail to identify the conditions necessary
and sufficient to explain unequivocally calcium-ion
efflux in the  brain or the other biological changes
caused by  modulated RF fields. The response to
specific frequencies and intensities is unusual and at
present unexplained.  This response  to amplitude-
modulated RF radiation or to sub-ELF signals alone
may be a true field effect at a very low SAR and at
biologically relevant frequencies,  i.e., in the range of
frequencies  normally present  in the  EEG.  The
frequency-specific nature of the responses provides
evidence against heat as the underlying cause. The
unusual, multiple-intensity-range response challenges
standard dose-response analyses,  and by its  very
nature,  may  prohibit the invocation  of threshold
levels.

Other areas of unresolved issues include comparisons
of CW vs. PW microwaves under identical exposure
conditions. Such studies would help determine if the
differences seen by Wangemann and Cleary (1976)
were due to different exposure conditions or to the
irradiation parameters (CW or PW). There is also a
paucity of information on the effects of RF radiation at
different frequencies,  particularly at frequencies of
environmental  importance.  Studies at  different
frequencies would help to determine the reasons for
differences in effects at similar SARs. Such studies
might  help explain why Wangemann and Cleary
(1976)  reported serum chemistry changes in rabbits
at 0.8 W/kg (2450 MHz), and why Lovely et al. (1977)
reported no change in serum chemistry values in rats
at 1 W/kg (918 MHz).

There are also data such as those reported by Boggs
et al.  (1972) where  the  results from microwave
heating to a predetermined temperature are different
from those resulting  from the same temperature
produced by other  means of heating. Perhaps there
are differences in the uniformity of heating or in the
rate of heating which  would account for these
differences. In addition, a study by Oeficis etal. (1979)
reported elevated serum triglyceride and/Hipoprotein
levels in mice exposed to 2450 MHz at 1.5, 3.3, or 4
mW/cm2,  but  not at 1  mW/cm2.  Because the
exposures were  conducted in a  multimodal cavity,
SAR values  were not reported  and cannot be
predicted. If this study is repeated, particular attention
should be given to dosimetry. An  alternative is to
                       5-92

-------
 make  or report  dosimetric  measurements  in  the
 exposure system used.
 The reported effects on thyroid function at 3.75 W/kg
 for 60 h contrasted with no effect at 6.25 W/kg for 16
 h (Parker  1973)  suggests that the total amount of
 energy absorbed may also be an important considera-
 tion.  Additional  studies could define further  the
 relative importance of dose rate compared with total
 dose.
Table 5-23.    Summary of Studies Concerning Biological Effects of Low Frequency Modulation of RF-Radiation
Effects
         RF     Modulation  Intensity   Time       SAR
Species  (MHz)  (Hz)         (mW/cm2)  (min)       (W/kg)
                                                  Reference
Altered calcium-ion efflux in
brain tissue in vitro

  frequency specificity          Chicken   147  6-20

  influence of pH and lanthanum Chicken   450  16

  frequency and intensity       Chicken   147  16
    specificity

  intensity specificity and
    sample spacing

  theoretical analysis of
    sample spacing

  intensity specificity


  two intensity ranges


  theoretical analysis of
    RF dependence


  test of predictions of
    theoretical analyses
  no effect for pulse
    modulation

  no effect for pulse
    modulation


  change in calcium efflux
    kinetics in synaptosomes
  frequency and intensity
    specificity in cultured
    neuroblastoma cells
Altered calcium-ion efflux in
brain tissue in vivo

  no effect for pulse
    modulation

  change in efflux kinetics
    from awake animal

Changed EEC patterns


No change in behavior

Suppressed T-lymphocyte
  activity

Altered calcium ion efflux
  in pancreatic slices
  in vitro
Chicken   147   9, 16


Chicken   147   16


Chicken   450   16


Chicken    50   16


Chicken    50   16
          147
          450

Chicken   147   16


Rat      1000   16*, 32*


Rat      1000   16*
         2450   8*, 16*.
                32*
Rat
          450   16
Human    915   16
being
Rat      2060   8*, 16*.
                32*
Cat       450   16


Rabbit      1.2   15
           5.0   14

Chicken   450   3, 16

Mouse    450   16-100
Rat
147   16
                  1-2

                  0.75

                  0.83


                  0.83


                  0.83


                  0.1-1


                  1.5
                  3.6
                 0.37
                 0.49

                 0.5-15
                  1, 10
                  1


                  0.5
                 0.5-10

                 3
20

20

20


20


20


20


20


20



20


20


20



10

30
20

60
0.002*

0.0035*
0.0014*


0.0014*


0.0014


0.0005-
0.005*

0.0013
0.0035

-0.001
0.0006
0.0008
0.15-4.35

0.29-2.9
0.3
                                       0.05
Bawin etal. (1975)
Bawin etal. (1978)
Blackman et at. (1979)

Blackman et al. (1980a)

Joines ef al. (1981)

Sheppard et al. (1979)

Blackman etal. (1980b)

Joines and Blackman (1980);
Athey (1981); Joines and
Blackman (1981)
Blackman et al. (1981)

Shelton and Merritt (1981)

Merritt ef a/. (1982)


Lin-Liu and Adey (1982)

Duttaef a/. (1984)
                 500 V/m   120 x 6wk  —
                 500 V/m   120 x 6wk  —

                 1,5       23

                 1.5        120
0.12-2.4    Merritt etal. (1982)

0.29       Adey etal. (1982)

           Takashima etal. (1979)
          0.2. 1.0*   Sagan and Medici (1979)
          -         Lyle etal. (1983)
                                      60-150     <0.075    Albert et al. (1980)
*Est. SAR.
'Square wave.
                                                                                 5-55

-------
            cs and Mutagenesis
            r/na/i

            ie branch of biology that deals with the
            variation of organisms. The biochemical
            dity lies in the sequence of bases found
            : acids, deoxyribonucleic acid (DNA), and
v.., a ,c« v,c,*os) ribonucleic acid (RNA). All living cells
have the biochemical machinery to  detect  the
sequence of bases  in the DNA  and  to transcribe
sections of DNA information into similar sequences
of bases in RNA. The RNA molecules then move to
other locations inside the cell, where their information
is translated into various series of amino acids joined
together in sequences that were precisely defined in
the  original DNA molecule. These  specifically
arranged amino acids form proteins, some of which
are enzymes. Enzymes, in turn, catalyze biochemical
reactions that  ultimately  result in the growth and
propagation  of intact organisms. Hereditary (genetic)
material can be either nucleic acids alone, as found in
bacteria and viruses, or a  nucleic acid in association
with proteins, which form the chromosomes found in
more complex  organisms, including man.
Heat is a physical agent that can disrupt  genetic
material by causing the temperature to rise above
normal physiological  range. The  effect  of heat, or
temperature rise, has been studied in many biological
systems. As examples, heat has been shown to cause
physicochemical damage in isolated DNA preparations
(Lindahl andNyberg 1974; Ginoza era/. 1964; Ginoza
and Miller 1965) and in bacteria (Pellon et al. 1980);
chromosome changes in  Drosophila melanogaster
(Grell 1971) and in the Locusts migrator/a (Buss and
Henderson 1971), including a change from diploid to
haploid  in pollen from maize (Mathur et al.  1980);
enhanced sensitivity to other agents in mammalian
cell  cultures  (Ben-Hur et al.  1974)  and in D.
melanogaster (Mittler 1979); reduced fertility in rats
(Bowler 1972;  Fahim et al. 1975); and mutations in
bacteriophages (Bingham et al.  1976),  in bacteria
(Zamenhof and Greer 1958) and in D. melanogaster
(Muller and Altenburg 1919). Because absorbed RF
energy is  usually dissipated as heat, all reports of
genetic and mutagenic changes caused by exposure
to RF  radiation  must  be examined  closely to
determine whether temperature rise, or some other
mechanism,  is  the causative agent.

Mutations are  relatively permanent changes in the
hereditary material involving either  a physical
change  in chromosomal relations or a biochemical
change  in the  sequence of nucleic acid bases that
make up the genes. These changes can be passed on
to future generations of cells. Two different types can
be affected:  germ cells, which are egg or sperm or
their antecedent cells; and somatic cells, which form
all other tissues in the body. Mutations in germ cells
can  be passed on to future generations  of the
organism, whereas mutations  in somatic cells of an
organism may lead to impairment of organ function
and, in the extreme case, to cancer. Most mutations
are considered harmful  because they disturb the
biochemical processes necessary for the survival of
the organism and for the propagation of the species.
Evaluation of the mutagenic potential of RF radiation
has generally followed the procedures established for
testing the mutagenic activity of chemicalsfHollaender
1971). This evaluative scheme utilizes

• molecular systems, to detect changes  in the
   hereditary material;

• single-cell organisms,  to detect  changes in
   structure and function that are transmissible to
   the next generation;
• multicellular systems,  including  plants and
   animals, to detect changes in reproductive
   potential;

• infra-human  primates,  to detect  changes in
   reproductive potential.

The evaluation scheme  starts with simple, well-
defined genetic systems that enable rapid analysis
and identification of suspected agents  meriting
further evaluation.  The  more complex biological
systems used in further testing require  substantially
larger investments of time and resources.
In summary, the following conclusions may be drawn
from the review of the literature on the genetic and
mutagenic effects of  RF radiation. Experiments
designed to examine the genetic consequences of
exposure to RF radiation have been conducted with a
variety of test systems,  including isolated DNA,
prokaryotic and eukaryotic cells, and whole  animal
systems.  Radiation-induced effects on biochemical
properties of DNA, on chromosomal structure, on
mutation induction, and on reproductive capabilities
have been investigated. The reports demonstrate that
   (1) RF  irradiation  of low-to-moderate  intensity
      does not cause mutations in biological systems
      in which temperature is adequately controlled,
      and
   (2) the only  exposures  that are potentially
      mutagenic are those at high power densities of
      CW or PW radiation, i.e., exposures that result
      in  substantial  thermal loading  at  sensitive
      sites or result  in extremely high electric-field
      forces.
These conclusions must be qualified because  (a) only
specific frequencies or  small regions of  the RF
spectrum  have been experimentally examined, (b)
few detailed dose-response analyses have been
performed, (c) the influence of modulated radiation has
not been  sufficiently examined, an  (d) potential
synergistic  reactions  with  other environmental
stresses have not been adequately studied.
                       5-94

-------
5.8.1 Effects on Genetic Material of Cellular and
Subcellular Systems
5.8.1.1 Physical and Chemical DNA and
Chromosome Studies
Mechanisms for the absorption of RF radiation by
DNA molecules are treated elsewhere  in this
document (Sec. 3.2, RF-Field  Interactions with
Biological Systems, and  Sec.  5.1, Cellular and
Subcellular  Effects). This section addresses the
experimental studies that have examined changes in
genetic material induced  by RF radiation. Purified
DNA or DNA extracted  from the testes of exposed
mice has been examined to determine whether the
physical properties of the molecule can be altered by
the  exposure to RF radiation. Hamrick (1973)
examined the thermal-denaturation profiles of
aqueous solutions of isolated DNA exposed in vitro at
37°C to 2450-MHz (CW) radiation (SAR = 67 W/kg).
No changes were found. Even elevated temperatures
(to 50°C) produced by microwaves (1 h at an SAR
estimated to be 160 W/kg) did not cause a difference
in the thermal denaturation profile. (See Sec. 5.1 for
details.)
Varma and Traboulay  (1976) reported radiation-
induced changes in  thermal denaturation  profiles
(i.e.,  shift to lower temperature in the midpoint of the
transition curve [Tm] and reduction in the  maximum
hyperchromicity) as well  as changes in the base
composition of testicular  DNA extracted from
anesthetized mice whose testes were irradiated in the
near field. Ten  animals were exposed individually
either to 1.7-GHz radiation at 50 mW/cm2 for 30 min
(SAR estimated at 2.4 W/kg for testes alone) or at 10
mW/cm2 for 80 min (SAR estimated at 0.48 W/kg for
testes alone), or to 0.985-GHz radiation at  10
mW/cm2 for 80 min (SAR estimated at 0.26 W/kg for
testes alone). The animals exposed to 1.7-GHz fields
at 50 mW/cm2 or to 0.985-GHz fields at 10 mW/cm2
were given a 1 -day recovery period before they were
subjected to euthanasia, and the DNA was extracted.
The animals exposed to 1.7-GHzf ields at 10 mW/cm2
were used  in  another test for 8  weeks before
euthanasia, and the DNA was extracted. Identical
results are reported  for  the 1.7-GHz exposure at 50
mW/cm2 in another  publication (Varma and
Traboulay 1977). They conclude  "that biological
damage may be due to the combined effect of thermal
and  nonionizing radiation."

To evaluate the  relative  contributions of the thermal
effects  versus  nonionizing  radiation per  se.  the
results reported by Varma and Traboulay must be
examined with careful attention  to the  control
experiments and to the  potential size of the thermal
insult. A critical examination of the base composition
and  hyperchromicity data reveals that the authors'
explanations are not supported by the data. Although
the  higher percent  adenine/thymine in the DNA
extracted from exposed animals could be responsible
for the drop in Tm, the variability normally inherent
in  the  extraction procedures and in the base
composition measurements is not given. Without an
explicit indication of this variability, one cannot
conclude that the differences between control and
exposed samples are significant;  nor  can it be
concluded that  the  differences result from  the
radiation exposure directly,  rather than from  the
extraction procedures. Similarly, the potential size of
the thermal insult must be examined. The SAR values
for testicular exposures were crude estimates based
on incomplete information and appear too  low to
produce the types of damage described in the report.
For example, in an anesthetized animal exposed in
the near field to 1.7-GHz radiation at 50 mW/cm2
(SAR estimated  at 2.4 W/kg) and shielded with
Eccosorb except  for the testes, a 1  to  2°C  recta I-
temperature rise was recorded following exposure
(Varma and Traboulay 1977). Since no description of
the exact shielding technique was  provided,  it is
possible that more than just the testes of the animal
was  exposed, thus accounting  for  the  rectal
temperature increase. In another report (Varma and
Traboulay 1975), animals were exposed under  the
same conditions as described above, except that the
exposure time varied between 30 and 40 min, and the
testes were examined histologically. The authors
report that "the lumens were empty with complete
disintegration of spermatids, Sertoli  cells and  the
delicate  connective  tissue  which surrounds  the
seminiferous tubules." This type of damage can be
caused  by abnormal  temperature elevation  of  the
tissue (Muraca et al. 1976). Since anesthesia has
been shown to impair  thermoregulation (Cairnie etal.
1980b), it  is possible  that  sufficient energy was
deposited in the  anesthetized animals to cause  the
temperature elevation responsible for many of  the
changes that are attributed to microwave-specific
effects. (See Sec. 5.3.3,  Reproductive Effects—
Testes,  for further discussion  of the influence of
abnormal temperature rise on radiation-induced
damage in the testes.) Thus,  an alternative explanation
for the results described by these authors is that the
causative agent is elevated temperatures, produced
at sensitive sites  in the  tissue by  the radiation
exposures. This explanation  is consistent with  the
available data, and there is no  need to advocate an
unknown mechanism for radiation-induced damage.
Thus, an evaluation of the existing  evidence from
physical studies on DNA indicates that RF irradiation at
low-to-moderate intensities not accompanied by
temperature rise causes no changes in DNA bases,
the fundamental unit of the genetic code. However, if
substantial elevations of temperature occur  during
exposure, disruptions  in  the  pairing of the two
complementary strands, as well as  other damage,
may result.
Other researchers have used cytogenetic techniques
to examine some physical and chemical properties of
                                                                     5-95

-------
chromosomes  in  intact  cells  to  determine  if the
relationship of various parts of the genetic material is
altered by RF-radiation exposure. Huang etal. (1977)
stated that no RF-induced chromosomal aberrations
were found in white  blood  cells from Chinese
hamsters exposed to 2450-MHz radiation at power
densities up to 45 mW/cm2 (SAR = 20.7 W/kg) for 15
min/day on 5 consecutive days; however, the data
are presented in a manner which does not permit
independent statistical analysis. McRee et al. (1978)
in a preliminary report found no sister chromatid
exchanges in bone-marrow cells of mice exposed to
2450-MHz fields at 20 mW/cm2 (SAR = 15.4 W/kg),
for 8 h daily, 28 days total. Alam etal. (1978)showed,
in great detail, that chromosomal aberrations
occurred in a Chinese-hamster-cell line  (CHO-K1)
exposed 30 min  to  2450-MHz  radiation from a
diathermy applicator, but only if the temperature of
the culture was  allowed to  rise to 49°C during
exposure. These authors demonstrated that irradiation
of cell lines at high (> 200 mW/cm2) power densities
(SAR estimated at 360 W/kg)  would cause  no
detectable cytogenetic effects,  provided proper
temperature control was maintained. Thus, heating
seems  to account for  the  observed cytological
changes.
Authors of one detailed study used a frequency well
below the 1.7- to 2.45-GHz range. McLees et al.
(1972)  exposed  rats  treated to undergo  liver
regeneration either to 13.12-MHz CW(4.45 kVp-p/m)
fields or to PW (44.1  kVp-p/m, 200-/us pulse width,
50-Hz pulse repetition rate [PRR]) radiation for 28 to
44 h. They examined the effects of radiation on liver
cell mitotic activity (i.e., the percentage of cells in
mitosis and the number of chromosomal aberrations).
Rat liver cells may be very sensitive to exogenous
stresses during this regenerative process because
they are normally nondividing cells  in  an  intact
animal unless challenged to divide in situ. However,
the authors found no RF-induced alterations in
chromosomal  morphology (SAR estimated at 1.2 to
1.3 W/kg). Their results indicate that no cytogenetic
changes would  be expected in the  range of
frequencies studied for low-intensity exposures.
Thus, although there are reports indicating that
exposure to RF radiation can cause  cytogenetic
changes,  these changes  appear to result  from
radiation-induced  elevations of temperature.
McRee et al. (1981) have recently published the paper
that was reviewed when in preliminary format
(McRee et al. 1978). Twelve 10-we.ek-old female mice
(CD-1 strain) were exposed to 2450-MHz CWfields at
20  mW/cm2 (SAR = 21 W/kg) for 8 h daily (4 h in
morning, 1 h delay, 4 h in afternoon) each  day for 28
days. Exposure was dorsal. Immediately after the last
exposure, procedures were initiated that  produced,
19  h later, labeled bone marrow cells that-were
processed to  give a measure of sister chromatid
exchanges. Upon  analysis,  no statistical differences
were detected  in  the mean  number of sister
chromatid exchanges per cell between the exposed
and either of the unexposed control groups (sham and
cage control). Based on the percentage of mitotic cells
in each group, the microwave treatment did not have
an effect  on the rate of cell proliferation in bone
marrow of mice. These results are consistent with
those given in the preliminary report.

5.8.1.2 Biological Studies of DNA and
Chromosomes
Bacteria have been used to study the mutagenic
potential of RF radiation, because the single-cell
system is simple, easy to culture, quick to test, and
relatively sensitive to the action of mutagenic agents.
By using  the bacterial system  and the biological
amplification it  provides for  any DNA change,
molecular biologists have been able to decipher the
genetic code and to identify a change in as few as one
DNA subunit out of 10'° to 1012 subunits. In contrast,
biologists using standard physical techniques such as
DNA melting curves usually can detect changes in no
better than 0.1  percent of the DNA. Blackman et al.
(1976) exposed growing cultures of the bacterium
Escherichia coli either to 1.7- or to 2.45-GHz (CW)
radiation for 3 to 4 h. Exposure at 1.7 GHz was in the
near field  at 88 V^s/m or ~ 250 Vp-p/m (SAR = 3
W/kg). The 2.45-GHz exposures were in the far field
at either 10 or 50 mW/cm2 (SAR = 15 or 70 W/kg,
respectively). Although exposure of growing cells
provided enhanced sensitivity to mutagenic agents,
no mutagenic activity was detected. A positive
control,  ultraviolet (UV) light,  caused mutations and
was used to demonstrate the  sensitivity of the assay
method. Dutta et al.  (1979a)  exposed  growing
cultures of various bacterial strains of Salmonella
typhimurium, commonly used in the Ames testing
procedures to detect chemical mutagens (Ames et al.
1975), to 2.45-GHz (CW) radiation for 90 min at 20
mW/cm2 (SAR =40 W/kg) and to 8.6-, 8.8-, 9.0-, 9.2-,
9.4-, and 9.6-GHz (PW) radiation (1 -yus pulse width, 1 -
kHz  PRR)  at 10  and 45  mW/cm2 average power
densities,  and  10,000 and 45,000 mW/cm2  peak
power densities.  (The  SAR  at  45 mW/cm2 was
estimated at 80 W/kg.) No mutagenic activity was
observed under any of these exposure conditions.
Another approach with bacterial systems is to test for
radiation-induced alterations in  genetic processes,
including cell death. Since most mutations are
detrimental, they might be detected indirectly by this
method. Corelli et al. (1977) exposed cultures of E.
coli to microwaves at frequencies swept between 2.6
and 4.0 GHz for  8 h (SAR =  19 W/kg). Although at
26°C these cultures were probably growing slowly,
no  change was noted  in the number of colony-
forming units  (CFUs) in the cultures following
irradiation, indicating no detectable  lethal  events
because  of the  exposure. These workers also
examined the  infrared (IR) spectrum of these cells
                       5-96

-------
when exposed to 3.2-GHz radiation for 11 to 12 h
(SAR is either  21 or 16  W/kg). There  was  no
observable effect on the molecular or conformational
structure of these cells, in contrast to results obtained
with a positive control, ionizing radiation. Two strains
of E. coli, one deficient in an enzyme needed to repair
damaged  DNA, were  tested for survival following
microwave  exposure (Outta et al.  1979b). The
exposure conditions were 8.6-GHz (PW) radiation (1 -
/us pulse width, 1 -kHz PRR, SAR  = 12 W/kg) for 1, 2,
4, or 7 h. There was no significant change  in the
relative growth patterns of these strains that could be
attributed to microwave-induced ONA damage that
was repaired in one  strain, but not in the  other.
Blackman eta/. (1975) exposed a  different strain of f.
coli in  log phase (actively dividing) and in lag phase
(undergoing metabolic activities preparatory to
division) at 32°C for 4 h to 2.45-GHz radiation at
0.005,  0.5,  5.0,  or 50 mW/cm2. (At 50 mW/cm2
power density the SAR = 75  W/kg.) Additional
experiments were conducted at 5 mW/cm2 and 25°C
to test for the influence of cold  stress, and in two-
culture media at 30 and 35°C to compare the relative
influence of a rich  medium with  a minimal medium;
the latter required greater utilization of the genetic
apparatus of the cell for growth  to occur. Except for
enhanced growth at 50  mW/cm2, which was
attributed to slight  temperature rises in the exposed
cultures, no change was found in the colony-forming
ability of the cultures due to the  exposure.

Few researchers have used single-cell systems more
complex than  bacteria specifically  to look for RF-
radiation-induced mutagenesis. Dutta et al. (1979a)
exposed a diploid strain of the yeast Saccharomyces
cerevisiae, a primitive eukaryote, to 2.45-GHz (CW)
radiation for 2 h at 20 mW/cm2 (SAR = 40 W/kg).
They found essentially no change in the number of
mutations at either of two  loci  affecting the
nutritional requirements for adenine or tryptophan.
These investigators conducted additional tests at 8.5-,
8.6-, 8.8-, 9.0-, 9.2-, 9.4-, and 9.6-GHz(PW)radiation
(1-/JS pulse  width, 1-kHz PRR) for  2  h  at  average
power densities of 1, 5, 8.9, 10, 15, 30, 35, 40, or 45
mW/cm2. Although no measurements of SAR are
cited, so that comparisons with  CW exposures are
difficult, the highest power density was reported to
raise the temperature of the culture by 12°C, which
would mean substantial absorption of the radiation.
(At 45-mW/cm2 power density the estimated SAR =
80  W/kg.) In no case did the exposures cause  a
change in the frequency of genetic  events, altering
the requirements for either adenine or tryptophan, in
the treated population  as compared with the control
population. Saccharomyces cerevisiae was also used
in two studies by Dardalhon and co-workers. In the first
report,  Dardalhon et al. (1979) exposed two haploid
and one diploid strain of the yeast at 20°C in the near
field to either 70.5- or 73-GHz CW fields at power
densities up to 60 mW/cm2 (SAR estimated  at 17
 W/kg) for durations up to 3 h. In the second report,
 Dardalhon et al. (1980) exposed a diploid strain of S.
 cerevisiae at 20°C in the near field of 9.4-GHz (CW)
 radiation (SAR estimated at < 2 W/kg) for 1 to 5 h, or
 in the near field of 17-GHz (CW) radiation at either of
 two power densities (SARs estimated  at 28 or < 6
 W/kg, respectively) for various times to  24 h.  No
 significant changes attributable  to the irradiation
 were observed in the  percent  survival, in  the
 induction of cytoplasmic "petite"  mutations, in the
 induction of mitotic recombinations, or in sporulation.
 Thus, no changes  were detected in complex single-
 cell systems used to examine directly the mutagenic
 potential of microwaves at frequencies of 9.4, 17,
 70.5, or 73 GHz (CW) and SARs of < 2 to  28 W/kg.

 Some work has been done with bacteria and yeast
 cultures  that compares the lethal and mutagenic
 effects  of microwaves with those  induced  by
 conventional  heating. Dutta et al.  (1980) examined
 the responses of various strains of the bacteria S.
 typhimurium and E. coli and those of a diploid strain
 of the yeast S.  cerevisiae, exposed to 8.6-, 8.8-, or
 9.0-GHz (PW) radiation (1 -kHz PRR, 1 -A»S pulse width)
 at average power  densities to 45 mW/cm2 (SAR
 estimated maximum at 80 W/kg). The bacteria were
 exposed 90 min at an ambient temperature of 37°C,
 whereas  the yeast were exposed  2 h at 30°C. The
 comparison  of these  irradiation  treatments with
 those obtained from treatments at elevated tempera-
 tures produced by conventional heating indicated that
 conventional heating could produce cellular damage
 leading to reduced survival in a manner similar to the
 changes caused by microwave-induced elevations of
 temperature (to 10°C above normal growth tempera-
 tures). Nd mutational events occurred in bacteria to
 42°C, or  in the yeast to 40°C, however, at 45°C, a
 slight increase in mutational events occurred in the
 yeast, and at 47°C,  in S. typhimurium. To obtain some
 indication of  the temperature rise associated with
 exposure to RF radiation, Dardalhon et al. (1979)
 developed a response curve for zygote formation, the
 production of a diploid cell from the union of two
 haploid cells, as a function of temperature.  Exposure
 to 70.5-GHz CW fields at 60 mW/cm2 (SAR estimated
 at 17 W/kg) produced a response in zygote formation
 that could result from a 3°C temperature  rise. This
 result demonstrated that biological systems could be
 used as sensitive indicators of temperature change
 under exposure conditions in which temperature can
not be readily measured. In another study, Dardalhon
 et al. (1980) exposed a diploid strain of S. cerevisiae
 for 3 h at 52°C. (Normal growth temperature is 30°C.)
 Large changes were observed In the percent survival,
 in the induction of cytoplasmic mutations, and in the
 induction of mitotic  recombinations after 3  h,
 compared to no changes in these end points during
 microwave exposure at 20°C to 9.4-GHz or 17-GHz
 fields (SAR  estimated  at <  28  W/kg), described
 above. Both  groups of  authors conclude  that  care
                                                                      5-97

-------
must be exercised in evaluating genetic changes in
microbial assay systems when elevated temperatures
accompany microwave exposure.
Since 1980, three groups have published reports in
this subject area. Dutta extended his previous work
(Dutta et at. 1979b, 1980) with two strains of E. coli,
one deficient in an enzyme needed to repair damaged
DNA, to compare bacterial growth under high-
intensity microwaves and conventional heating
(Hossain and Dutta 1982). Both bacterial strains were
grown for 2. 5,10, or 15 h either at 35.2 °C or42.2°C,
or exposed to 8.8-GHz  (PW) radiation (1-/as pulse
width, 1-kHz PRR, SAR = 12 W/kg) at 35.2°C (final
temperature of the sample was 42.2°C) and assayed
for  growth,  as  colony-forming  units.  The results
demonstrated that although the microwave exposure
caused  a change  in the bacterial growth when
compared with the 35.2°C control, it had no effect on
the growth of E. coli beyond that associated with the
radiation-induced temperature increase.
Dardalhon and co-workers (1981) exposed the yeast
S. cerevisiae, and both normal and repair-deficient
strains of the bacterium  E. coli, to microwave fields
and tested for survival and induction of mutations. In
diploid yeast, they tested  for sporulation capacity and
gene segregation during meiosis.  They  found no
evidence for altered survival in any of the bacterial
strains following near-field  exposure for  30 min to
9.4-GHz CW fields (SAR =  23 W/kg),  17-GHz CW
fields (up to SAR = 28 W/kg), or 70- to 75-GHz CW
fields (SAR = 9 W/kg), nor any evidence of mutation
induction following exposure for 30 min to 70- to 75-
GHz CW fields (SAR = 28 W/kg) or 17-GHz CW fields
(SAR = 6 W/kg). Exposure of haploid yeast strains to
70- to 75-GHz CW fields (SAR = 28 W/kg),  17-GHz
CW fields (up to SAR  =  28 W/kg), or 9.4-GHz CW
fields (SAR  = 23  W/kg) produced no detectable
changes in  survival,  induction of cytoplasmic
mutations,  induction of reversions, or genetically
altered colonies. Similarly, exposure of  diploid yeast
to 9.5-GHz CW (SAR = 23 W/kg) or 17-GHz CW (SAR
= 28 W/kg) fields for 48 h  produced no significant
alterations in the sporulation capacity nor in meiotic
gene segregation. Thus  with  both prokaryotic and
eukaryotic cells these investigators did not detect any
significant genetic effects from exposure to microwave
radiation at selected frequencies between 9.4 and 75
GHz.

In a  series  of  three  detailed  papers, Swedish
researchers studied the possible genetic effects in
bacteria  of  exposure  to high-frequency RF fields
(Anderstam  et  al.  1983;  Ehrenberg et al. 1983;
Hamnerius 1983). Tests  for forward mutations were
performed in two E.  coli strains and three S.
typhimurium strains, for backward mutations in three
E. coli strains and two S. typhimurium strains, and for
prophage induction in  three E.  coli  strains. The
bacterial samples were  exposed in the far field at
37°C to either 3.07-GHz pulsed fields (500 Hz PRR, 2-
fjs pulse width) at average power densities up to 210
mW/cm2 (SAR = 100 W/kg), or 2.45-GHz CW fields
(which were essentially  100-percent amplitude
modulated at 100 Hz) at power densities up to 170
mW/cm2 (SAR = 80 W/kg). Bacterial samples were
exposed similarly to 27.12-MHz CW field components:
electric fields at 72 V/m in the sample generated
between  two parallel plates (SAR = 3 W/kg), and
magnetic fields at 20 A/m in the sample generated by
a Helmholtz coil  (SAR = 20 W/kg). Exposures were
conducted for 1 to 7 h, followed by measurements of
cellular growth for viability, mutation or prophage
induction, or RNA synthesis. The battery of bacterial
test systems used in this detailed study was designed
to provide a measure of resolving power and thus to
place an upper confidence limit on the number of
mutagenic events that would be probable for a given
result. The authors state that the pooled mutation
frequency did not differ from that of the controls,
although a weak prophage-induction capability and an
ability to  modify the response  to UV-radiation (DNA
repair) were noted. The most prevalent finding was a
stimulation of growth, especially for certain strains
and as the cultures were entering stationary phase.
This result was also exhibited  in the RNA synthesis
studies. Although the authors discounted at length
the  influence of temperature difference as the
underlying cause of the growth stimulation, because
of the high SARs that were used, a 1 °C temperature
gradient was known  to exist between the bacterial
samples and the cooling bath. Other workers, notably
Livingston  er a/.  (1979),  have emphasized the
extreme caution that must be exercised in interpreting
data obtained under conditons where local hot spots
could occur. To support  their contention that
temperature change contributed in only a minor way
to the observed growth stimulation, the authors could
have conducted  exposures at temperatures below
and  above the optimum  growth temperature.  If
temperature increases were a major factor in the
field-induced growth-rate changes, one would expect
a reversal in the response between the lower and
higher temperatures.  The complex biochemical
changes  that occur as  the  cells transfer from
exponential to stationary growth phase adds further
uncertainty to any evaluation of altered growth
patterns. Nevertheless, with a comprehensive battery
of test systems,  the authors  were  unable  to
demonstrate any mutagenic activity due to exposures
to 27.12-MHz CW, 2.45-GHz CW, or 3.07-GHz pulsed
fields.

In another study to determine temperature change
during RF exposure, Dardalhon et al. (1979) exposed
one  diploid and two haploid strains of the yeast S.
cerevisiae, at 20°C in the near field to either 70.5 or
73-GHz CW fields at  power densities up to 60
mW/cm2 (est. SAR = 17 W/kg) for up to  3 h. No
changes were observed in cell survival, or in induction
                       5-98

-------
of mitotic recombination or of cytoplasmic mutations.
To obtain some  indication of the temperature rise
associated  with these  exposures, the authors
developed a biological indicator, namely a response
curve for zygote formation  as a function  of
temperature. Exposure to 70.5-GHz CW fields at 60
mW/cm2 (est. SAR = 17 W/kg) produced a response
that could result from a 3°C temperature  rise. This
result demonstrates the usefulness of biological
systems as sensitive  indicators of temperature
change under exposure conditions where temperature
can not be readily measured.
All these reports are consistent with the conclusion
reached above that no well-defined genetic effects
have  been  reported occurring from exposure to
microwave radiation that could not be attributed to
temperature rise. However, the work of Grundler and
Keilmann (1983)  identifies  possible  frequency-
specific alterations in the growth of yeast (Sec. 5.8.3).
The  mechanism for  this  response has  not been
identified.

5.8.2 Effects on Genetic Material of Higher-Order
Biological Systems

5.8.2.1 Plants
Plant  systems are sensitive to mutagenic agents and
thus can serve to identify conditions that should be
examined more  carefully in mammalian  systems.
Although several studies  are reported  in the
literature, none  has  sufficiently described  experi-
mental conditions to allow independent evaluation of
the work. Thus,  useful  mutagenic testing of RF
radiation with respect to plant systems has yet to be
accomplished and documented.
5.8.2.2 Invertebrates
Several investigators have studied D. melanogaster.
a standard model for testing mutations induced by RF
radiation.  No mutagenic  effects  were reported by
Hamnerius et al. (1979) in embryos  exposed 6 h in
water at 24.5°C to 2450-MHz (CW) fields  at 100
W/kg. Similar negative effects were reported by Pay
et al. (1972), who exposed adult males to 2450-MHz
fields for 45 min at 4600, 5900, or 6500  mW/cm2
(SARs estimated at 140,  190,  and  210  W/kg,
respectively) and mated each of the surviving males
individually with two  virgin females  during 15
consecutive  days. No changes were observed in the
generation time or sex-ratio pattern of the offspring.
These offspring were then mated and observed for
possible  sex-linked  lethal  mutations  in their
offspring. No  such mutations were found  at
frequencies  > 1 percent, a detection limit  based on
the small number of chromosomes (< 800) actually
evaluated. An additional caveat by the authors was
that the most sensitive stages to detect recessive sex-
linked lethal mutations (late spermatocytes and early
meiosis) were not tested in their study, because few
 sperm were still in meiosis at the time of exposure.
 Mittler (1976), who exposed adult males from various
 strains of D. melanogaster to 29-MHz (CW) fields at
 600 Vrms/m (SAR roughly estimated at 0.024 W/kg
 and to 146-MHz (CW) fields at  62.5 Vms/m  (SAR
 roughly estimated at 0.015 W/kg) for 12 h, mated
 them with virgin females for 12 h every 2  days in
 production of  4 or 5 broods. In these experiments,
 brood  4 was produced from sperm irradiated "in or
 about  meiosis." No mutations were induced by these
 treatments as  evidenced by the lack of chromosome
 loss, nondisjunction, or sex-linked recessive lethals.
 In addition, Mittler (1977) observed  no mutagenic
 effects (recessive  lethals) when  exposing adult
 females  to 98.5-MHz frequency-modulated fields
 (composed of standard commercial broadcast audio
 frequencies) at 0.3 Vm,,/m (SAR estimated at 0.0004
 W/kg), 134 h  per week for 32 weeks. Although the
 results  with  D. melanogaster are  difficult  to
 extrapolate to  the human condition, this test system
 as a qualitative index could not detect any mutagenic
 alterations by  RF radiation over a wide range  of
 frequencies.

 5.8.2.3 Vertebrates

 Experiments with vertebrates address the impact of
the genetic effects of RF radiation most directly be-
cause of the general biological similarities between man
and  other vertebrates.  However, there are very few
such experiments directed at this subject. Three
studies have been reported that are directly focused
on the mutagenic potential of  microwaves  in  mice
and  rats.  Varma et al. (1976) used the dominant-
lethal test to investigate the effects of 2.45-GHz (CW)
radiation. The  testes of  anesthetized mice were
treated either by a  single 10-min exposure  at 100
mW/cm2 (SAR estimated at 11.4 W/kg), or by three
exposures of 10 min each at 2-h intervals on the same
day at 50 mW/cm2 (SAR estimated at 5.7 WAg), or by
four exposures of 10 min each at 3-day intervals over 2
weeks  at 50 mW/cm2  (SAR estimated at 5.7 W/kg).
Varma  and Traboulay (1976) in a similar study exposed
the testes of anesthetized mice to a single dose of 1.7-
GHz radiation,  at  10  mW/cm2  for 80  min (SAR
estimated at 0.48 W/kg) or at 50 mW/cm2 for 30 min
(SAR estimated at 2.4 W/kg). Following a 24-h recovery
period,  the males were bred with virgin females—one
group of females per week for 6 to 8 weeks. In the 2.45-
GHz study, the authors used the week-6-group results
for comparison; they found a higher mutagenicity index
in the results of the 100-mW/cm2 group bred at week 1,
and in results of the 50-mW/cm2 group exposed three
times in one day and  bred at week 4. The authors
concluded that  a single intense exposure or multiple
exposures  during one day induced  a  number of
significant mutations in the  mice. The results of the 1.7-
GHz studies at 50 mW/cm2 for 30 min indicated a
radiation-induced increase in infertility,  in  pre-
implantation losses, and in the mutagenicity index of the
                                                                      5-99

-------
groups bred at 3, 4, 5, and 6 weeks. In addition, for the
group exposed to 1.7-GHz fields at 10 mW/cm2 for 80
min, an increase was seen in the mutagenicity index of
the groups bred at 1,  2, 3, and 6 weeks. The authors
concluded that the "biological damage may be due to the
combined effect of thermal and nonionizing radiation."
Both these  reports should be evaluated on the basis
described in Sec. 5.8.1.1, Physical and Chemical DNA
and Chromosome Studies, for the physical and chemical
studies reported by these authors. Basically, the reports
do not describe the exposure conditions in sufficient
detail to allow an unequivocal evaluation of the results.
Also, it is not possible to determine the extent of the
biological damage  caused by elevated temperature in
either of these reports, especially because anesthetized
mice  lack  an effective temperature-regulatuion
capability (Cairnie et a/.  1980). However, if the
histological  damage found  in  the testes of animals
exposed under similar  conditions (Varma andTraboulay
1975} is reflected in damage to sperm, it could account
for  many if not all changes in the mutagenic index,
fertility, and pre-implantation  loss reported by those
authors in later studies.

The conclusion that radiation in the frequency range
of 1.7 to 2.45 GHz could cause mutagenic changes in
mice was challenged in a detailed study by Berman et-
al. (1980). Unanesthetized male rats were exposed to
2.45-GHz (CW) radiation under three treatment
regimens: 4 h/day from day 6 of gestation to 90 days
of age at 5 mW/cm2 (SAR varied from 4.7 W/kg to
somewhat less than  0.9 W/kg at day  90, because
growth of the animals  changed their energy
absorption efficiencies); 5 h/day for 5 days beginning
on day 90 at 10 mW/cm2 (SAR estimated at 2 W/kg);
and 4 h/day, 5 days/week, for 4 weeks beginning on
day 90 at 28 mW/cm2 (SAR estimated at 5.6 W/kg).
During selected weekly periods after -treatment, the
exposed males were  bred to untreated  females that
were examined in  late pregnancy by the  dominant-
lethal assay.  No significant germ-cell mutagenesis
was detected under  any treatment condition, even
though significant  increases in rectal and testicular
temperature were observed during the 28-mW/cm2
exposure and were associated with a  concomitant
decrease in incidence of pregnancy during some of
the breeding periods, which indicates temporary
sterility. In  addition,  these authors reexamined the
data  of  Varma and concluded they had been
interpreted incorrectly because the effects of litter
size on fetal mortality had been ignored  and the
differences between treated and control groups had
been overemphasized when the control values were
not representative of normal values. Berman et al.
(1980) concluded that "it  still  remains to  be
demonstrated that microwaves, even at near-lethal
doses, can cause a dominant lethal mutagenic effect
in the  mouse or the rat."  Thus, well-designed
experiments with biological systems having complex
genomes similar  to  man's have demonstrated  no
mutagenic effects from low- to moderate-intensity RF
radiation  when  the  temperature of the biological
system is maintained in the physiological range. Most
of the reported  changes  are  consistent with the
effects expected from elevations  in  temperature,
caused by  CW  exposure  to high-intensity RF
radiation.  The other purported changes cited in this
section may have resulted from incomplete experi-
mental design.

Recently, dominant lethality was tested by Saunders
et al. (1983), who exposed the lower half of adult male
C3H mice one time  for 30  min in a waveguide at
2450-MHz CW fields (average SAR = 43 W/kg). The
exposed males were  then mated with pairs of virgin
females each week for 10 weeks. The uterine contents
of the females were examined either on day 14 or day
18. The post-implantation survival, defined  as the
number of living implants divided by the total number
of implants, was used as a measure of the dominant
lethality. There were no significant changes in the
dominant  lethal test for any of the 10 mating periods,
which indicated  that no stage of spermatogenesis
was preferentially sensitive to the RF treatment. This
result is in agreement with the other results cited in
this section. Other measures made by Saunders eta/.
are distussed  in  Sec. 5.3.3.


5.8.3 Unresolved Issues
Heating, which raises the temperature of biological
samples above normal physiological ranges and may
result in genetic and mutagenic changes, is a well-
known result of exposure to high-intensity RF
radiation. The choice of exposure conditions to study
the genetic and mutagenic effects of moderate to low
intensity   radiation is arbitrary because no well-
defined  mechanism  of RF  interaction other than
heating has been developed to address this biological
problem. The experimenter must select the frequency,
intensity,  and  duration of exposure, as well  as the
type and characteristics of modulation. To compound
the problem further, the number of biological systems
that could be selected for study is limitless. Several
guides can be developed from recent experimental
findings to help delineate a future experimental plan.
Although the conclusion has been stated above that
no strong evidence  exists in  the  cited reports  to
demonstrate activity  for low-intensity RF radiation
per se,  it should be qualified by the following
observations:

1. These experiments are limited to small portions of
   the RF region of interest, i.e., 0.5 MHz to 100 GHz.
   This 106-Hz frequency  range is so large  that if
   detrimental effects occurred over very  narrow
   frequency ranges, they might have gone unde-
   tected.

2. Many of  the  experiments were conducted with
   CW radiation. If effects exist, they may occur with
                       5-100

-------
   amplitude-, frequency-, or PW-modulated radiation
   at frequencies associated with ongoing biological
   or biochemical processes. An exposure to a short-
   duration, high-intensity PW radiation (radar-like)
   may not produce the same biological response as a
   CW exposure of the same average power density.
   Thus, the concept of averaging the intensity of PW
   radiation over time to  determine  an  effective
   intensity may not be generally valid.

3. The  general approach  in most experimental
   studies has been to assume that the higher the
   radiation intensity, the more  pronounced the
   biological effect.  It is  possible that  several
   mechanisms are operative during radiation
   exposure and that mechanisms with thresholds at
   relatively high intensities (e.g., heating) mask the
   effects, which may be caused by the more subtle
   mechanisms.
4. RF radiation may enhance the  effects of known
   mutagenic agents as a co-stressor. In environ-
   mentally relevant  situations,  biological systems
   are exposed simultaneously to many agents, both
   chemical and physical.
5. Standard test procedures that  have been
   optimized to identify  ionizing  and chemical
   mutagens are  not necessarily applicable  to
   examining the interaction  of RF radiation  with
   complex biological structures and genetic
   material. No  test system should be overlooked as
   potentially sensitive for detection of mutagenic
   activity.  For  example, plant systems have been
   shown to be extremely sensitive indicators  of
   environmental pollution.

Recent experimental developments support some of
the suggestions made above. Possible frequency-
specific responses were reported by Smolyanskaya
and Vilenskaya (1973), who described the induction of
colicin  in E. coli by extremely  high-frequency
radiation, near 37 GHz, at power densities of 0.001,
0.01,0.1, or 1.0mW/cm2for0.5,1, and 2 h. Colicin is
a protein usually not made by the cell. Its induction,
though not well understood, requires the transcription
of a new portion of the DNA molecule. This induction
demonstrates that  specific  microwave frequencies
can induce changes in the genetic processes of cells.
More recently,  Grundler et al. (1977) described
frequency-dependent growth responses in the yeast
S. cerevisiae  exposed for periods  to 11  h to
microwave frequencies between 41 and 42 GHz at 1
to 3 mW/cm2 (average SAR estimated at 4 to 11
W/kg).  (See Sec. 5.1, Cellular and  Subcellular
Effects, for details.) Although this work is controver-
sial,  it does support the frequency-specific-response
concept.

The  importance of examining  the effects of RF
radiation as a  co-stressor with  other agents  is
emphasized in the report by Dardalhon  et al. (1980)
 that was discussed earlier with respect to exposure
 either to  heat or to microwave radiation.  These
 authors exposed  haploid and  diploid  strains of S.
 cerevisiae tor 1 h to 17-GHz(CW) radiation, near field,
 at 50 mW/cm2 (SAR estimated at < 6 W/kg) and
 20°C, followed by exposure to various doses of UV
 light  (254  nm), which is known to cause damage
 directly to the DNA. Both haploid and diploid strains
 were affected. The exposure combination caused a
 tendency—but apparently not a statistically signifi-
 cant change—toward diminished survival, increased
 mitotic  recombination, and increased  cytoplasmic
 "petite" mutations, when compared with the effects
 caused by  exposure to the same doses of UV light
 alone. This possible microwave-radiation enhance-
 ment of effects due  to UV-light exposure would be
 unusual, because it was observed only at high doses
 of UV light,  where  additional  repair  systems are
 presumably operative. The authors found similar
 trends by  replacing  microwave-radiation exposure
 with temperature elevation to 46°C before treatment
 with  UV  light.  No  influence from  exposure to
 microwave radiation,- followed by UV  light, was
 observed for 17-GHz fields at 2.5 mW/cm2, or for 9.4-
 GHz fields  at  5 mW/cm2. Because the microwave-
 radiation  treatment caused only a  1°C global
 temperature rise in the sample that was 10°C below
 its normal growth temperature, the authors reasoned
 that if temperature rise is the basic agent enhancing
 the UV-light-induced response, it must be occurring
 selectively at specific sites within the biological
 system other than on the DNA itself, so that change
 is perhaps  caused in metabolic processes or  in the
 structural integrity of cytoplasm, or of membranes. In
 this case, a differential sensitivity among the various
 repair systems may be induced by the UV  light.
 Similar investigations  might reveal  biological
 processes, including repair of damage or detoxification
 of chemicals,  that may be  particularly sensitive to
 microwave-radiation exposure.

 In another area, radiation-induced efflux of calcium
 ions from in vitro brain tissues has been shown to
depend on the modulation frequency of a carrier wave
(Bawin et  al.  1975;  Blackman et  al.  1979). The
effective modulation frequencies were  in the same
frequency  range as the natural biological rhythms
associated with the EEG in the intact animal, and thus
the radiation  may have coupled with an  existing
oscillatory system. Further studies by Blackman et al.
(1979; 1980a,b) and Sheppard et al. (1979) indicated
that higher and lower intensities of radiation can lead
to the disappearance of the effect.  This result
suggests that examining biological processes only at
a single  power  density  may  provide misleading
information.

Three additional,  incomplete  reports have  been
included in this section. These studies are included
either because they present positive findings at
                                                                    5-101

-------
frequency ranges or intensities seldom studied or
because of a strong claim  by the  authors. Heller
(1970) and Mickey and Koerting (1970)  reported
chromosomal aberrations in cultured Chinese-
hamster lung cells exposed for 30 min to 19- or 21-
MHz fields, although no changes were observed after
exposure to 15- or 25-MHz (PW) fields (100-Hz PRR,
50-fjs pulse width) at field intensities to 300 kVp-p/m.
It is very likely either that the extremely high peak
voltages (up to 300 kVp-p/m) were producing intense,
rapid heating within the cells, or that the field per se
was  causing  major  stresses within the system.
These  reports  provide  insufficient  descriptions to
allow an estimate of the SAR values, and thus these
results cannot be  compared with those found by
others,. It  is possible that fields at  extremely high
intensities can account for some cytogenetic changes.
Manikowska et a/. (1979) reported a dose-indepen-
dent increase in chromosomal translocations and in
chromosomal  pairs  remaining as univalents at
Metaphase I in  the sperm cells of mice exposed 1
h/day, 5 days/week, for 2 weeks to 9.4-GHz (PW)
radiation at 200-, 1000-, 2000-, or 20,000-mW/cm2
peak power densities (0.5-//S pulse width; 1 -kHz PRR;
0.1-,  0.5-, 1-,  and  10-mW/cm2 average  power
densities, respectively). (Average SAR is roughly
estimated at 5 W/kg, and peak SAR roughly at 9000
W/kg  for the  highest power  density  used.)  The
authors do not describe the  relation of the animals'
testes  to the incident field, which  thus raises the
question of the actual quantities of energy coupled to
the target cells at 9.4 GHz, where tissues exhibit large
attenuation coefficients.  A description  of  the
environmental  conditions  during exposure is also
absent. These omissions prevent  any critical
assessment of the results. Furthermore, because of
the small number of animals used in the study, the
authors themselves state that "the findings...
obviously need  confirmation on larger numbers of
animals."

In  another study,  based on bacteria  exposed to
extremely high temperatures, an incomplete analysis
of  the distribution of temperatures  within  the
samples led to an unjustified conclusion. Blevins et
al. (1980), who exposed  several strains of S.
typhimurium commonly used in the Ames testing
procedures (Ames et al. 1975) to 2.45-GHz radiation
in a microwave oven at a calculated power density of
5100 mW/cm2 for periods of 2 to 23  s, examined the
cultures for  lethality and for  mutation induction.
These  exposures caused extremely large elevations
of temperature.  Corollary heating experiments were
performed in  high-temperature water baths to
determine the extent to which temperature change
alone contributed to lethality and mutations. Their
crnclusion that  microwave  radiation  is a  potent
mutagen, because it caused mutations  in excess of
those expected from the radiation-induced tempera-
ture rise, is not supported by their data. The authors
failed to demonstrate  that the  uniformity of
microwave heating was duplicated by the water bath
experiments. The authors acknowledge that "differ-
ences in the kinetics of water bath and microwave
heating  are possible" but  do not evaluate this
possibility further. Because of the larger temperature
increases in their  culture systems—apparently as
large as 46°C  in 14 s—more definitive temperature
distribution work must be  done  to  establish
accurately the  contribution that temperature change
makes to induction of  mutations  before additional
mutagenic properties are assigned to microwaves.
Blevins et al. (1980) reported using a power density
500 times greater than the current U.S. occupational
guidelines. Because the apparent mechanism (i.e.,
heating) is  based on  such high, nonphysiologic
temperatures, .the authors'  conclusions  are  not
applicable to the other cited studies, which have used
2.45-GHz radiation. However, the study may support
a reevaluation of the concept that brief, high-intensity
exposures can  be averaged over a longer time period
to define the  average  intensity.  For example,  an
exposure to 5100 mW/cm2 for 2 s is 28 mW/cm2 if
averaged  over 6 min,  and an exposure for 14 s
becomes 200 mW/cm2 if averaged over 6 min. The
Blevins et al. study would also support the concept
that'high-intensity  pulses of microwaves may affect
biological systems differently  from convection or
conduction heating.

Although  no solid evidence exists (i.e., there are no
independently  verified reports) to indicate that low-
intensity RF radiation is  mutagenic,  sophisticated
concepts and designs are just beginning to appear in
experiments concerning the biological effects of RF
radiation. Thus, although it is premature to proclaim
unequivocally that RF radiation is not mutagenic, the
majority  of evidence at present  indicates, in  the
absence of temperature elevation, that is the case.
The  above studies of the biological  effects  of  RF
radiation that relate to genetics and mutagenesis are
summarized in Table 5-24.

Since 1980, several recent  reports at scientific
meetings have supported  or extended the concerns
raised in this section. Kremer et al. (1983) examined
the puffing pattern of giant chromosomes from the
salivary  gland  of  the  midge Acricotopus lucidus
following a 2-h exposure to fields swept between 64
and 69 GHz CW, or at single, stabilized frequencies of
67.2 or 68.2 GHz CW. Power densities were less than
or equal to 5 mW/cm2 and resulted in temperature
rises of less than 0.03°C. There was up to a 10 times
reduction in the Balbiani ring BR2 in chromosome II;
that is, the normal mass of DNAfibers usually seen at
one  chromosome locus was dramatically retarded
following  microwave exposure. No change in  the
normal puffing pattern  was observed in either  the
sham treated samples or the sham plus heat treated
samples  (2.5°C temperature increase). This  result
                       5-102

-------
 Table 5-24.    Summary of Studies Concerning Genetic and Mutagentc Effects of RF-Radiation Exposure



                                                                                        Exposure Conditions











Ol
*•*
8





Effects
No change in thermal denaturation profile.
except at elevated temperature
Change in thermal denaturation profile and
hyperchromicity of DNA extracted from
testes following exposure
No chromosome aberrations in white blood
cells
No sister chromatid exchange in bone
marrow cells
No chromosome aberrations in CHO-K1 cell
line if temperature maintained
No chromosome aberrations or change in
mitotic activity in regenerating liver
cells in rat
No mutation induction
No mutation induction observed in Ames
tester strains
Reduction in survival concomitant with rise
in sample temperature
No reduction in survival or mutational
events
No reduction in survival or mutational
events
No detectable lethal events due to no
change in CPUs
No observable change in molecular
structures because no change in IR spectrum
No repairable DNA damage
No change in growth pattern;
enhanced colony-forming activity
No change in mutation frequencies at
either of two loci controlling requirements
for adenine and tryptophan
No mutagenic effects in exposed embryos
No changes in generation time, sex ratio,
or sex-linked lethal mutations in offspring
No mutations in adult males as evidenced by
chromosome loss; nondisjunction; or sex-
Species
DNA
Mouse
Chinese hamster
Mouse
Chinese hamster
Rat
E COll
E coli
S typhimurium
£ coli
S typhimurium
S cerevisiae
S cerevisiae
S. cerevisiae
E. coli
E coli
E coli
E coli
S. cerevisiae
D. melanogaster
D. melanogaster
D. melanogaster
Frequency
(MHz)
2.450ICW)
1.700ICW)
985 (CW)
2,450 (CW)
2,450 (CW)
2,450 (CW)
13.12 (CW)
1312 (PW)
2.450 (CW)
1,700(CW)
2.450 (CW)
8.600-9,600 (PW)
8,600-9,000 (PW)
8.600- 9,000 (PW)
8.600-9,000 (PW)
70.500 (CW)
73.000 (CW)
9,400 (CW)
1 7.000 (CW)
2,600 4,000 (CW)
3.2OO(CW)
8,600 (PW)
2.450 (CW)
2.450(CW)
8,500-9,600 (PW)
2,450 (CW)
2,450 (CW)
29 (CW)
146(CW)
Intensity
(mW/cm2)
134
<50
10
5-45
20
<200
4.45 kVD-p/m
44.1 kVp.p/m
10 or 50
250 Vp-p/m
20
10, 45
1-20
<45
<49
• 60


0005-50
20
1-45
—
4.600-6,500
600 Vm./m
62.5 V^/m
Duration
(days x mm)
1 x 960
1 x 80
5x 15
28x480
1 x 30
1 x 1,680-2,640
1 x 180-240
1 x210
1 x90
1 x90
1 x90
1 x90
1 x 120
1 x180
1 x 300
1 x 1,440
1 x480
1 x 660-720
1 x 60-420
1 x240
1 x 120
1 x120
1 x 360
1 x45
1 x720
1 x720
SAR"
(W/kg)
67 (est)
<2.4 (est-testes)
<0.26 (est-testes)
21
21
<360 (est)
1.3 (est)
1 5 or 70
3
40
18. 80 (est)
>50
<80
<80
£17 (est)
<28 (est)
19
21 or 16
12
0008-75
40
<80 (est)
100
150-2 10 (est)
0.024 (est)
0015 (est)
Reference
Hamrick(1973)
Varma and Traboulay
(1976. 1977)
Huang el al (1977)
McReeera/ (1981)
Mam eta/ (1978)
McLeesef at. (1972)
Blackman el al
(1976)
Dutta el al. (1979a)
Dutta el al (1980)
Dardalhon er al
(1979)
Dardalhon et al
(1980)
Corelli el al
(1977)
Corelli et at
(1977)
Dutta et al (1979b)
Blackman et al
(1975)
Dutta el al (1979a)
Hamnerius et al
(1979)
Pavel al (1972)
Mittler (1976)
linked recessive lethals

-------
     Table 6-24.    (Continued)
                                                                                                  Exposure Conditions
Oi
Effects
No mutagenic changes (recessive lethals)
in adult females
No significant germ-cell mutagenesis in
weekly breedings
No significant germ-cell mutagenesis in
weekly breedings
Same, except decrease in prec, lancies,
indicating temporary sterility caused
by elevated testicular temperatures
Induction of a repressed protein, colicin.
indicating a change in the genetic processes
Change in growth rate that was very fre-
quency specific, indicating an alteration
in the processes of the cell
Chromosome aberrations in lung cells in
vitro at two frequencies but not at two
closely related frequencies, 15 or 25 MHz
Increase in chromosome translocations in
sperm cells
Increased mutations and lethality
No change in growth when compared to
temperature controls
No change in survival or mutation
induction
No change in dominant lethality
No change in mutation induction
Higher mutagencity index perhaps
due to heating and RF
Species
D melanogaster
Rat
Rat
Rat
E coli
S cerevisiae
S cerevisiae
Chinese hamster
Mouse
S tvphimurium
E coli
E coli
S. cerevisiae
Mouse
E. coli
S. typhimurium
Mouse
Frequency
(MHz)
98 5(CW)
2.450 (CW)
2,450 (CW)
2,450 (CW)
37,000 (CW)
41. 000-42.000 (CW)
41,650-41, 825 (CW)
19(PW)
21 (PW)
9,400 (PW)
2,450 (CW)
8.800 (PW)
9,400 (CW)
1 7,000 (CW)
70.000-75.000 (CW)
9,400 (CW)
1 7.000 (CW)
70,000-75.000 (CW)
2.450 (CW)
27.12(CW)
2.450 (CW)
3,070 (PW)
27.12(CW)
2,450 (CW)
3.070 (PW)
2,450 (CW)
2,450 (CW)
1.700(CW)
Intensity
(mW/cm')
03 V,m,/m,
(FM) at
audio
5
10
28
0001-1
1-3
/10
up to
300 kVD-p/m
0 1-10
5,100




100
50
10
Duration
(days x min)
224 x 1,140
106 x 240
5 x 300
20 x 240
1 x 30-120
1 x 660
1 x 180
1 x 30
10 x 60
1 x 0.03-0.48
1 x 900
1 x 30
1 x 30
1 x30
1 x 30-2,880
1 x 30-2,880
1 x 30
1 x 30
1 x 60-400
1 x 60-400
1 x 60-400
1 x 60-400
1 x 60-400
1 x 60-400
1 x 10
3x 10
1 x80
SAR-
(W kg)
0.0004 (est)
4.7-0.9
2
56

4-11 (est, av)"

0.05-5 (est)

12
23
<28
9
23
28
28
43
-4
35-100
35-100
35-100
35-100
1 1.4 (est-testes)
5 7 (est-testes)
05 (est-testes)
Reference
Mittler (1977)
Berman et al. (1980)
Berman et al (1980)
Berman el al. (1980)
Smolyanskaya &
Vilenskaya (1973)
Grundler et al.
(1977)
Grundler and
Keilmann (1983)
Heller (1970) and
Mickey and
Koertmg (1970)
Manikowska et at.
(1979)
Blevins el al
(1980)
Hossain and Dutta
(1982)
Dardalhon et al
(1981)
Saunders et al.
(1983)
Anderstam et al
(1983)
Varma et al
(1976)
     "est = estimated, est-testes - estimated for testes only; av = average.

-------
 reinforced the concerns expressed above that specific
 interactions may occur in unexamined frequency
 domains. Edwards et al.  (1983) reported microwave
 absorption in isolated DNA molecules in salt solution
 when exposed over the 8- to 12-GHz range. This
 absorption was strongly dependent on the number of
 single strand breaks in each molecule. Since these
 breaks occur naturally, as during DNA replication or
 repair, this result provides a basis for assuming that
 certain  biochemical states of DNA could be highly
 susceptible to change by exposure to  microwaves.
 (For example, see discussion of co-stressors, above.)

 Manikowska-Czerska et al. (1983a,b)  reported,  in
 abstract form, further work with unanesthetized male
 IRC mice exposed to 9.4-GHz pulsed, 2450-MHz CW,
 and 915-MHz CW fields (Manikowska  et al. 1979).
 Exposures  were 30 min/day, 6 days/week, for 2
 weeks at SARs between 0.05 and 20 W/kg. Two
 studies  were performed. In one, increased chromo-
 somal translocations were observed in  meiotic cells
 from the testes of exposed mice. The numbers of
 translocations followed  the  same unusual dose-
 response profile described in the earlier preliminary
 report (Manikowska et al. 1979). In the other study,
exposed and sham treated males were mated with
unexposed  females each week for 3 weeks.  Post-
implantation  loss was assayed on day 13-14 of
gestation. Significantly increased loss was reported
for the group mated to the exposed males. This result
 is different from that  obtained  by Saunders et al.
 (1983),  who  exposed  at a higher SAR, 43 W/kg
 (reviewed  in Sec. 5.8.3.3), and found no change in
 post-implantation survival. The incompletely de-
 scribed  exposure conditions and  analyses in the
 Manikowska-Czerska abstracts, which  did not
 receive  peer review, preclude rigorous evaluation of
 the reports; however,  the  occurrence  of intensity-
 specific effects is consistent  with qualification 3,
 stated above. Conversely, if  the  animal  restraint
 produced additional stress in  the exposed  animals
 (Justesen  et al.  1971), the results may  address
 qualification 4. Nevertheless, because the procedures
 and exposure levels employed by these two research
 groups differed, additional experiments are needed to
 resolve the  influences of particular exposure
 geometries,  SAR,  stress produced by  animal
 restraint, ambient temperature and cooling capacity,
 and animal species to determine the general nature of
 this response.

 A recent publication  by Grundler and Keilmann
 (1983) has corroborated their earlier report (Grundler
 et al. 1977) of frequency-specific alterations in
 growth  rates of the yeast S. cerevisiae. In this new
 study, the growth rate of yeast was monitored
 spectrophotometrically during exposure by either one
 of two different antenna systems. Numerous narrow
 frequency ranges between 41.650 and 41.825 GHz
 (CW) caused up to 10 percent change in either growth
enhancement  or  retardation  compared to control
samples. There was substantial agreement in results
observed  from the two exposure  systems. The
emphasis  on  improved  frequency resolution also
revealed that a residual  modulation  up to 0.5 MHz
was present. No SAR was given by the authors nor
could  one be  estimated from the data presented;
however, the authors stated that this effect occurred
and saturated above a threshold  intensity that is
much  less than 10 mW/cm2. Because the exposure
i ncreased the sa mple temperature from 0.16 to 0.4°C
for the power (10 to 25 mW) usually applied, results of
a small  thermal increment on the growth rate were
established separately and displayed  along with the
results of exposure at different powers. These results
reinforce the concern expressed  above  that  low-
intensity, frequency-specific effects  may exist  but
have  not been widely detected because of inade-
quacies  either in the experimental design or in the
stability  of the exposure sources. Although there are
theoretical  models that are consistent with  these
results,  no mechanism  of action  has  yet  been
established.
                                                                    5-105

-------
5.9 Life Span and Carcinogenesis
William P. Kirk
The  generalizations that can be drawn from the
literature on the effects of RF radiation on life span
and carcinogenesis are:

• There is no convincing evidence that exposure to
  RF radiation shortens the life span  of human
  beings  or experimental animals or that RF
  radiation is a primary carcinogen (cancer inducer);
  however, (1) few studies have used longevity or
  cancer incidence as end points, and (2) human
  studies have lacked statistical power to exclude
  life shortening or cancer.

• There is evidence from one group of investigators
  that chronic exposure to RF radiation (SAR = 2 to 3
  W/kg)  resulted in cancer  promotion or  co-
  carcinogenesis in three different tumor systems
  in mice.

5.9.1 Life Span
5.9.1.1 Human Studies
There are few data and no definitive studies on which
to judge the long-term effects of RF-radiation exposure
on  human survival.  Two studies have evaluated
cause of death several years after exposure of the
study populations (Table 5-25). First, information has
been developed on the mortality experience of U.S.
Government employees assigned to the Moscow
Embassy during the period  1953 to 1976, when
microwave irradiation of the embassy by the Soviets
was taking place. Comparisons were made with a
group of employees who had been stationed at other
U.S. embassies in  Soviet Bloc  cities  (Budapest,
Leningrad,  Prague, Warsaw, Belgrade, Sophia, and
Zagreb). The comparison was chosen to be as similar
as possible to the 1800 employees in the Moscow
group for  selection,  i.e., posting,  criteria, and
environmental influences, except that the posts were
not subject to  microwave exposure (Lilienfeld et al.
1978). No evidence was found that the Moscow group
had experienced  any  higher  mortality or any
differences in specific causes of death up to the time
of the report.  Investigators noted that because the
population was relatively young  it was too early to
detect long-term mortality effects except for those
serving  in  the earliest  period of the  study. The
methods and other results of this study are detailed in
Sec. 5.10, Human Studies.

Robinette et al. (1980) examined the records of a
group of 40,000 U.S. Navy personnel who enlisted
during the period 1950  to 1954. Approximately
20,000 had job classifications with  maximum
potential exposure to  radar;  a similar  number of
employees, believed to have minimum potential for
exposure, were used as a comparison group. The
 authors found no  apparent difference in mortality
 patterns between the two exposure groups more than
 20 years post-exposure. This study is also described
 in more detail in Sec. 5.10, Human Studies. Exposure
 data for both of these studies are given in Table 5-25.


5.9.1.2 Animal Studies
Although there have been many median-lethality
experiments  to  determine the lethal effects  in
animals exposed to RF radiation, data on life span
effects of experimental animals exposed to low power
levels are scarce. Only one report is known of a study
in which  animals  were  exposed to microwave
radiation as the  sole stressor and observed
throughout their life span. Spalding  et  al. (1971)
exposed 24 adult female mice to 800-MHz fields at a
power density  of  43  mW/cm2 for 2  h/day, 5
days/week for 35  weeks. A waveguide  apparatus
was used to irradiate the animals, which were cooled
by forced air. No temperature  or relative humidity
data  are reported. The whole-body-averaged SAR is
estimated to be 12.9 W/kg. Four exposed mice died of
"thermal effects" during the experiment, and a fifth is
reported to have died during exposure because it
became too obese  for the exposure chamber. The
mean life span of the remaining 19 exposed mice was
664  ± 32.2 (SEM)  days, and that of the 24 sham-
irradiated mice was  645 ± 32.2  (SEM) days,
differences that are not statistically significant.
Prausnitz and Susskind (1962) studied pathological
and longevity effects on male Swiss mice exposed to
9270-MHz (PW) radiation (duty cycle = 0.001) at an
average power density of 100 mW/cm2 for 4.5 min
daily, 5 days/week for 59 weeks. The daily exposure
produced an average body temperature rise of 3.3°C.
This  daily dose is stated to be half the LDso for the
mice. Originally, there were 200 irradiated mice and
100  control mice. Five  percent of each group was
killed for pathological and hematological examination
7 months after the beginning of irradiation, and an
additional 10 percent of each group was killed at 16
months, within a  month  of the final  irradiation.
Because of a partial contradiction  in pathological
findings in the animals killed at 7 and 16 months, all
surviving mice (19  controls and 67 irradiates) were
killed 19 months after the beginning of irradiation (4
months after final  irradiation). The data from this
experiment are  summarized in Table 5-26. The
authors state that the longevity of the mice djd not
appear to be affected under the prevailing conditions
and suggested that "conceivably microwave irradia-
tion in this modality, with periodically induced slight
artificial fevers, is of some benefit to  the animal in
combating disease." The data  generally appear to
support this  statement—and perhaps  even the
stronger statement  that significantly more irradiated
mice than controls  survived until the termination of
the experiment. If simple 2x2 contingency tables are
                       5-106

-------
 Table 5-25.   Summary of Studies Concerning RF-Radiation Exposure Effects on Life Span/Carcinogenesis

                                                       Exposure Conditions
Effects
No effect on life span or cause
of death
No effect on mortality in a
military population followed
for 20 years

Slight increase in mean life span
Increased mean and maximum life
span for "irradiated mice with
tumors." Increased mean life span
but no change in maximum life span
of non-tumor-bearing mice. Delay
in development of tumors in
irradiated mice but no change in
ultimate number of tumors
Increased mean life span in
irradiated mice (concurrent
infection-pneumonia)
Species
Human adult
male & female
Human adult
male


Adult mouse
Infant mouse







Adult mouse


Frequency
(MHz)
2560-4100
600-9500
200-5000
(est) (PW)


800
2450







9270
(PW)

Intensity
(mW/cm2)
0.005 (max)
0.01 8 (max)
~ 1
(routine)
100
(occasional)
43








100


Duration
(days x min.)
8030* x 600
180 x 1200
730 x 480
(est)


175 x 120
4x20
fin uteroj






4.5 min/day
5 days/week
59 weeks
SAR
(W/kg)
2x 10~*(max)
7x lO^lmax)
<0.05 (est)

<5 (est)

12.9 (est)
35







40 (est)


References
Lilienfeld er al.

Robinette er al



Spalding et al.
Preskorn et al.








(1978)

(1980)



(1971)
(1978)







Prausnitz and Susskind
(1962)



 'The duration of 8030 days equals the number of years (22) of irradiation of the embassy and length of the study period, but the average
 exposure of individuals is estimated to be 2 to 4 years.
 used with data from Prausnitz and Susskind (1962,
 Figure 3) and from Table 5-26, ignoring all animals in
 sacrifice series #1 and #2, the difference in survival is
 not significant at 14 months into the experiment (x2 —
 2.2) but is significant at 19 months (x2 =4.9,0.05 > p
 >  0.025). The data provided are  not  sufficiently
 detailed to permit application of the  subsequently
 developed,  more  refined  methods  of analyzing
 survival (Peto et al. 1976,  1977).  The calculation
 would be perturbed slightly by the fact that sacrifice
 series #1 and #2 were not predesignated prior to the
 beginning of exposure but rather selected randomly
 from survivors at that point in the experiment. This

Table 5-26.   Summary of Prausnitz and Susskind Data (1962)
procedure would have the effect of decreasing the
proportion of survivors in both control and irradiated
groups.

There have been some reports of altered and possibly
enhanced immunological competence after  micro-
wave exposure. These reports are discussed in detail
in Sec. 5.2, Hematologic and Immunologic Effects. In
a relevant study, Preskorn et al. (1978) compared
tumor  development and longevity  in female mice
prenatally exposed to 2450-MHz microwaves with
sham-irradiated controls. The irradiated mice were
offspring of dams exposed for 20 min/day on days 11,
Series
Sacrifice Series #1


Sacrifice Series #2






Sacrifice Series #3
(all animals
surviving at 1 9
months)
Longevity Series

Spontaneous Deaths




When
Sacrificed No
(months) Animals
7 5


16 10
(1 month post-
irradiation)




19 19
(4 months post-
irradiation)

As 40
occurred
No data 26
given



Controls (100 Total)
% of No. with % This Series
Total "Leucosis" with "Leucosis"
5 None —
apparently
given
10 1 10






19 4 21



40 4 10

26 - _




Irradiates (200 total)
No. % of No with % This Series
Animals Total "Leucosis" with "Leucosis"
10 5 None —
apparently
given
20 10 6 30






67 33.5 12 18



60 30 21 35

43 22 — —




Comment
Reported as
negative for
"leucosis"
Reported as
positive for
"leucosis";
not statis-
tically sig-
nificant
XJ<1 df) = 1.49
Reported as
negative for
"leucosis"

See text

Autolysis too
advanced for
diagnosis of
cause of
death
                                                                        5-107

-------
12, 13,  and 14 of gestation in a multimodal cavity
(SAR = 35 W/kg). Both groups were injected with
sarcoma cells at age 16 days. Both the mean and
maximum  survival time of "irradiated mice with
tumors" exceeded those of "non-irradiated mice with
tumors" (p  < 0.05).  The mean survival time  of
"irradiates without tumors" was also greater than
"non-irradiated  mice  without  tumors." When  50
percent  of the controls had died, 67 percent of the
irradiates were  alive.  The  difference was not
significant (0.05 < p <0.1), however, and maximum
life span did not differ.

Rotkovska and Vacek (1972,1977) reported increased
survival  and increased LD50-30 in mice for sublethal
(600 to 750 R) acute exposures to 200-kV X rays if the
X-ray  exposures  were preceded (1,  3, 14 days)  or
followed (30 min) by a 5-min exposure to 2450-MHz
microwaves at  100 mW/cm2.  In their  initial
experiment, the  30-day survival  after an X-ray
exposure of  600 R was increased from 14 percent
(controls)  to 53, 88, or 90  percent  when such
exposure followed the microwave irradiation by 1, 3,
or 14 days, respectively. In their second experiment,
the LDso-so was increased by 100 R with  all
microwave-irradiated groups having significantly
higher  survival at 30  days  than their control
counterparts. These data, together with the interaction
with psychoactive drugs, are illustrative of the
potentially confounding variables that may have to be
considered in determining RF-radiation exposure
limits.

5.9.2 Carcinogenesis
5.9.2.1  General
Carcinogenesis is the process of inducing cancer or
malignant  neoplasia. Neoplasia 'is uncontrolled
growth  or  cell  division in a tissue; a malignant
neoplasm or cancer is a group of cells that replicate
uncontrollably and has the capacity to shed cells which
enter the blood and travel to other parts of the body to
colonize and  form  new tumors  (metastases).
Metastasis is characteristic of malignancy but is not
required for diagnosis which is done histologically A
tumor that  does not metastasize is considered
benign, although it may well grow to  sufficient size to
be life  threatening.  Consideration  of the various
theories of Carcinogenesis is beyond  the scope of this
discussion, except that most authorities believe that
the initiating event(s) on the cellular level involves
physical or chemical alterations of a  cell such that its
descendant cells are abnormal and may override the
body's cellular  proliferation  control processes.  In
recent years Carcinogenesis has come to be viewed
as a multistaged process that for simplicity can be
divided into two major stages, i.e., induction (actual
induction  of malignant transformation by genetic
damage) and promotion (enhanced growth and/or
survival of malignantly transformed cells). Promotion
is  increasingly seen as an essential phase in the
carcinogenic process since it is assumed that many
cells capable of resulting in malignant tumors simply
die, or remain dormant unless stimulated to divide, or
are somehow protected from potentially lethal DNA
damage.  Promoters are substances that result in
enhanced progression to malignant tumors  when
applied in close proximity to transforming events. A
given moiety, chemical or physical,  may act  as an
inducer,  promoter, or -both. (Ionizing radiation, for
example, is both a weak inducer and promoter.) The
role that the immune system  plays  in  primary
Carcinogenesis has not been well documented and is
somewhat controversial. However, it is thought that
neoplasia may arise through an epigenetic mechanism
in which immunosuppression allows the development
of tumors initiated by a genetic event
Potential carcinogenicity has been periodically
discussed in connection with RF-radiation exposure
since 1953  when J.R. Mclaughlin, a medical
consultant to the  Hughes  Aircraft Corporation,
submitted a report to the military, listing leukemia as
one of the possible effects  of radar exposure
(Mclaughlin 1953). The relevant literature, which is
sparse, has been reviewed by Baranski and Czerski
(1976), Justesen et al. (1978), and Dwyer and Leeper
(1978), with little supportable evidence that RF
exposures are likely to be carcinogenic. However, the
subject remains controversial for many reasons,
including:

(1)  the failure  of the public and  the media to
    distinguish between ionizing radiation, a proven
    carcinogen, and nonionizing radiation (specifically,
    RF radiation);
(2)  the existence of several anecdotal  and/or case
    reports associating cancer in humans with RF-
    radiation exposure;
(3)  report of studies indicating that RF exposure may
    cause leukemia (Prausnitz and Susskind  1962;
    Susskind 1962) or promote the development of
    several types of cancer in mice (Szmigielski et al.
    1980, 1982);
(4)  the lack  of well-designed human  or animal
    studies that have adequate exposure data that
    are free from  artifact and are statistically
    adequate to draw reliable conclusions.

5.9.2.2 Human Studies
The two relevant and acceptable studies in this area
deal with the health status of (1) U.S. Government
employees stationed at the U.S. Embassy in Moscow
(Lilienfeld et al. 1978) and (2) U.S. Navy personnel
exposed to radar duringthe Korean War(Robinette et
al.  1980). For these two studies, only those aspects
dealing with Carcinogenesis are discussed here.
In the Moscow Embassy study, the Moscow  group
experienced less overall mortality from all causes of
                       5-108

-------
 death than did the comparison group; however, the
 death rate in Moscow group females from malignant
 neoplasms was slightly, but not significantly, higher
 than expected, and when exposure to microwaves
 was considered, it was found that the death rate was
 highest  for those  having the least exposure to
 microwaves. The incidence of malignant neoplasms,
 other than of the skin, was significantly higher in the
 Moscow female  group. In both cases, the numbers
 were very small. In the former case, 7 different cancer
 sites were involved in 8 cases, which according to the
 authors, virtually eliminates a single causal factor; in
 the latter case,  there were 10  cases involving  7
 different sites.
 One of the end points examined in the U.S. Navy study
 (Robinette et at.  1980) was the effect of exposure
 from 1950 to 1954 on the relative incidence of cancer
 during 1950 to 1976.  No statistically significant
 differences between the  high- and  low-exposure
 groups were evident for malignant neoplasms as a
 cause of death from 1950 to 1974, or as a cause of
 hospitalization in Navy or VA hospitals. Within the
 high exposure group, three subgroups were developed
 to provide a gradient of potential exposure. (Sec. 5.10,
 Human  Studies,  gives details.) There were  signifi-
 cantly more malignant neoplasms of the respiratory
 tract in the subgroup rated  as highly exposed. A
 significant trend for increased death from increased
 exposure was also noted when these subgroups were
 used. However, information relevant to the develop-
 ment of lung cancer, such as smoking  histories, was
 not available because the study used  death
 certificates to obtain data  on cause of death. Also,
 when many statistical comparisons are made, as was
 done in this study, with a p=0.05 criterion, one or two
 positive findings would be expected by chance.

 5.9.2.3 Animal Studies
The only animal  study reported to date that  was of
 sufficient  duration and employed end  points
 appropriate to detect the induction of cancer by RF
 exposure was that of Prausnitz and Susskind (1962).
They conducted  an experiment  to determine the
 pathological and life-shortening effects of chronic
 microwave exposure of mice. Two hundred male
 Swiss mice were irradiated with 9270 MHz(PW) at an
 average power density of 100 mW/cm2 for 4.5 min
 daily.  This exposure produced  an average body
temperature increase of  3.3°C.  Exposures were
 conducted 5 days/week for 59 weeks; and survival,
 body mass, blood parameters, and certain postmortem
pathological findings were compared with those of a
concurrently sham-irradiated group of 100 animals.
Tissues taken for histopathology included liver,
spleen, lymph  nodes, kidneys, adrenals, gut, lungs,
and testes. No data were given regarding ambient
temperature during  exposure. Between  exposures,
the mice were housed 10 to the cage in a cabinet
maintained between 21 and 24°C. Animals were
 sacrificed at 7,17, and 19 months afterthe beginning
 of irradiation. In addition, 23 percent of the animals
 were lost to the study by spontaneous death and
 autolysis prior  to  necropsy.  The remaining 100
 animals  (40 controls,  60 irradiates) formed a
 "longevity" study group. One of the findings reported
 in this group was "cancer of the white cells," defined
 as monocytic or lymphocytic leucosis, or lymphatic or
 myeloid leukemia. Leucosis was defined as a
 noncirculating neoplasm of the white cells, whereas
 leukemia was defined as a circulating leucosis. Data
 were  grouped  and  reported as  "leucosis."  A
 contemporary medical dictionary (Taber 1953)
 defines leucosis as (1) unnatural pallor, (2) presence
 of an abnormal  number of leukocytes  in blood, (3)
 increase in leukocyte-forming  tissue; however, it is
 not  a  usual  pathological term. Although the liver,
 spleen, and lymph nodes were removed for examina-
 tion at necropsy,  they were not  considered in
 determining lymphoid infiltration, even though these
 organs are usually involved in this kind of reaction.
 Data describing  the differential composition of the
 blood cell types are not given. A summary of the fate
 of the 200 irradiates and 100 controls is shown in
 Table 5-26.
 The results as reported are confusing.  The authors
 performed a  third sacrifice series at 19 months to
 resolve a  perceived contradiction  in the "leucosis"
 findings from the first and second sacrifice series,
 even though a 2 x 2 contingency table analysis to test
 the  prevalence  of leucosis  in  the control and
 irradiated animals sacrificed  in the  second series
 yields a x2 of only 1.49 (uncorrected for continuity)
 with one  degree of freedom  (p > 0.20). Thus, the
 prevalence of leucosis in irradiated animals was not
 significantly different from that in the control animals
 for any of the "sacrifice" groups.
 Although  leucosis in the "longevity" series is highly
 significant (p <  0.005, x2  >  8.0), these  data are
 severely compromised by (1) the marked loss due to
 autolysis of animals that would have fallen in this
 group, (2) the differential loss of control animals due
 to infection, and  (3) the premature  sacrifice  of all
 remaining animals to  resolve an apparent paradox.
 Comparing data  from two  groups of animals that
 selected themselves (by dying prior to 19 months and
 being found in time for necropsy) is not an acceptable
 statistical analysis  of experimental data. An
 acceptable  method  would  be to compare the
 prevalence rates of leucosis  in the controls and
 irradiates  in combined groups that  include all the
 animals  at  risk  in the original  longevity  group
 (longevity and third sacrifice series).  If that is done,
 the following  results are obtained:
Group
With Leucosis  Without Leucosis  Total
Control    8 (a)
Irradiated  33 (c)
Total      41 (a+c)
             51 (b)
             94 (d)
             145 (b+d)
59 (a+b)
127 (c+d)
186(n)
                                                                     6-109

-------
To test whether the leucosis rates are different in
controls and irradiates, the following statistic may be
used:

    S = n(ad - bc)2/[(a + b) (c + d) (a + c) (b + d)J

If the animals comprising the table represent a random
sample  from a larger population, S  approximately
follows a x2 distribution with one degree of freedom.
Upon substitution, the value of 3.62 (p ~ 0.06) is
obtained. Since x2 < 3.84 (p = 0.05), the rates would
not be declared different at the 5-percent significance
level. Furthermore, if  a continuity correction were
used in computing this statistic, the  resultant x2 -
2.93, with a probability of 0.09 (not significant).

There are now better statistical methods available to
analyze survival data (Peto era/. 1976,1977), but they
require detailed documentation  of the fate of  each
animal during the course of the experiment. Given (a)
the loss of histopathological data on animals because
of autolysis and infection, (b) the further complication
that different  pathologists were used for different
sacrifice  series, and (c) the absence of historical
data on the incidence of leucosis in the mouse strain
used, it  is  not feasible  to  apply these methods
retroactively. In summary, because of the previously
described problems with the biological protocol, the
lack of sound statistical methodology in experimental
design and data analyses,  and the  questionable
significance of what was reported, the 1962 study of
Prausnitz and Susskind is  of limited value in defining
the  pathological effects of chronic wholebody
microwave  radiation. As noted previously, the data on
survival  times (Sec. 5.9.1.2) are of considerable
interest due to similar findings by Spalding  el al.
(1971) and  Preskorn et al. (1978).

There have been several acceptable reports dealing
with the effects of RF exposure in the development of
experimental,  spontaneous, or chemically  induced
tumors in mice (Preskorn et al. 1978;  Szmigielski et
al.  1980, 1982). In their paper,  discussed in  Sec.
5.9.1, Life Span, Preskorn et al. (1978) reported that
the development of tumors following injection of
sarcoma cells into  16-day-old CFW mice  was
significantly delayed if the mice had been exposed to
2450-MHz radiation in utero on day 11,12,13, and 14
of gestation (20 min/day, SAP = 35 W/kg). However,
there was no  difference in the ultimate number of
tumors.

Szmigielski et al. (1980,  1982) demonstrated that
repeated exposure of mice to 2450-MHz radiation (far
field) at 5 or 15 mW/cm2 (SAR = 2 to 3 or 6 to 8 W/kg),
2 h per day, 6 days per week for varying times up to 10
months accelerates the appearance of spontaneous
mammary cancer (female C3H/HeA mice) and of skin
cancer  in  male Balb/c  mice  treated  with  3,4-
benzopyrene during or after microwave treatment.
Animals were exposed in plastic cages containing 10
mice per cage. The  chamber temperature  was
maintained at 22 to 23°C with humidity of 60 to 70
percent. Stress produced by chronic confinement in
compartments 5  x 6  x 10 cm, 1  animal  per
compartment, 20  compartments/cage  produced
approximately  the  same effect as the 5-mW/cm2
exposure. An additional finding is that this exposure
regimen increases the number of neoplastic nodules
developing in the lungs of Balb/c mice injected with
Li sarcoma cells (2 x 105  cells) and examined 14 days
later. Challenge with the sarcoma cells  after a 3-
month exposure to 5 or 15 mW/cm2 or confinement
produced the following results.
    Dose Groups
Nodules (X ± SO)
       Cage Controls
     Sham Irradiates
         5 mW/cm2
        15mW/cm2
Chronic Confinement
   2.8 if 1.6
   3.6 ± 2.2
   6.1 ± 1.8
  10.8 ±2.1
   7.7 ± 2.0
It seems clear that exposure of mice to RF radiation in
these experiments resulted in the acceleraction,  or
promotion, of three completely different tumors.
Whether the effect is by direct action of RF radiation
at the cellular or  subcellular level, a nonspecific
stress reaction (caused by RF radiation, crowding,  or
thermal stress), or  a  general  effect  on immune
response  cannot  be said with any confidence.
Exposure of 10 animals in a cage is certainly open  to
criticism on grounds of nonuniformity of energy
absorption and resulting thermal load.

Comparison of the Preskorn etal. (1978) data with the
Szmigielski et al. (1980, 1982) data is difficult since
the biological models and exposure conditions were
substantially different (offspring of pregnant female
mice  exposed at 35 W/kg,  20 min  daily for 4 days,
resulting in elevation in body temperature by several
degrees, as compared with months of exposure for 2
h daily at 2 to 3 or 6 to 8 W/kg, levels that  do not
produce noticeable  heating). The  Preskorn  et al.
conditions resulted in slower development  of
experimental sarcomas in mature offspring irradiated
in utero. whereas the Szmigielski et al. work resulted
in increased incidence of cancer, and/or accelerated
growth of experimental,  chemically induced,  or
spontaneous tumors.

5.9.2.4 Anecdotal and Other Unsupported
Reports
Much of the impetus for public and media belief in the
carcinogenicity of RF radiation can be traced  to the
popularization of anecdotal and other unsupportable
reports by Brodeur (1977), Zaret (1976,1977), Dwyer
and Leeper (1978),  and others. Most of these reports
deal with a perceived increase in cancer incidence in
one of several  groups of defense contract personnel
working in RF-radiation research and development,
usually in situations having  the potential for
substantial X-ray exposure  from the RF generators
                       5-110

-------
being used. No reliable reports on these incidents
have appeared in the scientific or medical literature.

A good illustration of the kind of misunderstanding/
misinterpretation that has occurred is the so-called
North Karelia Connection. In the early to mid-1970's,
the Finnish government, with the  World Health
Organization, conducted  a program known as the
North Karelia Project, designed to decrease morbidity
and mortality from the high levels of cardiovascular
disease (CVD)  in eastern Finland by identifying
causative  factors,  devising  means for primary
prevention, and strengthening treatment and
secondary prevention (Puska et al. 1978; Keys 1970).
The proximity of North Karelia to the Soviet border
prompted  Dr. Milton Zaret, who  is known for his
investigations into causation of cataracts and lens
opacities from radiation, to speculate in a presentation
(at the  1973 Warsaw Symposium on the Biologic
Effects and Health Hazards of Microwave Radiation)
that microwave radiation from Soviet communications
or radar might  be contributing to the incidence of
CVD.  His remarks do not appear in the Symposium
record. This CVD hypothesis was restated, however,
in a published article (Zaret  1976) and letter to the
editor (Zaret 1977), both of which include allegations,
without data or reference other than a newspaper
article, of the emergence of an increased incidence of
cancer in North  Karelia. The Zaret article (1976) was
later misinterpreted  by Dwyer  and  Leeper (1978),
who presented  the  North Karelia project as being
designed to test the hypothesis that RF radiation
caused or contributed to heart attacks or cancer. The
Finnish government (K. Jokela, personal communication
to M.  Hattunen, Scientific Counselor  of Embassy of
Finland, 2133 Wisconsin Avenue N.W., Washington,
DC 20007, November 22, 1978) specifically denies
the existence of an abnormal cancer incidence in
eastern Finland and knowledge of any  possible
linkage to  microwave fields.

5.9.3 Unresolved Issues
Because few RF radiation studies in man or animals
have employed life span or cancer as end points and
none has  had sufficient statistical power and
adequate quality control to place an upper limit of risk
at less than twice control incidences, the questions of
RF-radiation carcinogenesis or life shortening are
still open.  None of the complete  reports in the
literature  presents a convincing  case for  the
existence of a significantly increased risk of cancer
induction or life shortening in exposed populations.
Neither theory nor the existing data on RF-radiation
mutagenesis support the notion  of  a  role for RF
radiation in cancer induction except, very remotely, at
exposures causing substantial tissue heating. The
data of Szmigielski era/. (1980,1982), however, raise
the possibility that RF radiation may act as a cancer
promoter even at levels within the physiologic limits
of thermal regulation of  the animal. In a  different
system with higher exposures, Preskorn et al. (1978)
observed  a delay in  development of experimental
sarcomas. The  existence of  either promotion or
inhibition is yet to be confirmed in other systems and
laboratories and the mechanisms of action determined.

Two letters to the editor have appeared recently in the
New England Journal of Medicine  suggesting an
association of polycythemia vera with occupational
microwave exposure  (Friedman  1981)  and of
leukemia  with occupational exposure to a variety of
electric and magentic fields, including radio, TV, and
other electronic  devices (Milham 1982).  Neither of
these  letters meets the literature selection criteria
established for this review but should lead  to more
elaborate studies to resolve the questions raised. A
critical issue is the difficulty of developing exposure
data or information in human population studies.
The relationships of RF radiation to other physical and
chemical  moieties in carcinogenesis or other
physiologic interactions having the potential to affect
survival have not been characterized sufficiently. For
example,  (1) microwave-induced hyperthermia has
been demonstrated to increase the effectiveness of
X-ray treatment  of various cancers and to increase
the LDso for mice  exposed  to X rays; (2) in
nontherapeutic  situations, as noted above,  both
inhibition and promotion of growth of experimental
and chemically induced tumors  in  mice have been
reported;  (3) Riddle et al. (1982) demonstrated that
exposure (2450 MHz, CW) at 20 or 30 mW/cm2 (SAR
= 12 to 18 W/kg) after the injection of Salmonella
typhimurium lipopolysaccharide  (IPS) in male  mice
significantly decreased the LPS dose required to kill
50  percent of the mice. Irradiation prior to the LPS
injection did not affect the LDso; (4) Chang et al.
(1981) found that exposure to 5  mW/cm2 of 1-GHz
radiation for 20 min significantly  inhibited the ability
of methotrexate  (MTX) to delay the  development of
CNS leukemia in mice when irradiation occurred after
MTX treatment.

It is easy to  postulate feasible scenarios where
interactions of RF radiation with neurologically active
drugs (tranquilizers,  medication for treatment of
various cardiovascular problems,  etc.) or synergistic
effects with increased temperature, both noted in Sec.
5.5.5 (Interactions with Other Stimuli), result in life-
threatening situations. In view of the relatively few
multiple-agent-effects studies involving RF radiation
as one of the agents, the foregoing reports could well
be the harbingers of future developments in the study
of the biological effects of RF radiation.
                                                                      5-111

-------
5.10 Human Studies
Doreen Hill
The general approach used in this document for the
evaluation of the health-effects literature is stated in
Sec. 2. Rigorous criteria were established and applied
to the reports of experimental results in order  to
establish what is believed to be a credible data base.
However, it is difficult to apply these selection and
review  criteria uncompromisingly to the human
studies or epidemiological literature, largely because
the research  method is  population based and
observational rather than experimental. The papers
included here were selected because they were
judged  to present relatively more information on
exposure parameters  and/or  more  rigorous  or
analytical study designs. The amount of detail  in
reporting and the degree of specification of study
methods and procedures (use of controls, statistics,
control of confounding variables, and so forth) were
considered important. Case reports are not reviewed.

From this review, the following general conclusions
can be drawn:

• The currently available epidemiological data on RF
  radiation are  very  limited and  not useful for
  deriving environmental exposure limits.  This
  conclusion is based on various problems in the
  data  on human beings, as discussed in  Sec.
  5.10.5, Unresolved Issues.

• Two recent exploratory studies of physiotherapists
  who  use RF equipment in  their  occupation
  describe potentially significant findings of heart
  disease, primarily ischemic heart disease, in adult
  males and poor pregnancy outcome in female
  physiotherapists. However, neither study provided
  quantitative information on radiation levels in the
  work environment.
• Although some studies have associated ocular-
  lens defects with microwave radiation  exposure,
  no data would at present support a conclusion that
  low-level, chronic exposure to microwave
  radiation induces cataracts in human beings.
Table 5-27 summarizes those studies on human
beings for which SAR values could be estimated.
5.10.1  Occupational Surveys/Clinical Studies
The  majority of reports in the  literature concern
people  occupationally exposed  in  military  or
industrial settings.  A wide  variety of conditions,
symptoms, and clinical measurements are usually
evaluated. The health conditions investigated usually
are pre- or subclinical instead of overt or diagnosed
disease. Studies of this type that focused on single
rather than multiple end points are described later.
Barron  er a/. (1955) presented  results of a  study
conducted to evaluate changes  in various physical
characteristics of radar  personnel employed  by an
airframe manufacturer. A total of 226 exposed workers
were initially included in the medical surveillance
program. The radar workers were characterized by
their duration of exposure, as shown in Table 5-28.
Controls totaling 88 subjects, stated to have had no
industrial radar exposure,  were also examined.
Methods of selection of cases or controls were not
specified. The age distribution of all subjects ranged
from 20 to more than 50 years, with the majority
under 40 years of age. However, the controls were
older (Table 5-29).

A decrease in the number of polymorphonuclear cells
below 55 percent was observed in 25 percent of the
radar workers vs. 12 percent of the controls, but this
change  did not occur in a second study (Barron and
Baraff 1958). An increase in monocytes and eosinophils
was also observed for the exposed group, but in the 1958
report these effects were  attributed to a technical
error. Platelet counts and urinalyses were similar in
the two groups.  Ophthalmological examinations
revealed ocular anomalies of several diverse types
among 12 exposed persons vs. 1 case in the control
group.  The medical surveillance program was
extended to permit periodic reexaminations (Barron
and Baraff 1958). No significant changes in physical
health status were noted.

The radar bands of exposure included S-band (2880
MHz) and X-band (9375 MHz);  the majority  of
personnel worked with  or around APS-45 and
AN/APS-20-B and E radars. Exposure times and
power  densities  for  individuals  could not  be
developed, but zones at various distances from the
antenna were specified and used to estimate three
ranges  of power densities.  The minimal average
power density in Zone A was 13.1 mW/cm2. The field
strength in Zone B ranged from 3.9 to 13.1 mW/cm2,
and in Zone C the levels were <3.9 mW/cm2. The
authors  stated that because of the relatively low
power densities,  personnel working in Zone C were
eliminated from  the  study.  By establishing these
zones of potential exposure, steps were taken to limit
entry of  personnel to Zone A. Zone B was judged to be
safe for occasional but not constant exposure. As a
result, most  subjects continuing in or added to the
examination  program  are believed to have had
incidental  exposures to power  densities  less than
13.1 mW/cm2.

The strengths of  this study are the attempts to
estimate potential exposures and to reexamine the
workers periodically. But  there  are some major
problems  with  study methods and analyses. For
example,  there were fewer control subjects than
radar workers.  Also, the comparability of the two
groups  is questionable because of age differences
(Table 5-29). None of the observed results was tested
statistically.
                       5-112

-------
Table 5-27.   Summary of Selected Human Studies Concerning Effects of RF-Radiation Exposure

                                                    Exposure Conditions
Effects
No significant change in health status
of exposed personnel


No differences in three major
diagnostic categories between
the two groups of microwave workers
No differences observed in clinical
evaluations; more subjective
complaints in exposed group
No effect on life span or cause
of death


No effect on mortality in a
military population followed for
more than 20 years

Decreased number of sperm/ml
of ejaculate; reduced percentages
of normal and motile sperm in
ejaculate
Species
Human
adult
male

Human
adult
male
Human
adult
male
Human
adult male
and female

Human
adult
male

Human
adult
male

Frequency
(MHz)
400 (PW)
2880 (PW)
9375 (PW)

Radar


Radar


2560-4100

600-9500

200-5000 (est, PW)



3600-10,000



Intensity
(mW/cm2)
~4 (avg)
—4 (avg)
~4 (avg)

<0.2 (avg)
X).2 (avg)
6.0 (max)
<5


0.005 (max)

0.01 8 (max)

"""*!
(routine)
100
(occasional)
Tens to
hundreds
/M//cm2

Duration
(Years)
0-13



1-10


5-10


22*

0.5

2



1-17
8 (avg)


SAR (W/kg)
(est)
-0.16
-0.12
-0.12

<8x10"3
>8x10'3
0.24 (max)
<0.2


2x10"
(max)
7x10"
(max)
<0.05

<5

0.3-4x10"2



Reference
Barren and Baraff
(1958)
(See also Barren el al
1955)
Czerski ef al. (1974);
Siekierzynski et al.
(1974a,b)
Djordjevic et al. (1979)


Lilienfeld et al. (1978)



Robinette et al. (1980)



Lancranjan et al. (1975)



•Number of years of irradiation of the embassy and length of the study period, but the average exposure of individuals is estimated to be 2 to
 4 years.
Table 5-28.
Distribution of Years of Exposure for 226 Radar
Workers*
Years of Exposure
           No. of Workers   Percent of 226
    Oto2
    2 to 5
    5 to 13
               106
               83
               37
47
37
16
'Data from Barron et al. 1955.
Table 5-29.
Age Distribution of 226 Microwave Workers
and 88 Controls*
                 Radar
                                    Controls
Age
20 to 29
30 to 39
40 to 49
50+
% of 226
34
49
13
4
Cumulative % % of 88 Cumulative %
34
83
96
100
14
40
27
19
14
54
81
100
*Data from Barron et al. 1955.
It is in occupational surveys and industrial health
surveillance programs that the potential for neuro-
logical and behavioral changes has been investigated.
The Soviet and Eastern European literature frequently
describes a collection of symptoms reported to occur
in personnel  industrially  exposed to microwaves.
These collective symptoms, which have been
variously  called the "neurasthenic syndrome," the
"chronic overexposure syndrome,"  or "microwave
sickness," are based on subjective complaints that
include  headaches, sleep disturbances, weakness,
decrease of  sexual activity (lessened libido),
impotence, pains in the chest, and general poorly
defined  feelings of non-well-being (Baranski and
Czerski  1976, p.  157).  Also described are  labile
functional cardiovascular changes including brady-
cardia (or occasional tachycardia), arterial hyperten-
sion  (or  hypotension),  and  changes in cardiac
conduction; this form of neurocirculatory asthenia is
also attributed to nervous system influence (Silverman
1980).
Czerski,  Sierkierzynski, and co-workers in  Poland
evaluated 841 men occupationally exposed to pulse-
modulated microwave radiation (Czerski et al. 1974;
Sierkierzynski et al.  1974a,b).  The radiation
frequencies were not explicitly stated, but one can
infer  from the references to pulse  modulation and
radiolocation that  the working  environment  dealt
with radar frequencies. The age distribution of the men
ranged from 20 to 45 years. The men  had worked
various periods of time, with some employed over 10
years. An unexposed control population comparable
in general working conditions and socioeconomic
status could not be established; therefore, the study
group was subdivided into two groups on the basis of
level of microwave exposure. One group consisted of
507 men exposed  to mean power densities greater
than  0.2 mW/cm2, with short-term exposures
estimated to reach 6 mW/cm2 The other group was
334 men exposed to mean power density levels less
than 0.2 mW/cm2.

The health end points evaluated covered three major
categories: neurotic  syndrome, digestive-tract
functional  disturbances, and cardiocirculatory
disturbances with abnormal electrocardiogram (ECG)
                                                                      5-113

-------
findings. According to Polish occupational exposure
criteria, these conditions are considered contraindi-
cations for work in a  microwave environment. The
neurotic syndrome was defined by a variety  of
symptoms such as  fatigue, headaches, sleep
disturbances, and difficulties in memorizing and
concentrating. Psychologic examinations were given.
Comparisons were  made between  and within
exposure groups according to age and duration  of
occupational exposure. The two groups were found to
be similar with respect to the distribution of these
symptoms and conditions. There was no dependence
on the duration of exposure.

In 1979,  Djordjevic  et al.  reported  on medical
evaluations of radar workers, aged 25 to 40 years,
with a  work history  of  5 to  10 years. Specific
frequencies of exposure were  not cited but were
stated to be within the whole range used in radar
operations. Evaluation of the working environment
was  undertaken, including power-density measure-
ments. Although the environmental analyses are not
given in the paper, it was concluded that the workers
were exposed to pulsed microwaves within a wide
range of intensities but generally at levels less than 5
mW/cm2. The lower limit of this exposure may have
been 1  mW/cm2, but the discussion does not clarify
whether this estimate refers to the workers included
in this study or to radar station personnel in general.
The control group consisted of 220 persons reported
to be similar in  age, character of work  regime, and
socioeconomic status.  The controls  did not have
work experience with microwave sources. Selection
criteria  or further descriptive information was not
given for either the cases or controls.

Ten major end points or diagnoses were covered in the
clinical evaluation, including ophthalmologic exam-
inations. The medical  evaluations were conducted
by clinical specialists who followed the same scheme
for classification of abnormalities. The two groups did
not differ with respect to the 10 diagnostic categories.
Functioning  of the nervous and cardiovascular
system was analyzed in greater detail. Electrocardio-
gram results, multiple  biochemical and  hematologic
indicators, frequency of sleep disturbance, inhibition
of sexual activity, and  impairment of memory were
not different  between  the groups. Radar workers
reported more subjective complaints of headache,
fatigue, and  irritability. Based on their  survey of
working conditions, the authors  attribute the latter
result to specific problems such as poor lighting, poor
ventilation, and high noise levels in the environment
of the  radar workers as well as the need to
concentrate on the radar screen. If that is true, the
working environments for the two groups were
evidently not similar in all  respects, even though the
authors stated that the controls  had  been matched
on the  basis of comparable work regimes. Other
factors could also have been operating to produce
more subjective complaints. One possibility is
reporting bias on the part of the radar workers, e.g.,
enhanced awareness of the  possible "microwave
sickness" syndrome. Another environmental factor
may have been present that may not have been
measurable  in  the type of  environmental survey
conducted or detectable through  these clinical
evaluations; however, microwave exposure cannot
be clearly excluded.


An exploratory study was published recently of male
physical therapists who use RF,  infrared, and
ultrasound diathermy equipment in their occupation
(Hamburger et al.  1983). The  study population
consisted of 3004 men who  responded to a mail
questionnaire survey.  The cohort was divided into
subgroups according to the types  of exposures they
had experienced, that is, by their  use of the various
types of diathermy  equipment and treatment
modalities—ultrasound, microwave, shortwave, and
infrared. High and low exposures were approximated
by considering length  of employment, frequency of
treatments per week, and combinations thereof. An
association between heart disease (primarily
ischemic heart disease) and exposure to shortwave
(27-MHz) radiation was the only consistently
significant finding in comparisons between high and
low exposure groups. In general, the prevalence rates
for heart disease were lower than rates in a general
population comparable with respect to sex, age, and
race. The lower prevalence rates  were attributed to
higher socioeconomic status in the study group, to a
health care occupation in a medical setting, and to a
"healthy worker effect." The authors considered their
report to be  an  exploratory  analysis because the
response rate to the  questionnaire was low (58
percent), the nonresponderits could not be character-
ized, and individual exposures could not be estimated.

5.10.2 Mortality Studies
Lilienfeld and associates (1978) completed a broad
survey of the mortality and morbidity experience of
Foreign  Service employees and their dependents to
assess  the  potential health consequences  of
microwave irradiation  of the U.S.  Embassy  in
Moscow. The health status of  Foreign  Service
employees and those from other agencies who had
served in the Moscow Embassy during 1943 to 1976
was compared with that of employees at eight other
embassies or consulates in Eastern Europe over the
same time period.

The microwave irradiation of the  Moscow Embassy
was first detected in 1953 and subsequently varied in
intensity, direction,  and frequency over time. The
frequencies ranged from 0.6 to 9.5 GHz (U.S. Senate,
Committee on Commerce, Science,  andTransportation
1979; Pollack 1979). The measured average power
densities oyer time are given in Table 5-30.
                       5-114

-------
 Table 6-30.   MicrowaveExposureLevelsattheU.S. Embassy
            in Moscow*
                 Exposed           Power Density and
 Time Period         Area of Chancery     Exposure Duration
 1953 to May 1975     West Facade        Max of 5 /M//cm2
                                 9 h/day
 June 1975 to Feb. 1976 South and East Facade 18 /yW/cm2
                                 18 h/day
 Since Feb. 7, 1976     South and East Facade Fractions of a^W/cm2
 	18 h/day	
 •Data from Lilienfeld et al. 1978.

 Extensive efforts were launched to identify and trace
 the populations. Information on illnesses, conditions,
 or symptoms were sought from two major sources: (1)
 employment medical records, which  were fairly
 extensive, given examination requirements  for
 foreign duty,  and (2) a  self-administered health
 history questionnaire. Questionnaire responses were
 validated for a stratified sample by review of hospital,
 physician, and clinic records.  Death certificates were
 also sought, although other sources also were used
 to ascertain mortality status.

 Standardized mortality ratios for various subgroups
 were calculated for each cause of death, were
 standardized for age and calendar period, and were
 specific for  sex.  Similar  procedures were used to
 develop summary indices of morbidity.

 A total of 4388 employees  and 8283  dependents
 were studied. More than 1800 with 3000 dependents
 were employed at the Moscow Embassy and 2500
 with more than  5000 dependents worked at the
 comparison posts.  Ninety-five percent  of the
 employees  were traced. Receipt of  completed
 questionnaires was less successful, with an overall
 response rate  of 52  percent for State Department
 personnel.

 Based  on information in medical records, various
 health  problems  were generally similar, with two
 exceptions.  Moscow  employees had a threefold
 greater risk of acquiring  protozoal infections than
 comparison-post employees. In general, both sexes in
 the Moscow group had somewhat higher frequencies
 of most of the common kinds  of health conditions
 reported. Lilienfeld et al.  (1978) stated, "However,
 these most common  conditions represented a very
 heterogeneous  collection  and it is difficult  to
 conclude that they  could have  been related  to
 exposure to microwave radiation since no consistent
 pattern of increased frequency in the exposed group
 could be found."

Some excesses were reported by Moscow employees
in the  health  history questionnaire. Both sexes
reported more eye problems due to correctable
refractive errors. More psoriasis was reported by men
and anemia by women. The Moscow  employees,
especially males,  reported more symptoms such  as
irritability, depression, difficulties in concentration,
and loss of memory. It is possible, however, that a bias
 due to awareness of potential adverse effects is
 operating, since the strongest differences were
 present in the subgroup with the least exposure.

 The observed mortality was less in both male and
 female  employees  than  expected,  based on  U.S.
 mortality rates; the  male  employees  had more
 favorable experience than female employees. In both
 sexes, cancer was the predominant cause of death.
 The Moscow and comparison groups did not differ
 appreciably  in overall  and specific  mortality.
 However, the population was relatively young; it may
 have been too early to detect  long-term mortality
 effects.

 The authors concluded that no convincing evidence
 was discovered to implicate microwaves in the
 development  of adverse health effects at the time of
 the  analysis. But they also carefully  discussed
 the limitations  inherent in the study: uncertainties
 associated with the reconstruction of the employee
 populations and dependents, difficulties of obtaining
 death certificates, the low percentage of responses
 for the questionnaire, and the statistical power of the
 study. The limitation most critical for consideration in
 a document such as this relates to ascertainment of
 exposure. Problems relative  to individual mobility
 within the embassy and variation of field intensities
 within the building are present in this study as in any
 other. No records were available on where employees
 lived or worked, so one had to rely on questionnaire
 responses to estimate an  individual's potential for
 exposure. The highest exposure level (18 /M//cm2)
 was recorded for only 6 months in 1975-76; thus, the
 group exposed  to the most  intense fields had the
 shortest cumulative time  of exposure and of
 observation in the study.

 Robinette and Silverman (1977) and Robinette et al.
(1980) examined mortality and morbidity among U.S.
 naval personnel occupationally exposed  to  radar.
 Records of service technical schools were used to
select subjects for the study; the men graduated from
technical schools during the  period from 1950
through 1954. Exposure categorizations were made
on the basis of occupational specialty. The exposure
group  (probably highly exposed) consisted of
technicians involved in repair and maintenance of
 radar equipment. The controls (probably minimally
exposed) were involved in the operation of radar or
radio equipment. It was estimated  from shipboard
monitoring that radiomen and radar operators (in the
 low-exposure group) were generally exposed at  less
than 1 mW/cm2, and gunfire control and electronics
technicians  (in  the high-exposure group) were
exposed to higher  levels  during their duties. Over
40,000 veterans were included in  the study, with
about equal numbers in these two  major exposure
classifications. The mean age in 1952 of the low-
exposure group was 20.7 years and of the high-
exposure group, 22.1 years. In conjunction with naval
                                                                     5-115

-------
personnel, an effort was also made to develop an
index of potential exposure, termed Hazard Number,
for a limited portion of the population. This number
was based on the duty months multiplied by the sum
of the power  ratings (equipment output power of
gunfire-control  radars (ship)  or  search radars
(aircraft)) where technicians were assigned.
Medical information was obtained through Navy and
Veterans  Administration records. Records  were
searched for information on four major end points: (1)
mortality, (2) morbidity via in-service hospitalizations,
(3) morbidity via VA hospitalizations, and (4) disability
compensation.
Mortality was ascertained through the VA beneficiary
system. Strokes, cancers of the  digestive tract and
respiratory system, and leukemias were elevated for
the high exposure group, but none of the increases
was statistically significant.  The authors noted that
the differences in mortality from malignant neoplasms
of the lymphatic and hematopoietic system were not
statistically significant. As seen  in  Table  5-31,
comparisons  were also made within the high
exposure group across Hazard Number categories. In
this case, only two comparisons were statistically
significant: (1) the difference in  respiratory tract
cancer between those with a Hazard Number smaller
than 5000 vs. larger than 5000  and (2) the test for
trend for all diseases combined. These results maybe
fortuitous since one or two positive findings might be
expected  when  many statistical comparisons  are
made. Furthermore, additional information relative to
the  development of  lung cancer, e.g., smoking
histories, could not be obtained; the mortality data
were obtained from death certificates, and obtaining
background information  from next of kin was not
feasible.  Differential  health risks with  respect to
hospitalized illness around the  period of exposure
were not apparent. Subsequent VA hospitalizations
and disability awards provided incomplete information.

Because the study focused largely  on the use of
automated VA record systems, it was not possible to
determine non-Navy or non-VA hospitalizations,
nonhospitalized conditions, reproductive histories, or
subsequent employment  histories.  Since actual
individual exposure  could not be  reconstructed
retrospectively,  only an  estimate of the potential
exposure of the individuals was possible.

5.10.3 Ocular Effects
The potential of RF radiation to induce cataracts and
less  significant lens defects and opacities has been
cited in both U.S. and foreign  literatures. Ocular
effects have, in fact, been the major end point of study
in U.S. research, primarily in military populations. For
this  health end point more than  others,  much
attention has been devoted to careful, detailed clinical
examinations and to use of standard procedures or
protocols. However, population selection criteria are
not well elaborated;  therefore, the  possibility of
selection bias cannot  be excluded for most of these
surveys. Ocular studies have largely taken the form of
cross-sectional  clinical surveys in actively working
Table 5-31.    Number of Deaths from Disease and Mortality Ratios* by Hazard Number: U.S. Enlisted Naval Personnel Exposed to
            Microwave Radiation During the Korean War Periodf
Number of Deaths
Cause of Death
All diseases
Malignant neoplasms
Digestive organs
Respiratory tract
Lymphatic and
hematopoietic
system
Other malignant
neoplasms
Diseases of
circulatory system
Other diseases
International
Classification
of Diseases
(8th Rev.)
000-796
140-209
150-159
160-163
200-209
Residue
390-458
Residue
High Exposure
Low
Exposure
325
(1.04)
87
(0.96)
14
(0.85)
16
(0.85)
29
(6.83)
37
(1.19)
167
(1.07)
71
(1.08)
Total
309
(0.96)
96
(1.04)
20
(1.14)
24
(1.14)
26
(1.18)
26
(0.82)
150
(0.93)
63
(0.92)

0
63
(0.82)
22
(0.99)
6
(1.49)
4
(0.82)
6
(1.09)
6
(0.78)
36
(0.94)
5
(0.30)
Hazard Number
1-5000
160
(0.91)
45
(0.90)
11
(1.14)
10
(0.86)
12
(1.04)
12
(0.70)
73
(0.83)
42
(1.13)

5000+
86
(1.23)
29
(1.44)
3
(0.78)
10
(2.20)
8
(1.64)
8
(1.17)
41
(1.17)
16
(1.08)
"Mortality ratio (in parentheses) standardized for year of birth; the combined experience of the low and high exposure groups is taken as the
 standard.
tTable from Robinette et al. 1980.
                       5-116

-------
 populations and, as such, have focused on minor lens
 changes. Longitudinal studies of populations with
 known  exposure, with  follow-up of  either a
 prospective or retrospective nature, are nonexistent
 despite the identification of potential cohorts from the
 cross-sectional surveys. The research to date has not
 determined whether  lens changes  lead to overt
 clinical eye disease or conditions, e.g., the functional
 and clinical significance of minor lens changes in
 general  and over time. Thus, although it appears
 reasonable  that RF radiation at high levels could
 influence cataractogenesis, it cannot be confirmed
 without  more rigorous follow-up studies in already
 identified exposed populations. This would  be a
 reasonable line of future research if such epidemio-
 logical studies prove feasible.
 More than 50 cases of cataract development  have
 been attributed to microwave exposure (Zaret 1974).
 These instances have generally been related to acute
 exposure to high-intensity fields in the workplace. For
 most cases, work history  information sufficient to
 estimate dose rates or exposure conditions is lacking
 (Zaret  1970).
 Utilizing  Veterans Administration hospital records
 and military personnel records, Cleary et al. (1965)
 conducted a case-control study to examine cataract
 formation among U.S. Army and Air Force veterans.
 This is the only  case-control  study of clinically
 diagnosed cataracts. Cases were defined as white
 males born after 1910 who were treated for cataracts
 between 1950 and 1962, based on diagnoses given in
 VA hospital records. The restriction on year of  birth
 resulted  in a study population under 55 years of age.
 According to the authors, this was done to minimize
 dilution of the  sample with senile cataracts. Also,
 cases with certain specific cataract diagnoses were
 not included in the study because the diagnoses are
 unrelated to microwave exposure. These cases were
 congenital cataracts, cataracts in Down's Syndrome,
 traumatic cataracts, and diabetic cataracts. Controls,
 also born after 1910,  were drawn from the same
 sources by selection of men with adjacent hospital
 register numbers; therefore, the control group had a
 sample of diagnoses made in the same hospitals at
 the same time the cataract cases were diagnosed.
 Military occupational specialties, as abstracted from
 service records,  were used  to denote radar or
 nonradar workers. Job classification was then used
 as an indicator of potential exposure. The distribution
 of cases and controls according to radar exposure is
 seen in Table 5-32. The  frequency of  microwave
 exposure as denoted  by  radar  work  history  was
 similar between cases  and controls; the odds ratio
 was less than one (0.67) and was not statistically
 significant. The distribution  in different age groups is
 shown in Table 5-33. Excess risk was noted for U.S.
Air Force veterans, but the numbers were small. Only
40 radar workers were found in over 5000 cases and
controls.
 Table 5-32.    Classification by Military Occupation of World
             War II and Korean War Veterans With  and
             Without Cataracts, Based on Discharges from
             VA Hospitals*
                         Veterans' Cataract Status
Military Occupationt
Yes
           No
           Total
Radar Workers
Nonradar Workers
TOTAL
19
2625
2644
21
1935
1956
40
4560
4600
           Odds Ratio = OR = (19>(1935) = 0.67
                        (21) (2625)
*Data are from Cleary et al. 1965.
tBased on Military Occupational Specialties.
Table 5-33.   Estimated Relative Risk of Cataracts Among
            Army and Air Force Veterans by Age Group and
            Occupation*
Age
Group
Total
20-39
40-49
50-59
Cataracts
Yes
No
Yes
No
Yes
No
Yes
No
Radar
Workers
19
21
3
4
7
13
9
4
Nonradar
Workers
2625
1935
418
522
1517
1101
699
316
Odds
Ratio
0.67
0.94
0.39
1.02
X1
1.26NS
0.08 NS
3.39 NS
0.09 NS
'Data are from Cleary et al. 1965.
Cleary and Pasternack (1966) examined and scored
subclinical or minor lens changes in 736 microwave
workers and 559 controls. The population was drawn
from 16 microwave installations of various types at
different locations. Sampling constraints did  not
permit matching on age. The mean age of each group
was very similar (32.8 years for microwave workers
and 33.2 for controls), but the age distributions were
different. The age frequency distribution for controls
was broader than for the exposed group, with more
younger and older individuals, whereas more of the
exposed group were 26 to 30 years of age. Slit lamp
examinations  were  performed,  and a grading
procedure revised by Zaret and Eisenbud (1961) was
used to provide a relative measure (eye score) of
minor types of lens changes. Detailed occupational
histories were taken. The work environment in terms
of microwave exposure parameters was  not reported;
that is perhaps the greatest drawback of the study.
However,  a relative microwave  exposure index or
score was developed  by use of various parameters
such  as power  output and distance from the
microwave-generating source. The  exposure score
was used to compare lens findings with occupational
microwave exposure.

By  linear regression techniques, the number of
defects was found to increase significantly in both
groups with  increasing age, and microwave workers
were found to have more lens changes than controls
(Figure 5-7). The authors  suggested that the
                                                                      5-117

-------
Figure 5-7.   Linear regressions of eye score on age for workers
           exposed to microwave radiation and controls
           (from Cleary and Pasternack 1966.  Figure 3).
   12-


   11-


   10-


   9-


   8-


   7-
 o>
 8 
-------
The survey was later extended to six other installations
with  results reported  by Appleton in 1973.  The
exposure  of military  personnel  was  classified as
"likely" vs.  "unlikely"; the latter group served as
controls. Masking procedures were used for the
examining ophthalmologists; that is, they were not
aware of the exposure classification of subjects. The
same team of examiners performed all tests, except at
one location, to minimize interobserver variation. The
same three end points used in the 1972 study were
also  used  in the 1973  study.  Older  age  groups
demonstrated a trend  of greater opacities among
exposed personnel, but since the numbers in some
groups  were  small, the validity of this result is
questionable.
Odland (1973) reported results of a survey of ocular
anomalies in personnel from eight military installa-
tions.  The population consisted of 377 exposed
individuals and 320 controls. Exposure conditions
were not specified. Exposed personnel were defined
as  individuals whose primary duties  involved the
operation or maintenance of radar equipment. The
selection criterion for controls was duty that did not
permit actual or potential exposure to radar, although
the  actual  work  assignments  were  not stated.
Medical histories were taken, and ophthalmic
examinations were performed in a manner which
masked the exposure classification of the individuals.
The frequency of occurrence of lens anomalies was
similar in the two  groups; however, the frequency of
anomalies between control and exposed groups was
different for individuals who had a family history of
diabetes, nontraumatic cataract,  glaucoma, or
defective vision. Lens changes were  noted  in 29
percent of the exposed individuals with such a history
vs 17 percent in the controls with a family history of
eye problems. No statistical tests were applied to any
of the reported frequency distributions.
Shacklett et al. (1975) reported eye examinations of
817 military and civilian personnel. There were 477
persons with a history of microwave exposure and
340 controls without exposure drawn from eight U.S.
Air Force  bases   between  November 1971 and
December 1974.  The authors stated that detailed
work  histories were recorded (including time spent
with different types of equipment), but information on
typical exposure settings  was not given. Local unit
commanders selected the subjects by using criteria
established  by the examining team. Standard
diagnostic  criteria were  established. The  same
ophthalmologists  performed all examinations and
were not aware whether a subject was considered as
exposed or  a  control.  No differences were  noted
between the two groups in the frequency of opacities,
vacuoles,  and posterior subcapsular  iridescence.
Differences  in results were  reported as not  being
statistically significant, but the type of statistical test
was not stated. An age-dependent increase in lens
changes was noted in both groups.
 The study by Siekierzynski et al. (1974b) discussed
 earlier  also compared  lens opacities  in the two
 exposure groups (<  0.2  mW/cm2 and 0,2 to 6
 mW/cm2).  Ophthalmologic examinations  were
 performed. Lens translucency was assessed with a
 slit lamp after pupil dilatation and according to criteria
 established  for  five grades.  No  differences  were
 reported between the exposure groups nor within the
 groups for duration of exposure.

 5.10.4  Congenital Anomalies and Reproductive
 Effects
 In 1965 Sigler era/, reported a history of occupational
 exposure to radar and more military service among
 fathers  of children with  Down's Syndrome. The
 association with radar exposure  was an ancillary
 observation; this study was specifically directed
 toward  examination of the relationship between
 ionizing radiation exposure and Down's Syndrome. A
 case-control approach was used,  and 288 children
 born with Down's Syndrome in Baltimore between
 January 1946 and October 1962 were identified for
 inclusion in the study.  The investigators selected
 controls by matching each case of Down's Syndrome
 with another normal birth for (1) hospital of birth, (2)
 sex, (3)  date of birth, and (4) maternal age at birth of
 child. Of the original 288, 216 matched pairs were
 available for final analysis. Eliminations occurred for
 various reasons  such  as non-cooperation or
 equivocal diagnoses. Occupational histories and
 other data were obtained by interview of the parents.
 The fathers of the children with Down's Syndrome
 had more military service experience than  control
 fathers, but the difference was  not statistically
 significant. A more frequent history of radar exposure
 was reported by case fathers; this difference was
 statistically significant. Exposure to radar occurred
 primarily in job assignments as radar operators or
 technicians. (Some radar operators may be exposed
 to only low levels if the place of operation is distant
 from the power-generating equipment or the micro-
 wave source.)
 Cohen et al. (1977) extended  the study. The case
 series  was  expanded to include 128  additional
 verified  cases and their matched pairs born through
 1968.  To serve as an  independent replication,
 essentially the same procedures were applied along
 with the following expansions: (1) more  extensive
 questions on microwave/radar exposure and military
 service, (2) validation  of exposure histories by
 searching of armed  service records, and (3)
 chromosomal studies. The previously noted differences
 disappeared in the extended analysis (Table 5-34).
 Results  of the cytogenetic analyses are not yet
 available.

The authors stated that although the extended study
did not confirm excess radar exposure among Down's
case fathers,  the possible relationship of  such
                                                                    5-119

-------
Table 5-34.
Case Seriest
Paternal Radar Exposure Before Conception of
Index Child (from Interview and/or NAS)"

              Down's Cases   Controls
                No.
No.   %
1. Current    Exposed        20   15.7   27   21.3
            Exposure Known  127  100.0  127  100.1
2. Original    Exposed        36   18.6   30   15.2
            Exposure Known  194  100.1  198  100.0
3. Combined  Exposure       56   17.4   57   17.5
            Exposure Known  321  99.9   325  100.0
•Data are from Cohen et at. 1977.
t1, 128 pairs; 2, 216 pairs; 3, 344 total pairs.


exposure to increased risk of Down's offspring cannot
completely be ruled out. They added further that the
most challenging aspect of the investigtion was the
definition of  radar "exposure." In discussing
explantions for lack of  confirmation  of a radar
exposure factor (if one does exist), the authors offer
several possibilities. As mentioned above, inaccurate
exposure estimates could distort results. The role of
maternal factors in Down's Syndrome is so important
that paternal factors could be masked.  It  was also
suggested that detecting a paternal relationship in a
retrospective study of Down's cases may be difficult.
If radar exposure does pose some increased health
risk to the fathers, the most highly exposed males
may have  poor survivorship or an increased risk of
germinal tissue damage. This factor could result in
elimination of such men from reproductive experience.
It was further suggested that a prospective approach
might then prove more fruitful.  Although not
discussed  by  the  authors, radar equipment and
military occupational specialties could have changed
over time  in such a way as to lower risk. It has also
been  suggested that since  microwave-generating
equipment, especially of an older vintage,  may have
emitted more ionizing  radiation  than  modern
equipment, the ionizing radiation presents the actual
risk  factor, with  microwave radiation  possibly
operating  as a covariable.

Lancranjan era/. (1975) studied 31 adult males with a
mean age  of 33 years and a mean exposure of 8  years
(a range of 1 to 17 years) to electromagnetic  fields
that  "frequently  were in the  range  of tens to
hundreds of  //W/cm2." The frequencies  were
defined as microwaves of wavelengths between  3 and
12 cm and frequencies between  10,000 and 3,600
MHz. No details on the RF source(s) were provided. A
group of 30 men of similar mean age and  no known
exposure to microwaves served as a control group for
the  analysis of spermatic fluids and hormones.
Statistical  analysis of  the  results showed no
differences in  urinary content of 17 ketosteroids (as
an indirect measure of Leydig cell function) or total
gonadotropin  between the exposed and  controls.
Slight but statistically  significant decreases  were
reported for exposed personnel  in the number of
 sperm per milliliter of semen, percent of motile sperm
 in the ejaculate, and the percent of  normal sperm.
 The goal of this study was application of objective
 measures to assess the subjective  reports of
 decreases in libido or of other sexual disturbances.
 This goal was accomplished,  but the exposures are
 poorly defined, and the number of men evaluated was
 very small. Spermatogenesis  improved in two-thirds
 of  the subjects after cessation of exposure. The
 investigators felt this supported an  argument for
 microwave influence on the observed alterations. But
 the values obtained by the semen analyses for both
 exposed and control groups might be considered to be
 low-normal or below normal values.
 Kallen et al. (1982) examined the pregnancy outcome
 of  2018  Swedish  women  who  were certified
 physiotherapists at the time of pregnancy  during
 1973  to  1978. There were 2043 infants born,
 including 25 pairs of twins. To  identify the mothers
 and the births,  the  investigators linked data  from
 computerized national registers of certified physio-
 therapists, of deliveries,  and of major congenital
 malformations. It is likely that  use of these record
 systems resulted in reasonably complete ascertain-
 ment of the study population.

 Overall, this cohort had a better than expected delivery
 outcome as measured by eight different end  points.
 All measures of infant health were near or below the
 expected national values,  which were adjusted for
 age and parity distribution, as  well as  hospital of
 delivery, to control for diagnostic variability.
To examine whether a "healthy worker" effect  may
have been operating in this group and to obtain more
detailed work histories, a nested case-control study
was also  conducted.  The cases were  defined as
infants with a major malformation plus all perinatal
deaths. Two controls  per  case were selected,
matched for maternal age, parity, and date of delivery-
After elimination of ineligible women, 33 cases and
63 controls remained. A work history questionnaire
was mailed to the study group, and the response rate
was 93 percent.

 Of several end  points  explored, the only positive
 finding was a  higher frequency of  shortwave
 equipment use among the women with a dead or
 malformed infant than among the control physiother-
 apists. This difference was statistically significant in
 both unmatched and matched pair analyses.  Use of
 ultrasound equipment followed a similar pattern but
 was not statistically significant. The authors stressed
 that exposure to the  two types of  equipment was
 heavily associated. Information on wbrk  with X rays
 was not reported, although questions on X rays were
 included in the questionnaire. Measurement of the
 radiation fields  in the work  environment was not
 performed  in conjunction with the health  study.
 Furthermore, no  information  was presented on
                       5-120

-------
shortwave  radiation levels typical of the various
physiotherapy work environments or on the types of
equipment  most commonly used  in Sweden.


5.10.5 Unresolved Issues

There are some general issues surrounding the use
and applicability  of epidemiological  research
methods to study the effects of environmental agents,
including RF radiation.  Examples  of  such  issues
include the  ability of epidemiological studies to detect
low-level risks, to  separate the effects of multiple
causes,  and  to identify  and control  confounding
factors.  Moreover, specific problems are common in
the current epidemiologic literature on RF radiation
that limit the ability to draw inferences from this body
of data  and its usefulness in establishing environ-
mental exposure limits for the general population.
These problems are briefly discussed below.


5.10.5.1 Exposure Assessment
It is difficult to determine actual exposure and dose
for individuals, and even for groups; that  is perhaps
the largest single problem in epidemiological studies
on RF radiation. In general, there  are no continuous
surveillance programs in the  workplace  that could
yield data for use in epidemiological studies.  It is a
formidable task to reconstruct RF-radiation exposure
data for  health studies that are begun after exposure
has taken  place. The  study by  Robinette and
Silverman (1980)  is a good example of attempts to
deal with the difficulties of retrospective exposure
assessment, e.g., two approaches (job type vs. Hazard
Number) were used, and an exposure gradient was
obtained with the Hazard Number which simultane-
ously considered the RF sources and the length of
service  assignment.  Despite these  efforts,  the
estimates remain  those  of  potential  rather  than
actual exposures.  In many of  the other studies, the
levels and frequencies of exposure are not known, not
estimated, or not reported. Also, in studies of military
populations, information relevant  to  exposure
conditions might be classified. When well-developed
exposure data  are  not  available, it is difficult to
analyze possible  dose-response relationships, to
interpret the significance of the findings,  and to use
the data in  establishing protective exposure limits.


5.10.5.2 Documentation and Methods

Another problem in the RF radiation literature  on
human beings relates to documentation of methods
and procedures. Degree of detail in reporting seems
to be a major  difference between studies done in the
U.S. and those conducted in  other countries. The
paper  "Guidelines  for Documentation of  Epidemio-
logic Studies"  (Epidemiology Work Group 1981)
suggests the types of topics that are  useful to
document when reporting an epidemiologic study,
especially one used to support regulatory decisions
These major elements include a  statement of
background and objectives, methods of selection and
characterization of study and comparison subjects,
data collection procedures, and analysis. Few reports
on human beings exposed to RF radiation adequately
include such necessary information. It is frequently
difficult to tell whether certain  research methods
were not applied or simply not reported. For example,
the criteria for selecting controls are frequently not
stated, but controls are often said to be comparable in
all  respects except exposure;  analyses or data to
support such statements may not be supplied. Also,
common practices such as development of standard-
ized rates or use of procedures to control confounding
variables,  e.g., age ajdustment, are  usually not
reported. Statistical  power  is rarely evaluated or
discussed,  but it is difficult to estimate power if the
underlying prevalence  or incidence of  the disease
under study is not well known, especially for some of
the  conditions and symptoms, e.g., functional
disturbances,  studied  in relation to RF radiation
exposure.
5.10.5.3 Health End Points: Design and
Populations
Another issue is the  medical  significance  of  any
changes that may be induced by exposure to RF
radiation. For example, the studies on ocular effects
usually have examined a subclinical end point, e.g.,
lens opacities, which may not necessarily be an early
marker or risk factor for cataractogenesis and visual
problems (Silverman 1979). Further, most studies
present only single measures at one point in time; the
study populations have not  been  followed long
enough to permit development of longitudinal data
upon  which  to base a  determination of whether
symptoms and subtle changes lead to disability or
disease. A similar concern surrounds the problematic
information on functional changes and nervous
system effects  reported  in some  of  the  Eastern
European literature. The potential for neurological or
behavioral  effects has  not been thoroughly  and
rigorously evaluated, and standardized questionnaires
and more objective medical measurements need to
be used in studies on these effects.


Most studies concern occupational groups of
relatively  young healthy males.  It cannot be
presumed  that  sensitivity, or lack thereof,  to  RF-
radiation exposure would be the same in the general
population as  in working groups. The general
population  is more diverse, with the full range of
ages,  sexes,  races, and other  factors that could
influence  health status  or the development of
disease. To  resolve this  issue, specific exposed
populations could be identified and evaluated if such
research appeared feasible.
                                                                     5-121

-------
5.10.5.4 Summary
Serious methodological problems in the human
studies literature make results equivocal. These
problems are not necessarily insurmountable. But at
present, the data on human beings exposed to RF
radiation are not adequate nor sufficiently developed
to be very useful in determining exposure limits for
the general population.
                       5-122

-------
                                           Section 6
                                 Summary and Conclusions
Joe A. Elder
This  document  presents  a critical review of the
literature on the biological effects of RF radiation with
the  objective of determining the  relation  between
biological effects and  exposure to RF  radiation  in
terms of (1)  absorbed dose rate (SAR) and (2)
increased body temperature. In Sec. 3, the interaction
of RF radiation with biological systems is discussed,
and in Sees. 4 and 5, research reports are reviewed
for the presence or absence of biological effects of RF
radiation. The major conclusions or generalizations
that can be drawn from these three sections are now
presented. In  addition, the biological effects  of RF
radiation are discussed in relation to  SAR and  to
increased body temperature in an attempt to analyze,
synthesize, and consolidate the available information
from this multidisciplinary field  into  a coherent
picture.

6.1  Major Conclusions and
Generalizations
1. RF radiation is a form of nonionizing electromag-
netic radiation  of very low photon energies  and
frequencies (0-3000 GHz), as distinguished from the
very high photon energies and frequencies associated
with ionizing electromagnetic radiation, e.g., X and
gamma rays. Included in the RF-radiation spectrum
are AM and FM radio, UHF and VHP TV, radar, and
microwave communication frequencies. The frequency
range of concern in this document is 0.5  MHz to 100
GHz, which includes  nearly all the significant sources
of population exposure except 60-Hz electrical power
systems. However, there is very little information on
effects in human beings at any of these frequencies
and limited data on responses of animals exposed at
frequencies above 10 GHz and below 10 MHz; most of
the animal research is concentrated in the range of
900 MHz to 3 GHz.
2. RF-energy absorption by biological systems is  a
complex  function of  frequency and the dimensions,
orientation, and dielectric properties of the absorber
and the complexity of the  incident radiation fields.
Resonant frequencies (and their related wavelengths)
for an absorber  are  those  at which maximum RF
energy  is coupled  into the  absorbing system.
Resonance occurs when the long dimension of the
absorber is approximately 0.4 times the wavelength
of the incident RF radiation, if the object is located in
free space. Under these conditions, the whole-body
resonant frequency range for humans (from adults to
infants) is approximately  30 to 300 MHz. In this
document, whole-body average specific absorption
rate (SAR) is used to normalize the rate of energy
absorption across the frequency range, 0.5 MHz to 100
GHz, and to quantitate the relation between biological
effect and dose rate of RF radiation. SAR is the mass-
normalized rate at which the energy of an electro-
magnetic field is coupled into an absorbing body; the
units of SAR are watts per kilogram  (W/kg).
3. High level  RF radiation is a source of thermal
energy (e.g., microwave  ovens) that carries all of the
known implications of heating for biological systems.
At a given incident field strength, maximal heating
occurs at the  resonant  frequency (D'Andrea  et al.
1977). In  general, the data are  consistent with the
hypothesis that  the  SAR  required  to  raise  body
temperature of laboratory animals decreases as body
mass increases.
4. In most of the animal studies that report a biological
effect of RF radiation, exposures occurred at ambient
temperatures of 20 to 25°C and relative humidities of
50  to 70  percent. At  more thermally stressful
conditions, e.g., higher ambient temperature and the
same or  higher relative  humidity, the experimental
results  show that  lower  SARs cause a similar
biological effect. For example, Rughefa/. (1974)found
that the lethal dose of 2450-MHz radiation for mice
was inversely related to the temperature-humidity
index (Figure  4-22).  Gage (1979b)  showed that a
2450-MHz exposure  at  22°C resulted in a reduced
behavioral response rate in rats at 3 W/kg, but that
similar exposures at 28°C caused reduced rates at 1,2
and 3 W/kg.

5. No consistent biological effect has yet been found
with molecular and subcellular  systems  exposed in
vitro to RF radiation other than  effects occurring at
SARs that cause general  temperature increases.
Conclusions regarding effects of in vitro exposure of
higher-order biological systems,  such as single cells
and brain tissue, are given below.

6. The electrophysiological properties of single  cells,
especially the firing  rates of neurons  in isolated
preparations, may be affected by  RF radiation at SARs
                                               6-1

-------
as low as 1 W/kg in  a manner different from
generalized heating.

7. In general, no changes in chromosomes, DNA, or
reproductive potential of RF-exposed animals have
been reported and corroborated in the absence of
significant rises of temperature. Similarly, RF
radiation does  not  appear  to cause  mutations or
genetic changes in bacterial test  systems unless
temperatures  well above the normal physiological
range are produced.

8. Effects on the  hematologic and immunologic
systems have been  reported at SARs > 0.5 W/kg;
however, there is a  lack of convincing  evidence for
RF-radiation effects on these systems without some
form of thermal involvement. Some of the reported
effects of RF radiation  on  the  hematologic and
immune systems are similar to those resulting from a
stress  response involving the hypothalamic-hypo-
physeal-adrenal axis or following administration of
glucocorticoids. In  those few cases  where  the
reversibility of RF radiation effects on the hematologic
and immunologic systems has been examined, the
effects have proved to be transient.

9. RF  radiation is teratogenic at high  SARs (> 15
W/kg) that approach lethal levels for the pregnant
animal. High maternal body temperatures are known
to be associated with birth defects. There appears to be
a threshold for the  induction of experimental birth
defects when a maternal rectal temperature of 41 to
42°C is  reached. Any agent capable of producing
elevated  internal  temperatures in  this range,
including RF radiation, is a potential teratogen.

10. Reduced fetal mass seems to occur consistently
in rodents exposed during gestation to teratogenic
levels of RF radiation, or at SARs somewhat less than
those which cause death or malformation.

11. There is evidence that exposure of rodents during
gestation to  RF radiation  may cause functional
changes  later  in life. For example, Johnson  et al.
(1978) observed lower body weight at weaning and in
young  adult rats exposed at 2.5 W/kg for 20 h/day
during  19 days of gestation, and Chernovetz et al.
(1975) found increased  postnatal  survival of mice
exposed at 38 W/kg for 10 min during  gestation.

12.  Permanent changes  in reproductive efficiency
have been directly associated with RF-radiation
exposures that caused temperatures in animal testes
greater than 45°C. At temperatures of 37 to 42°C
mature sperm may be killed with a temporary loss of
spermatogenic epithelium. Irradiation of rats  at an
SAR of 5.6 W/kg, which produced a core temperature
of 41 °C, resulted in temporary infertility.

13.  Neurons in the central nervous system (CNS) of
experimental animals have been reported to be
altered by acute high-level and by chronic low-level
exposures (> 2 W/kg). Pulsed RF radiation may have
a potentiating effect on drugs that affect nervous
system function. Some of the early reports of RF-
radiation effects on the blood-brain barrier (BBB) at
SARs < 2 W/kg have not been substantiated by later
investigations.

14. An increased mobilization of calcium ions occurs
in  brain  tissue exposed  in vitro  to  RF radiation,
amplitude modulated at frequencies recorded in the
electroencephalogram (EEC) of awake animals. The
response appears to be based on the intensity of the
electric field within the tissue, which can be related to
SAR; the lowest effective SAR in in vitro samples is
estimated to be 0.0013 W/kg> Calcium-ion efflux is a
nonlinear effect in terms of both AM frequency and
field intensity; that is, the response occurs at specific
frequencies  and  electric field  strengths. The
physiological significance of this effect has not been
established.
15. Some types of animal behavior are disrupted at
SARs that are approximately 25 to 50 percent of the
resting metabolic rates of many species. For example,
changes in locomotor behavior in rats occur at an
SAR of 1.2 W/kg, and alterations in thermoregulatory
behavior in squirrel monkeys occur at an SAR of 1
W/kg. Decreases in other  operant or  learned
behavioral  responses during  exposure have been
found at an SAR of 2.5 W/kg in the rat and at 5.0
W/kg in the rhesus monkey. The reported behavioral
alterations appear to be reversible with time.
16. Changes  reported in endocrine gland  function
and blood chemistry  are  similar to those observed
during increased thermoregulatory activity and heat
stress, and are generally associated with SARs > 1
W/kg. Exposures of sufficient intensity to produce
whole-body heating produce an increase in heart rate
similar to that caused by heating from other sources.
Changes in  whole-body metabolism  have been
reported following exposures at thermal levels (~ 10
W/kg), and brain energy metabolism  is altered at
levels as low as 0.1 W/kg following irradiation of the
exposed surface of the brain of anesthesized animals.
17. A single acute exposure of the eye  to high-
intensity RF radiation, if applied for a sufficient time,
is cataractogenic in some experimental animals. In
the rabbit,  the  animal most  often used in ocular
studies, the cataractogenic threshold for a  100-min
exposure is 150 m W/cm2 (138 W/kg peak absorption
in  the  lens).  The cataractogenic  potential of
microwave radiation varies with frequency; the most
effective frequencies for cataracts in the rabbit eye
appear to be in the 1 - to 10 GHz range. Cataracts were
not produced in primates exposed acutely to RF-
radiation conditions that  caused cataracts in lower
mammals  such as  the rabbit.  The absence of
cataracts in the primate is attributed to the different
facial structure that  caused a different pattern of
absorbed energy in the eye. No cataracts have been
reported in rabbits after whole-body, far-field RF-
                       6-2

-------
 radiation exposures, even at near-lethal levels (SAR
 = 42 W/kg for 15 min). No data at present support a
 conclusion  that low-level, chronic  exposure to
 microwave  radiation  induces cataracts in  human
 beings, although  some studies have associated
 ocular-lens defects with  microwave radiation
 exposure.
 18. Pulsed RF radiation in the range 216 to 6500 MHz
 can  be heard by some  human  beings. The sound
 associated with the "RF hearing" varies with pulse
 width and pulse-repetition rate and is described as a
 click,  buzz,  or  chirp. The  threshold for  human
 perception of this effect  is approximately 40/uJ/cm2
 (incident energy density per pulse). The most
 generally accepted  mechanism for the RF-auditory
 sensation is that  the  incident pulse induces a
 minuscule but rapid thermoelastic expansion within
 the  skull, which results in a pressure wave that is
 conducted by the bone to the cochlear region of the
 ear.
 19.  For the broad range  of frequencies between 0.5
 MHz and 100 GHz, cutaneous perception of heat and
 thermal pain may be an unreliable sensory mechanism
 for  protection against potentially harmful  RF-
 radiation exposure levels. Many frequencies deposit
 most of their energy at depths below the cutaneous
 thermal receptors.
 20. There is no convincing evidence that exposure to RF
 radiation shortens the life span of human beings or
 experimental animals or that RF radiation is a primary
 carcinogen (cancer inducer); however, (1)fewstudies
 have used  longevity  or cancer  incidence  as  end
 points, and (2) human studies have lacked statistical
 power to exclude life shortening or cancer. There is
 evidence from one group of investigators that chronic
 exposure to RF radiation (SAR = 2 to 3 W/kg) resulted
 in cancer promotion or  co-carcinogenesis in three
 different tumor systems in mice; the  incidence of
 cancer  was comparable to that  observed in mice
 exposed to chronic stress conditions only.
 21. Human data are currently limited and incomplete
 but do not indicate any obvious relationship between
 prolonged low-level RF-radiation  exposure  and
 increased mortality or morbidity, including  cancer
 incidence.

The  prospects for revision and  refinement  of  the
major conclusions and generalizations stated above
are considerable because of our limited knowledge of
(1) effects of most frequencies in the range 0.5  MHz to
100 GHz; (2) effects of chronic low-level exposures on
human  beings and  laboratory animals; (3)  which
segments of the population are most sensitive; (4) the
influence of ambient environmental conditions and of
potential synergistic interaction with other agents;
(5) the implication of  nonhomogeneous RF-energy
deposition; (6) the  existence and significance of
frequency-specific effects and power-density
 windows;  and  (7) the  physical mechanisms of
 interaction at low exposure levels,  including field-
 specific phenomena.

 6.2 Specific Absorption Rate
 Reports on the biological effects of  RF radiation
 usually specify the frequency and power density of
 the field applied to the biological system. Individually
 or  together,  these  parameters do  not provide  a
 reasonable correlate with biological effects because
 RF-energy  absorption is  known to depend  on  the
 relation between wavelength and absorber size and
 orientation. However, when frequency and power
 density are combined with a knowledge of absorber
 size and dielectric property, an estimate of SAR can
 be made. The whole-body average SAR is currently of
 limited use because it is an  estimate of absorbed
 energy averaged over the whole-body mass. It does
 not address the existence of localized areas of
 increased energy deposition  in exposed biological
 systems, nor the considerable differences between
 species in their capacity to regulate  a given energy
 burden. However, work is currently under way to
 extend the  usefulness of this concept in both areas.
 The SAR is currently the parameter used most
 frequently  to  describe the energy absorbed by a
 biological system exposed to RF  radiation. In  the
 discussion below, whole-body average SAR is used as
 a correlate  for the observed biological responses.
 The data cited in the preceding sections show that
 acute whole-body exposure of laboratory animals at
 SAR  values  > 30  W/kg is  lethal. For  example,
 Chernovetz  et  al.  (1977) reported that a 20-min
 exposure (2450-MHz) at 31  W/kg was lethal to 23
 percent of the rats. Appleton et al. (1975) found that
 3000-MHz  exposures at 42 W/kg for 30 min or 70
 W/kg for 15 min were lethal to rabbits. A shorter
 exposure (15 min) at 42 W/kg caused  acute ocular
 changes, but  no cataracts, and  cutaneous burns
 around the eye (see Table 5-15). During a 15-min
 exposure at 14 W/kg, rabbits showed signs of acute
 heat stress and struggled out  of the field. Localized
 exposure to very high intensity RF radiation can result
 in burns, as  shown by Kramar et al.  (1978), who
 reported second- to third-degree burns on the face of
 rhesus monkeys exposed for 22 min to 2450-MHz
 radiation that produced an SAR in the eye of ~ 115
 W/kg.

 Brief exposures (< 5 min) of the whole  body to high
 intensities result in significant  biological damage, as
shown by Rugh (1976a)and Hugh era/. (1975), who
 irradiated pregnant  mice  at SAR values near 100
W/kg and found birth defects, increased embryonic
and fetal  resorptions, postnatal weight  decrements,
and reduced survival  upon re-irradiation  of the
offspring of mice exposed during pregnancy.
 Laboratory animals have generally been found to be
affected by  SARs in the 10- to 30-W/kg range, even
                                                                      6-3

-------
when the exposures have been  short (minutes to
hours). Fetotoxic effects in mice exposed at 22 W/kg
have been reported by Berman et al. (1978). Some
components of the immune system appear to react
with  an increased response,  whereas others
demonstrate diminished responses. Components
reported to  increase are PMN levels  (Kitsovskaya
1964; Michaelsoh  et al. 1964; Lappenbusch et al.
1973; Liburdy 1977), splenic lymphocytes (Wiktor-
Jedrzejczak  et  al.  1977a,b,c; Sulek et al.  1980),
lymphocyte response to mitogen stimulation (Huang
and Mold 1980; Wiktor-Jedrzejczak era/. 1977a,b,c),
and lymphocyte transformation to the lymphoblast
stage (Huang et al. 1977). Decreases  are reported
in the primary antibody response to sheep  red blood
cell (SRBC) (Wiktor-Jedrzejczak et al. 1977a,b,c), the
CPU for the erythroid granulocyte-macrophage series
in bone marrow (Huang and Mold  1980), lymphocyte
traffic from lung  to spleen  (Liburdy 1980), and
circulating lymphocyte  levels (Lappenbusch et al.
1973;  Liburdy 1977). Roszkowski et al. (1980)
reported a general  immunosuppressive effect at 35
W/kg; they found an increase in lung cancer colonies
and an inhibition of contact sensitivity to oxazlone in
mice.

Some biological end points appear to be unaffected
even  at  10  to  30 W/kg.  Examples are  postnatal
survival, adult body weight, and longevity (Guillet and
Michaelson 1977; Spalding et al.  1971).
In contrast to the results of animal  studies, in vitro
studies at 10 to 30  W/kg have shown few effects on
the irradiated  systems if temperature is  properly
controlled, e.g., no mutation  induction in bacteria
(Blackman et al. 1976; Corelli et al. 1977; Dutta et al.
1979a, 1980), no effects on enzyme activities (Ward
et al. 1975; Bini et al. 1978; Allis and Fromme 1979),
no  effects on  the physical characteristics  and
structure of biomolecules such as nucleic acids and
proteins (Allis  1975; Allis et al. 1976;  Corelli et al.
1977), and no change in lymphocyte transformation
(Smialowicz 1976).  Notable exceptions are Ismailov's
(1971, 1977, 1978) reports  that the  human  RBC
exhibits increased electrophoretic  mobility,  K+ efflux,
Na* influx, and hydrogen exchange at SARs between
5 and 45 W/kg; and Seaman and Wachtel's (1978)
observation  of a decreased  firing  rate of Aplysia
pacemaker neurons at  SARs as low as  1 W/kg.

The majority of the experimental studies that met the
criteria for inclusion in this document employed SARs
< 10 W/kg. Many of these studies are differentiated
by observed  "effects"  (without regard  to biological
significance) or "no effects," and arranged within
these two groupings by decreasing SAR values in
Tables 6-1 and 6-2.

Biological variables  that do not appear to be  sensitive
to RF radiation at < 10 W/kg include organ weight,
litter size, teratology and growth (Table  6-1).
Physiological systems and parameters that appear to
be sensitive include behavior, the central nervous
system, and hematology and immunology (Table 6-2).
In Table 6-2, several positive findings at SARs < 10
W/kg are listed for a number of biological variables.
A comparison of Table 6-1 with Table 6-2 reveals that
many reports  of  "no effects" occur in the same
biological systems (e.g., behavior and hematology/
immunology)  over a similar SAR range. Negative
findings are important to define the lower exposure
limits of biological effects and  the sensitivity of
specific biological end points; however, in general,
they cannot displace the reports of positive findings
that  are  the  principal concern in a  review of the
biological effects of RF radiation.

It seems appropriate at this point to cite  the
conclusions of the subcommittee that developed the
rationale for  the recently published ANSI  RF-
radiation (0.3 MHz to 100 GHz) exposure guidelines
(ANSI  1982). The subcommittee  completed  its
review of the literature in February 1979. Note that
about half of the reports on behavior shown in Tables
6-1 and 6-2 were  published prior to 1979. The ANSI
subcommittee  concluded the following:

  The most sensitive measures of biological effects
  were found to be based on behavior.
  The whole-body-averaged SARs associated with
  thresholds of reversible behavioral disruption were
  found to range normally between 4 and 8 W/kg in
  spite of considerable differences in carrier
  frequency (600 MHz to 2.45 GHz), species (rodents
  versus primates), and mode of  irradiation (cavity,
  waveguide, and plane wave).

  Because  of the paucity of reliable data on chronic
  exposures, the subcommittee focused on evidence
  of  behavioral disruption under acute exposures,
  even that of a transient  and fully reversible
  character.  The  assumption is  that  reversible
  disruption during an acute exposure is tantamount
  to  irreversible injury during chronic exposure.
  The concensus remained that reliable evidence of
  hazardous effects is associated with whole-body-
  averaged SARs  above 4 W/kg.

Other significant effects in laboratory animals which
occur at SARs of 4 to 8 W/kg are:
1. Temporary sterility in male rats exposed at an SAR
   =  5.6 W/kg for 4 h/day for 20 days (Berman et al.
   1980).

2. Bradycardia in  rats after whole-body exposure
  (SAR = 6.5 W/kg) (Phillips et al. 1975b). Exposure
  of the head of the rabbit at an SAR of 3  W/kg
   caused tachycardia (Birenbaum et al. 1975).
Potentially  significant biological  effects that have
been reported at SARs < 4 W/kg include  the
following:
                       6-4

-------
Table 6-1.    Selected Studies Reporting

Biological Variable
'No Effects" at SARs < 10 W/kg Grouped by Biological Variable

         Relevant Studies
SARs
Growth
(Food, Water Intake)





Behavior










Mortality/Life Span


Hematology/lmm unology





Mutations/Chromosomal
Aberrations

Teratology

Litter Size

Organ Weight

Blood Chemistry
Fertility
Hormones



Neurotransmitter Levels
Metabolic Rate
Cardiovascular System
Stavinoha et al. (1975)
McAfee et al. (1973)
Kaplan (1981)
Michaelson et al. (1978)
Johnson et al. (1978)
Lovely et al. (1977)
D'Andrea et al. (1979)
D'Andrea et al. (1976)
Moe era/. (1976)
Mitchell etal. (1977)
Sanza and de Lorge (1977)
Gage(1979a)
Scholl and Allen (1979)
de Lorge (1976)
Lovely etal. (1977)
Robert! era/. (1975)
Schrotefa/. (1980)
Thomas etal. (1979)
Kaplan (1981)
Johnson etal. (1978)
Spaldingef a/. (1971)
Smialowicz et al. (1981 a)
Smialowicz etal. (19796)
Liburdy(1980)
Smialowicz (1981 a)
Spa Iding etal. (1971)
Djordjevic et a/ (1977)
Blackman et al. (1975, 1976)
McLees et al. (1972)
Bermanera/. (1980)
Johnson et al. (1978)
Bermanera/. (1978)
Michaelson et al. (1978)
Johnson etal. (1978)
Michaelson et al. (1978)
Mikolajczyk(1976)
Wangemann and Cleary (1 976)
Berman etal. (1980)
Mikolajczyk(1976)
Parker (1973)
Lu etal. (1977)
Milroy and Michaelson (1972)
Merritt etal. (1976)
Ho and Edwards (1977a,b)
Phillips et al. (1975b)
6.3
6.0-8.0
3.4
2.5-10.0
2.5
1 .0-2.5
1.0-2.5
5.0-6.0
3.9
2.3
2.1-4.7
0.3
1.6
1.1-1.4
1.0
0.2-8.3
0.2
0.2
3.4
2.5
1.7
10
>4.0
3.8
2.0-3.0
1.7
1.0
0.08-7.5
1.3
0.9
2.5
2.0-8.1
2.5-10.0
2.5
2.5-10.0
1 .0-2.0
0.8
1 .0-2.0
1 .0-2.0
2.0-6.5
2.5
0.25-2.5
3.0
5.5
4.5
1. The decrease in behavioral response rates cited in
  ANSI (1982) were based  on studies done at
  ambient temperatures of  20 to 25°C.  Gage
  (1979b) has shown  that  similar changes in
  behavior occur at lower SARs when exposures are
  conducted at higher ambient temperature; that is,
  at 22°C the effective SAR was 3 W/kg, whereas at
  28°C SARs of 1 and 2 W/kg were effective.

2. Although RF radiation does not appear to be a
  primary carcinogenic agent (cancer inducer),
  there is evidence from one laboratory that RF
                     radiation acts as  a cancer promoter or co-
                     carcinogen in three different tumor systems in
                     mice at an SAR of 2 to 3 W/kg (Szmigielski et al.
                     1980, 1982).

                  3.  A decrease in the number of Purkinje cells in the
                     brain of rats  exposed at an SAR of 2 W/kg was
                     reported by Albert et al. (1981 a).

                  4.  Endocrine gland  function and  blood chemistry
                     changes are similar to those observed during heat
                     stress and are generally associated with SARs>1
                     W/kg (Sec. 5.7.1).
                                                                      6-5

-------
Table 6-2.    Selected Studies with Reported
Biological Variable
'Effects" at SARs < 10 W/kg Grouped by Biological Variable

       Relevant Studies
SARs
Behavior
Central Nervous System
Hematology/lmmunology
Hormones


Drug Potentiation

M utations/Chromosome
  Aberrations
Neurotransmitter Levels
Fertility
Clinical Chemistry
Cardiovascular System
       Gordon (1983a)
       Gordon (1983b)
       deLorge(1976)
       Moeeta/ (1976)
       Gage(1979a)
       de Lorge and Ezell (1980)
       D'Andrea et at. (1980)
       Mitchell etal. (1977)
       Gage(1979b)
       Thomas etal. (1975)
       D'Andrea et a/. (1979)
       Stern etal. (1979)
       Thomas et at. (1976)
       Rudnevef a/. (1978)
       Adair and Adams (1980a,b)
       Schrot et al. (1980)
       Frey and Feld(1975)
       Albert and DeSantis (1975, 1976)
       Switzer and Mitchell (1977)
       Albert (1979b)
       Albert et al.  (1981 a)
       Seaman and Wachtel (1978)
       Albert and Kerns (1981)
       Sanders et al. (1980)
       Goldstein and Sisko (1974)
       BaranskM 19726)
       Blackman et al. (1979, 1980a,b)
       Takashima et al. (1979)
       Sulekera/ (1980)
       Smialowicz et al. (1979a, 1982)
       Liburdy(1979)
       Michaelson  et al. (1964)
       Huang and Mold (1980)
       Huang et al. (1977)
       Djordjevic and Kolak (1973)
       Deichmann et al. (1963)
       McRee  et al. (1980a)
       Czerski  (1975)
       Szmigielski et al. (1975)
       Baranski(1971)
       Prince et al.  (1972)
       Lu etal. (1977)
       Lotz and Michaelson (1978)
       Mikolajczyk (1976)
       Edelwejn (1968)
       Thomas and Maitland (1979)
       Manikowska et al. (1979)

       Merritt  et al. (1977)
       Berman et al. (1980)
       Wangemann and Cleary (1976)
       Kaplan  et al. (1971)
       Phillips etal. (1975b)
       Birenbaum etal. (1975)
7.0
5.3
5.0
3.6
27
2.5-4.9
2.5
2.3
2.0
1.4-1.5
1.2
1.1
1.1
1.0
1.0
0.7
0.4

5.0-12.5
2.3
2.0
2.0
1.0
0.9-2.0
0.1
0.1-0.3
0.4-2.5
0.0013
0.0001
5.0-11.8
5.0-7.0
4.6
4.0-6.0
3.6-10.0
2.3-30.7
2.0
1.5-2.2
1.5
0.5-0.8
0.5
0.5
0.4-2.0
5.0
3.0-12.8
1.0-2.0
1.0
0.2
0.05-5.0

6.0
5.6
1.6-4.0
8.0
6.5
3.0
 5.  Effects on  the hematologic  and  immunologic
    systems occur at SARs > 0.5 W/kg and appear to
    resu It from some form of therma I i nvolvement due
    to absorbed RF energy (Sec. 5.2).
 6.  Changes in cellular energy metabolism in the  rat
    brain have been reported at an SAR =0.1 W/kg;
                    the data support the conclusion that the effect is
                    frequency specific (Sanders et al. 1980).
                 7.  Results of studies of amplitude-modulated (AM)
                    radio-waves, particularly AM frequencies near or
                    at  16  Hz, have shown changes in  calcium-ion
                    efflux from chick brain tissues in vitro. The effect
                           6-5

-------
   has been shown to be frequency and  intensity
   specific and to occur at SARs as low as 0.0013
   W/kg (Blackman et at. 1979, 1980a,b). In  1984
   Dutta et al. reported that 16-Hz AM microwave
   radiation  caused changes in calcium-ion efflux
   from human brain cells in culture at an SAR of 0.5
   W/kg.

Human data are limited and not useful for developing
a quantitative relation between  effect and  SAR;
however, two recent  studies of physiotherapists
suggest two  potentially significant health effects
associated with work with RF equipment. The first is
heart disease (primarily ischemic)  in males; (Ham-
burger et al. 1983) the second is pregnancy outcomes
in female physiotherapists (Kallen et al. 1982). Both
studies are considered to be exploratory, and neither
provides quantitative data on the RF radiation levels
in the work environment, so that SAR values cannot
be estimated.
In  summary,  the data  currently  available on the
relation of SAR to biological effect show evidence for
biological effects at an  SAR of about 1 W/kg. This
value is lower by a factor of 4 than 4 W/kg, the value
above which reliable evidence of hazardous effects
was found by ANSI (1982) following a review of the
literature in February 1979. The above conclusion is
based on:
1.  the  findings that  more thermally stressful
   conditions result in lower threshold  SARs  for
   behavioral changes similar to those  changes
   determined by ANSI  (1982)  to be the  most
   sensitive measures of biological effects
2.  the effects on endocrine gland function, blood
   chemistry, hematology, and immunology that
   appear to result from some  form of thermal
   involvement due to  absorbed RF energy
3.  data from  one  laboratory showing  that  RF
   radiation  can act as a cancer  promoter or co-
   carcinogen and results from another laboratory
   describing changes in brain cellularity.
The experimental evidence suggests that the central
nervous system  is particularly sensitive to  RF
radiation. Two other areas of research that may prove
to  be highly significant are calcium-ion efflux and
brain energy metabolism.

6.3 Core Temperature
Heating is  the  least controversial explanation for
most  RF-radiation  effects, and it is appropriate  to
examine the biological effects in relation to an
increase in body temperature in the range of 0 to 6°C
above normal. The average rectal (core) temperature
is approximately  37.0°C (98.6°F) for human beings
and  ranges  from 36 to 38°C for most mammals
(Schmidt-Nielsen  1979). In  general,  the lethal
temperature is approximately 6°C above the average
core temperature (cf. Table 4-2). Prolonged elevation
of core temprature at  5°C above normal (42CC  -
107°F)  is associated with  heat  stroke  and brain
lesions; the temperature 41,2°C (106.2°F) occurs m
only 1 of 1000 humans during fever (Folk 1974) With
RF-radiation exposures, even brief periods (1  5 to 20
min) at core temperatures of 41.5 to 42 5°C in rats
can result in  increased fetal resorptions,  decreased
fetal body weights (Chernovetz  et al. 1977), and
significant hemolysis and K" efflux in red blood cells
of adult rats (Peterson et al. 1979). In healthy young
men whose mean body temperature was increased to
41 °C by RF-radiation exposure for up to 3 h, sperm
numbers had decreased by 60 percent by 40 to 60
days after treatment (MacLeod and Hotchkiss 1941).
At this core temperature, temporary infertility in male
rats has also  been  reported (Berman et al. 1980)
Berman et al. (1981 (showed that when RF-radiation
exposure produced  a maternal colonic temperature of
<41°C in the pregnant rat, no effects were detectable
in the fetus; O'Connor (1980)  reached a  similar
conclusion. The mouse fetus  is apparently more
sensitive to  increased  maternal  body temperature
than is the fetal rat.  Berman era/. (1978) reported that
a 0.8°C temperature differential  between exposed
and  sham-irradiated dams produced a  10-percent
decrease in  the body mass of the exposed  mouse
fetus.
Several RF-radiation-induced effects have been
reported for other  stages of life  in various  animal
species at core temperature increases of 1 to 3°C.
These effects include an increased response of  the
adrenal glands in irradiated infant rats (Guillet and
Michaelson  1977); increased serum glucose and
blood urea nitrogen (BUN)  levels in adult  rabbits
(Wangemann and Cleary  1976);  increases in
circulating neutrophils.and T and B lymphocytes in
spleens of mice, along with a decrease in circulating
lymphocytes  (Liburdy  1977, 1979); changes in
lymphocyte circulation between bone marrow,
spleen,  liver,  and lung  similar to that produced by
stress-related steroids  (Liburdy  1980);  increased
responsiveness of lymphocytes to mitogen stimulation
in monkeys (Prince et al.  1972); increased white blood
cell (WBC) count and decreased S9Fe uptake in mice
(Rotkovska and Vacek 1975); decreased exploratory
activity of rats (Hunt et al. 1975); decreased vigilance
in monkeys (de Lorge 1976); and work stoppage in
rats (D'Andrea et al. 1977).

RF-radiation exposures that produced core temperature
increases of 1 to 2°C are reported to be associated
with decreased serum thyroxine and  increased
corticosteroid levels (Lu era/. 1977). Michaelson era/.
(1964)  found decreased lymphocyte, neutrophil, and
eosinophil levels and hemoconcentration at tempera-
tures of 1.0  to 1.7°C  above normal  in dogs. In
association with a temperature  rise of 1.0°C in
irradiated rats, Djordjevic and Kolak (1973) observed
                                                                      5-7

-------
increases in RBC. hematocrit, and hemoglobin which
may also be a case of hemoconcentration. Pazderova-
Vejlupkova and Josifko (1979) reported decreases in
the hematocrit,  WBC,  and lymphocyte numbers in
rats at a temperature change of only +0.5°C.
As a  general  rule,  exposures that produce a core-
temperature rise of £ 0.5-C do not cause detectable
effects  on reproduction, fetal  weight,  growth,
development, hematological and immunological end
points, hormonal levels, and clinical blood chemistry
(Berman et al.  1978,  1980; Johnson et  at.  1978;
Michaelson et  al.  1978; Guillet and  Michaelson
1977; Djordjevic et al. 1977; Smialowicz eta/. 1979b;
Milroy and Michaelson 1972,  Lovely et  al.  1977;
Wangemann and Cleary 1976).
In most of the animal studies that  report a biological
effect of RF radiation,  the  exposures occurred at
ambient temperatures  of 20 to 25°C and relative
humidities of  50 to 70 percent. At more  thermally
stressful conditions, e.g., higher ambient temperature
and  the  same  or  higher relative humidity,  the
experimental results show that lower SARs will cause a
similar  biological effect.  In  other words, it is
reasonable to expect that all  the above  biological
effects described in Sec. 63 that are associated with
an increase of core temperature > 0.5°C will occur
also at lower SARs if the combination of RF radiation
exposure and  ambient  conditions  result in a similar
core-temperature increase.
In Sec. 6.2, the biological effects relative to SAR were
presented, and the important conclusions from ANSI
(1982) relative to  hazardous levels of RF radiation
were  discussed. In this section, it is instructive to
use the threshold  limit value (TLV) for deep-body
temperature  of  workers  as an  introduction to  a
discussion of temperature changes in human beings
exposed to RF radiation. The TLV of  the American
Conference of Governmental Industrial Hygienists
(1983) states that workers should not be permitted to
continue their work when their  deep-body tempera-
ture exceeds 38°C  , that is, 1 °C above the average
normal temperature for adult  human beings. One
may  conclude  therefore  that  exposure  to  an
environmental agent such as RF radiation that may
cause a  1°C  rise  in core temperature  should be
considered hazardous to relatively healthy individuals.
The  application  of this  reasoning to the general
public, which varies in health  status, age, etc., is
discussed below.

There are few  data on the RF-radiation  exposure
conditions that cause core temperature changes of 0
to  1 °C in human beings. One of the first attempts to
consider the addition of RF energy as a thermal load in
a model of man  was done by Guy  et al. (1973), who
later  used a thermodynamic model of the human
body to determine increase in temperature (Guy et al.
1978). They reported that an exposure of 4 W/kg at
79 MHz (resonant  frequency for  an adult human
being in free space) for about two hours caused the
hypothalamic temperature to rise 1°C. If no unusual
hot spot occurred in the hypothalamic area, one could
assume that the temperature within the head was a
good estimate of core temperature, that is, the core
temperature rose about 1&C also. Using a mathema-
tical model of the human body, Spiegel etal. (1980a)
predicted than an 80-MHz exposure (Ta = 30CC, RH =
30 percent) at an SAR - 2.3 W/kg would produce a
1 ~C rise in core temperature; and SAR of 1.4 W/kg
resulted in a 0.5'C rise in core temperature. In their
report, in which they examined the  relation  of
biological effects to  increase in body temperature.
Tell and Harlen (1979)  concluded that whole-body
exposure of  human  beings to 3 W/kg at resonant
frequencies would raise a resting individual's rectal
temperature by about 1 °C for exposure durations of
the order of 1 h or more. In summary, the results of
Guy et al. (1978), Spiegel et al. (1980a), and Tell and
Harlen (1979) show that SARs of 1 to 4 W/kg for short
durations (1 h) produce significant increases in
human body temperature at ambient temperatures of
25 to 30°C It is reasonable to conclude that increases
in body temperature are likely to occur at lower SARs
if exposure takes place  under more thermally
stressful  conditions,  e.g., higher ambient tempera-
ture and/or higher relative  humidity.

Data from experimental studies with primates are in
good agreement with the  results of the modeling
experiments described  above.  In 1976, de Lorge
found that a 2450-MHz exposure of the  rhesus
monkey for 1 h (Ta = 21 to 24°C; RH = 70 percent) at an
SAR  of 2.2 to 2.9 produced a temperature  rise of
0.5°C; an SAR of 4.3 W/kg caused an increase of
1 °C.  Exposure of the rhesus monkey to 225 MHz, a
frequency close to resonance, at only 1.2 W/kg (Ta =
24°C) caused a 0.5 to 0.6°C  rise in rectal temperature;
at  1290 MHz, an  SAR of 3 W/kg caused a  similar
temperature rise  (Lotz  1982;  Lotz and Podgorski
1982). In these two studies, the ability of the animals
to  dissipate heat may  have been compromised
because  of  various  factors in the experimental
design.

Two additional points must be  given serious
consideration.
1.  When a measurable increase in body temperature
   occurs, it can be interpreted as an indication that
   the body is under some stress.  Exposure to RF
   radiation at  SARs below those that cause an
   increase  in core  temperature may activate the
   heat-dissipating thermoregulatory mechanisms
   of the body to maintain  core temperature within
   the normal range.
2.  Under many conditions of  RF exposure, some
   regions of the body will probably experience an
   increase  in  temperature due to  localized RF-
   energy absorption while the core  temperature
                       6-8

-------
   remains near normal (Guy et al. 1978; Spiegel et
   al. 1980a). For example,  Spiegel et al. (1980a)
   showed that an SAR  of 1.4 W/kg produced  a
   3.6°C rise in temperature of the thigh while the
   core temperature rose only 0.5°C.
Consideration is now given to the health impact of the
thermal extremes that  can exist in the U.S. during
summer months. During  the period 1952-67, heat
waves occurred in five different years. Ellis (1972)
examined  the Vital Statistics Reports of the U.S.
Public Health Service for  these years  and drew the
following conclusions:
1. During the five heat-wave years  1952-55 and
   1966, excess deaths due to heat-aggravated or
   heat-precipitated illness during the heat wave
   were conservatively estimated to be at least 10
   times those to be expected in this category in
   the General Mortality Tables.
2. In either June or July there were excess numbers
   of deaths due to vascular accidents of the central
   nervous system in 1952-55 and 1966, and similar
   but smaller excess numbers of diabetic deaths in
   1952, 1954, 1955, and 1966.
3. Excess deaths from  arteriosclerotic and degene-
   rative heart disease, including coronary disease,
   were observed in 1955 and 1966, and there was
   an  excess of deaths  from hypertensive heart
   disease in May, June, or July in each of the heat-
   wave years but not in  10 of the other 11  years.
4. Infants below 1 year of age are the most heat-
   illness-prone age group below 50  years  of age;
   adults  above 50 years are more heat-illness-
   prone  than infants and  become  progressively
   more so with advancing age.

The  evidence indicates that  exposure of  human
beings at  frequencies in the resonant region at an
SAR of approximately 1 W/kg  produces significant
changes in body temperature  under some environ-
mental conditions. This conclusion should  not be
interpreted as conservative because, as stated above:
1. when an increase in body temperature occurs, it
   can be  interpreted as an indication that the body is
   under some stress;
2. under many conditions of RF exposure, it is very
   likely that some regions of the body increase in
   temperature due to localized RF-energy absorption
   while the core temperature remains near normal;
3. the general population has groups  of individuals
   particularly susceptible to heat.

6.4 Summary

The review of the currently available literature on RF
radiation provides evidence that biological  effects
occur at an SAR of about 1 W/kg; some of them may
be significant under certain environmental conditions.
This value was derived from two perspectives; (1) the
relation of biological effects to SAR (dose rate) and (2)
the relation of biological effects  to increased  body
temperature caused by absorption of RF energy.
                                                                       6-9

-------

-------
                                         References
Abramson, D.I., A.J. Harris, and  P. Beaconsfield.
  1957. Changes in Peripheral Blood Flow Produced
  by Short-Wave  Diathermy. Arch. Phys.  Med.
  Rehab., 38:369-376.
Adair, E.R. 1971. Displacements of Rectal Tempera-
  ture Modify Behavioral Thermoregulation. Physiol.
  Behav., 7:21-26.
Adair, E.R. 1976. Autonomic Thermoregulation in
  Squirrel Monkey when Behavioral Regulation is
  Limited. J. Appl. Physiol., 40:94-700.
Adair, E.R. 1981. Microwaves and Thermoregulation.
  In: USAF  Radiofrequency Radiation Bioeffects
  Research Program—A Review, -J.C. Mitchell, ed.
  Review 4-81, USAF School of Aerospace Medicine,
  San Antonio, Texas, pp.145-158.
Adair, E.R., and B.W. Adams.  1980a. Microwaves
  Induce Peripheral Vasodilation in Squirrel Monkey.
  Science, 207:1381-1383.
Adair, E.R., and B.W. Adams.  1980b. Microwaves
  Modify Thermoregulatory Behavior in  Squirrel
  Monkey. Bioelectromagnetics, 1:1-20.
Adair, E.R., and B.W. Adams. 1982a. Adjustments in
  Metabolic  Heat Production by Squirrel Monkeys
  Exposed to Microwaves. J. Appl. Physiol., 52(4):
  1049-1058.
Adair, E.R.,  and B.W.  Adams. 1982b.  Behavioral
  Thermoregulation in the Squirrel Monkey: Adapta-
  tion Process During Prolonged Microwave Exposure.
  Behav. Neurosci., 97:49-61.
Adair, E.R.,  D.E. Spiers,  J.A.J. Stolwijk, and C.B.
  Wenger.  1983. Technical Note:  On Changes in
  Evaporative Heat Loss that Result from Exposure to
  Nonionizing Electromagnetic  Radiation. J. Micro-
  wave Power, 18(2):209-211.
Adey, W.R. 1981. Ionic  Nonequilibrium Phenomena
  in Tissue Interactions With Electromagnetic Fields.
  In: Biological Effects of Nonionizing Radiation, K.H.
  Illinger, ed. ACS Symposium Series, 157:271 -297.

Adey, W.R., S.M. Bawin, and A.F. Lawrence. 1982.
  Effects of Weak Amplitude-Modulated Microwave
  Fields on Calcium Efflux From Awake Cat Cerebral
  Cortex. Bioelectromagnetics, 3:295-307.
Airborne Instruments Laboratory. 1956. An Observa-
  tion on  the Detection by the Ear of Microwave
  Signals. Proc. IRE, 44 (Oct.):2A.
Alam, M.T., N. Barthakur, N.G. Lambert, and S.S,
  Kasatiya. 1978. Cytological Effects of Microwave
  Radiation in Chinese Hamster Cells In Vitro. Can. J.
  Genet. Cytol., 20:23-30.
Albert, E.N.  1977. Light and Electron Microscopic
  Observations on  the Blood-Brain  Barrier after
  Microwave Irradiation. In: Symposium on Biological
  Effects and  Measurement of Radio Frequency/
  Microwaves, D.G. Hazzard, ed. HEW Publication
  (FDA) 77-8026,  Rockville, Maryland, pp. 294-304.

Albert, E.N. 1979a. Current Status of Microwave
  Effects on the Blood-Brain Barrier. J. Microwave
  Power, 14:281-285.

Albert,  E.N. 1979b. Reversibility  of Microwave-
  Induced Blood-Brain Barrier  Permeability. Radio
  Sci., 14:323-327.
Albert, E.N., and M. DeSantis. 1975. Do Microwaves
  Alter Nervous System Structure? Ann. N.Y. Acad.
  Sci., 247:87-108.
Albert, E.N., and  M. DeSantis. 1976. Histological
  Observations on  Central  Nervous System: In:
  Biological Effects of Electromagnetic Waves, Vol. I,
  C.C. Johnson and M.L Shore, eds. HEW Publication
  (FDA) 77-8010,  Rockville, Maryland, pp. 299-310.
Albert E.N., and J.M.  Kerns. 1981.  Reversible
  Microwave Effects on  the  Blood-Brain  Barrier.
  Brain  Res., 230:153-164.

Albert, E., C. Blackman, and F. Slaby. 1980. Calcium
  Dependent Secretory Protein Release and Calcium
  Efflux  During RF Irradiation of  Rat  Pancreatic
  Tissue Slices.  In: Ondes Electromagnetiques  et
  Biologie, A.J. Berteaud and B. Servantie, eds. Paris,
  France, pp. 325-329.

Albert, E.N., M.F. Sherif,  and  N.J. Papadopoulos.
  1981 a. Effect of  Non-ionizing Radiation on the
  Purkinje Cells of  the Uvula  in Squirrel Monkey
  Cerebellum.  Bioelectromagnetics,  2:241 -246.
Albert, E.N., M.F. Sherif,  N.J.  Papadopoulos, FJ.
  Slaby, and J.  Monahan. 1981b. Effect of Nonionizing
  Radiation on the  Purkinje Cells  of the Rat
  Cerebellum.  Bioelectromagnetics,  2:247-257.

Alfsen, A., and A.J. Berteaud, eds. 1976. Water and
  Biological Systems, International Colloquia of the
  National Center for Scientific Research, No. 246,
  Paris, France. 321 pp.
                                              R-1

-------
Allen, S.J., W.D. Hurt, J.H. Krupp, J.A. Ratliff, C.H.
  Durney, and C.C. Johnson. 1976. Measurement of
  Radiofrequency Power Absorption in Monkeys,
  Monkey Phantoms, and Human Phantoms Exposed
  to 10-50 MHz Fields. In: Biological Effects of
  Electromagnetic Waves,  Vol.  II, C.C. Johnson and
  M.L Shore, eds. HEW Publication (FDA) 77-8011,
  Rockville, Maryland, pp. 83-95.
Allis, J.W. 1975.  Irradiation of  Bovine Serum
  Albumin with a Crossed-Beam Exposure-Detection
  System. Ann. N.Y. Acad. Sci., 247:312-322.

Allis,  J.W., and  M.L.  Fromme.  1979.  Activity of
  Membrane-Bound Enzymes Exposed to Sinusoidally
  Modulated 2450-MHz Microwave Radiation. Radio
  Sci., 14(6S):85-91.
Allis,  J.W., M.L Fromme,  and D.E. Janes. 1976.
  Pseudosubstrate Binding to Ribonuclease During
  Exposure to Microwave Radiation at 1.70 and 2.45
  GHz. In: Biological Effects  of Electromagnetic
  Waves, Vol. I, C.C. Johnson and M.L. Shore, eds.
  HEW  Publication  (FDA) 77-8010,  Rockville,
  Maryland, pp. 366-376.

Allis, J.W., C.F.  Blackman,  M.L Fromme,  and S.G.
  Benane. 1977.  Measurement  of  Microwave
  Radiation Absorbed by  Biological Systems, 1,
  Analysis of Heating and Cooling Data. Radio Sci.,
  12(6S):1-8
Altman,  P.L., and D.S. Dittmer. 1972. Growth. In:
  Biological Handbook,  Fed. Am. Soc.  Exp.  Biol.,
  Washington, D.C. pp. 537-538.
American  Conference  of  Governmental  Industrial
  Hygienists. 1983. TLV's—Threshold Limit Values
  for Chemical Substances and Physical Agents in
  the Workroom Environment with Intended Changes
  for 1980. American Conference of Governmental
  Industrial Hygienists, Cincinnati, Ohio.
Ames, B.N.,  J.  McCann, and  E. Yamasaki. 1975.
  Methods for Detecting Carcinogens and Mutagens
  with the Sa/mo/7e//a/Mammalian-Microsome
  Mutagenicity Test. Mutat. Res., 31:347-364.
Anderstam, B., Y. Hamnerius, S. Hussain, and L.
  Ehrenberg.  1983. Studies of Possible Genetic
  Effects  in  Bacteria of High Frequency Electro-
  magnetic Fields. Hereditas, 98:11-32.

ANSI. 1974. An American National Standard, Safety
  Level of Electromagnetic Radiation with Respect
  to Personnel (C95.1-1974). American  National
  Standards Institute, New York, New York.

ANSI. 1982. American National Standard  Safety
  Levels with Respect to Human Exposure to Radio
  Frequency Electromagnetic Fields, 300 kHz-100
  GHz (ANSI C95.1-1982). American National
  Standards Institute, New York, New York.
Appleton,  B. 1973. Results of Clinical Surveys of
  Microwave Ocular Effects. HEW Publication (FDA)

                      R-2
  73-8031. U.S. Dept.  of  Health, Education, and
  Welfare, Rockville, Maryland. 13 pp.
Appleton, B., and G.C. McCrossan. 1972. Microwave
  Lens Effects in Humans. Arch. Ophthal., 88:259-
  262.
Appleton, B., S.E. Hirsch, and P.V.K. Brown. 1975.
  Investigation of Single Exposure Microwave Ocular
  Effects  at  3000  MHz. Ann. N.Y. Acad. Sci.,
  247:125-134.

Arber, S.L 1976. Effect of Microwaves on Resting
  Potential of Giant  Neurons  of Mullusk  Helix
  Pomatia. Electronnaya Obrabotka Materialov,
  6:78-79.

Aschoff, J. 1981. Thermal Conductance in Mammals
  and Birds:  Its Dependence on Body Size and
  Circadian Phase.  Comp. Biochem. Physiol., 69A:
  611-619.
Ashman, R.B., and AJ. Nahmias. 1978. Effect of
  Incubation Temperature on Mitogen Responses of
  Lymphocytes from Adult Peripheral Blood and from
  Cord Blood. Clin. Exp. Immunol., 33:319-326.

Asian, E.E. 1970. Electromagnetic Radiation Survey
  Meter. IEEE Trans. Instrum. Meas., IM-19:368-
  372.
Asian, E. 1976. A Low Frequency H-Field Radiation
  Monitor. In: Biological Effects of Electromagnetic
  Waves, Vol. II, C.C. Johnson and M.L. Shore, eds.
  HEW  Publication (FDA) 77-8011, Rockville,
  Maryland, pp. 229-238.

Asian,  E.  1979. The Maturing  of Electromagnetic
  Radiation Hazard Instruments. Microwave J.,
  22:83-90.
Athey, T.W.  1981. Comparison of  RF-lnduced
  Calcium Efflux From Chick Brain Tissue at Different
  Frequencies: Do the Scaled Power Density
  Windows Align? Bioelectromagnetics, 2:407-409.
Atkinson,  R.M. 1975.  Screening Medicines for
  Teratogenicity: Problems of Interpretation. In:
  Teratology: Trends and Applications, C.L. Berry and
  D.E.  Poswillo, eds. Springer-Verlag, New York,
  New York. pp. 136-146.

Austin, G.N., and S.M. Horvath. 1954. Production of
  Convulsions in Rats by High Frequency Electrical
  Currents. Am. J. Phys. Med., 33:141-149.

Barahski, S.  1971.  Effect of Chronic Microwave
  Irradiation on the Blood Forming System of Guinea
  Pigs and Rabbits. Aerospace Med., 42:1196-1199.

Barahski, S. 1972a. Effect of Microwaves on the
  Reactions of the White Blood Cells  System. Acta
  Physiol. Polon., 23:685-695.
Barahski, S. 1972b.  Histological and Histochemical
  Effect of Microwave  Irradiation on the Central

-------
  Nervous System of Rabbits and Guinea Pigs. Am. J.
  Phys. Med., 51(4):182-191.
Barahski, S., and P. Czerski. 1976. Biological Effects
  of Microwaves. Dowden, Hutchinson,  and Ross,
  Stroudsburg, Pennsylvania. 234 pp.
Barahski, S., and Z. Edelwejn. 1968. Studies on the
  Combined Effect of Microwaves and Some Drugs
  on Bioelectric Activity of Rabbit Central Nervous
  System. Acta Physiol. Polon., 19:31-41.
Barahski, S., and Z. Edelwejn. 1974. Pharmacologic
  Analysis  of Microwave Effects  on the  Central
  Nervous  System  in Experimental Animals. In:
  Biologic Effects and Health Hazards of Microwave
  Radiation, P. Czerski, K. Ostrowski, M.L. Shore, C.
  Silverman, M.J. Suess, and fi. Waldeskog, eds.
  Polish Medical Publishers, Warsaw, Poland, pp.
  119-127.
Barahski,  S., K. Ostrowski,  and  W.  Stodolnik-
  Baranska. 1972.  Functional and Morphological
  Studies of the Thyroid Gland in Animals Exposed
  to Microwave  Irradiation. Acta  Physiol. Polon.,
  23:1029-1039.
Barber,  D.E. 1962.  The Reaction of Luminous
  Bacteria to Microwave Radiation Exposures in the
  Frequency Range of 2608.7-3082.3  Me. IEEE
  Trans. Biomed. Electronics, BME-9:77-80.
Barber, P.W., O.P.  Gandhi, M.J. Hagmann, and I.
  Chatterjee. 1979. Electromagnetic Absorption in a
  Multi-Layered Model of Man. IEEE Trans. Biomed.
  Eng., BME-26(7):400-405.
Barnes, F.S., and C.J. Hu. 1977. Model for Some
  Nonthermal Effects of Radio and Microwave Fields
  on Biological Membranes. IEEE Trans. Microwave
  Theory Techniques, MTT-25:742-746.
Barren, C.I., and A.A. Baraff. 1958. Medical Con-
  siderations of Exposure to Microwaves  (Radar). J.
  Am. Med. Ass., 168:1194-1199.
Barren, C.I., A.A.  Love, and  A.A. Baraff.  1955.
  Physical  Evaluations of Personnel Exposed to
  Microwave Emanations. J. Aviat. Med., 26:442-
  452.

Bassen, H.  1977. Internal Dosimetry and External
  Microwave Field  Measurement Using  Miniature
  Electric Field. Probes. In: Symposium on Biological
  Effects  and Measurement of Radio Frequency/
  Microwaves, D.G. Hazzard, ed. HEW Publication
  (FDA) 77-8026, Rockville, Maryland, pp. 136-151.
Bassen,  H.,  M.  Swicord,  and J. Abita. 1975.  A
  Miniature  Broad-Band Electric Field Probe. Ann.
  N.Y. Acad. Sci., 247:481 -493.
Bassen, H., P. Herchenroeder, A. Cheung, and S.
  Neuder. 1977a.  Evaluation of an Implantable
  Electric-Field Probe Within Finite Simulated
  Tissues. Radio Sci., 12(6S): 15-25.
Bassen, H., W. Herman, andR. Hoss. 1977b. EM-Probe
  with Fiber Optic Telemetry System. Microwave J.,
  20:35-47.
Bassett, H.L, H.A.  Ecker, R.C. Johnson,  and A.P.
  Sheppard. 1971. New Techniques for Implementing
  Microwave Biological  Exposure Systems. IEEE
  Trans. Microwave Theory Techniques,  MTT-
  19(2): 197-204.
Bawin, S.M.,  and W.R. Adey. 1976. Sensitivity  of
  Calcium Binding  in Cerebral Tissue  to Weak
  Environmental Electric  Fields Oscillating at Low
  Frequency. Proc. Natl. Acad. Sci. USA, 73:1999-
  2003.
Bawin, S.M., and W.R. Adey. 1977. Calcium Binding
  in Cerebral Tissues. In: Symposium on Biological
  Effects and Measurement  of Radio  Frequency/
  Microwaves, D.G.  Hazzard, ed. HEW Publication
  (FDA) 77-8026, Rockville, Maryland, pp. 305-313.

Bawin, S.M.,  R.J. Gavalas-Medici, and W.R. Adey.
  1973. Effects of Modulated Very High Frequency
  Fields on Specific Brain Rhythms in Cats. Brain
  Res., 58:365-384.
Bawin, S.M., L.K. Kaczmarek, and W.R. Adey. 1975.
  Effects of Modulated VHF  Fields on  the Central
  Nervous System. Ann. N.Y. Acad. Sci., 247:74-81.
Bawin, S.M., W.R. Adey, and I.M. Sabbot. 1978. Ionic
  Factors in Release  of 46Ca2* From  Chicken
  Cerebral Tissue by Electromagnetic  Fields. Proc.
  Natl. Acad. Sci. USA, 75:6314-6318.
Belkhode, M.L., D.L Johnson, and A.M.  Muc. 1974a.
  Thermal  and Athermal Effects of  Microwave
  Radiation on the Activity of Glucose-6-Phosphate
  Dehydrogenase in Human Blood. Health Phys.,
  26:45-51.
Belkhode, M.L, A.M. Muc, and D. L Johnson. 1974b.
  Thermal and Athermal  Effects of  2.8  GHz
  Microwaves on Three Human Serum Enzymes. J.
  Microwave Power, 9:23-29.
Belsher, D.R. 1975. Development of Near-Field
  Electric Energy Density Meter Model EDM-2. HEW
  Publication  (NIOSH) 75-140, U.S. Department  of
  Health,  Education, and  Welfare, Public Health
  Service, Cincinnati, Ohio. 61 pp.
Ben-Hur, E.,  M.M.  Elkind, and B.V. Bronk,  1974.
  Thermally Enhanced Radio-response  of Cultured
  Chinese Hamster  Cells: Inhibition of Repair  of
  Sublethal Damage and Enhancement of Lethal
  Damage. Radiat. Res., 58:38-51.
Benzinger, T.H. 1969. Heat Regulation: Homeostasis
  of  Central  Temperature in Man. Physiol.  Rev.
  49:671-759.
Berman, E., and H.B. Carter. 1984. Decreased Body
  Weight in Fetal Rats After Irradiation with 2450-
  MHz (CW) Microwaves.  Health Phys., 46:537-542.
                                                                    R-3

-------
Berman,  E., J.B. Kinn, and  H.B.  Carter. 1978.
  Observations of Mouse Fetuses after Irradiation
  with 2.45 GHz Microwaves. Health Phys., 35:791-
  801.
Berman, E., H.B. Carter, and D. House. 1 980. Tests of
  Mutagenesis  and  Reproduction  in Male Rats
  Exposed to 2450-MHz (CW) Microwaves. Bioelec-
  tromagnetics, 1:65-76.
Berman,  E., H.B. Carter, and  D.  House. 1981.
  Observations of Rat Fetuses after Irradiation with
  2450-MHz (CW) Microwaves. J. Microwave Power,
Berman,  E.,  H.B.  Carter, and D.  House. 1982.
  Observations of Syrian Hamsters after Exposure to
  2450-MHz  Microwaves. J. Microwave Power,
  17(2): 107- 11 2.
Bermant, R.I., D.L Reeves, D.M. Levinson, and D.R.
  Justesen. 1979. Classical Conditioning of Micro-
  wave-Induced Hyperthermia in Rats. Radio  Sci.,
  14(6S):201-207.

Bingham, P.M., R.H. Baltz, LS. Ripley, and J.W.
  Drake. 1 976. Heat Mutagenesis in Bacteriophage
  T4: The Transversion Pathway. Proc. Natl. Acad.
  Sci. USA 73(11 ):41 59-41 63.
Bini, M.,  A.  Checcucci, A. Ignesti,  L. Millanta,  N.
  Rubino, C.  Camici, G.  Manao, and G. Ramponi.
  1 978. Analysis of the Effects of Microwave Energy
  on Enzymatic Activity of Lactate Dehydrogenase
  (LDH). J. Microwave Power, 13:95-99.

Birenbaum, L., G.M. Grosof,  S.W. Rosenthal, and
  M.M. Zaret. 1969a. Effect of Microwaves on the
  Eye. IEEE Trans. Biomed. Eng., BME-16:7-14.

Birenbaum,  L., I.T. Kaplan, W.  Metlay, S.W.
  Rosenthal,  H. Schmidt, and M.M. Zaret.  1969b.
  Effect of Microwaves on the  Rabbit  Eye.  J.
  Microwave Power, 4:232-243.
Birenbaum,  L., I.T. Kaplan, W.  Metlay, S.W.
  Rosenthal, and M.M. Zaret.  1975. Microwave and
  Infra-Red Effects on Heart Rate, Respiration  Rate
  and Subcutaneous Temperature of the Rabbit. J.
  Microwave Power, 10:3-18.
Blackman, C.F., and J.A. Black. 1 977. Measurement
  of Microwave Radiation Absorbed by Biological
  Systems, 2. Analysis by Dewar-Flask Calorimetry.
  Radio Sci., 12(6S):9-14.
Blackman, C.F., S.G. Benane, C.M. Weil, and J.S. Ali.
  1975. Effects of Nonionizing Electromagnetic
  Radiation on Single-Cell Biologic  Systems. Ann.
  N.Y. Acad. Sci., 247:352-366.

Blackman, C.F., M.C. Surles, and S.G. Benane. 1976.
  The Effect  of Microwave Exposure  on  Bacteria:
  Mutation  Induction.  In: Biological Effects  of
  Electromagnetic Waves, Vol. I, C.C. Johnson and
  M.L Shore, eds. HEW Publication (FDA) 77-8010,
  Rockville, Maryland, pp. 406-413.
Blackman, C.F., J.A. Elder, C.M. Weil, S.G. Benane,
  D.C. Eichinger, and D.E. House. 1979. Induction of
  Calcium-Ion  Efflux from  Brain  Tissue by Radio-
  Frequency Radiation: Effects of Modulation
  Frequency and Field Strength. Radio Sci., 14(6S):
  93-98.
Blackman, C.F., S.G. Benane, J.A. Elder, D.E. House,
  J.A. Lampe,  and J. M.  Faulk. 1980a.  Induction of
  Calcium-Ion  Efflux from Brain Tissue by Radiofre-
  quency Radiation: Effect of Sample Number and
  Modulation  Frequency  on the Power-Density
  Window. Bioelectromagnetics, 1:35-43.

Blackman, C.F., S.G.  Benane, W.T. Joines,  M.A.
  Hollis, and D.E. House. 1980b. Calcium-Ion Efflux
  from Brain Tissue: Power-Density Versus Internal
  Field-Intensity Dependencies at 50 MHz RF
  Radiation. Bioelectromagnetics,  1:277-283.
Blackman, C.F., W.T. Joines, and J.A. Elder. 1981.
  Calcium Ion  Efflux Induction in Brain Tissue by
  Radiofrequency Radiation. In: Biological Effects of
  Nonionizing  Radiation,  K.H.  Illinger, ed.  ACS
  Symposium Series, 157:299-314.
Blackman, C.F.,  S.G.  Benane, L.S.  Kinney,  W.T.
  Joines, and D.E. House. 1982. Effects of ELF Fields
  on Calcium-Ion Efflux From Brain Tissue In Vitro.
  Radiat. Res.,  92:510-520.
Blasberg, R.G. 1979. Problems of Quantifying Effects
  of Microwave Irradiation on  the  Blood-Brain
  Barrier. Radio Sci., 14(6S):335-344.
Blecha, F., R.A. Barry, and K.W. Kelley. 1982. Stress-
  Induced Alterations in  Delayed-Type Hypersensi-
  tivity to SRBC and Contact Sensitivity to DNFB in
  Mice. Proc. Soc. Exp. Biol. Med., 169:239-246.

Blevins, R.D., R.C. Crenshaw, A.E. Hougland, andC.E.
  Clark.  1980.  The Effects of Microwave Radiation
  and Heat on Specific Mutants of  Salmonella
  typhimurium LT2. Radiat. Res., 82:511-517.

Bligh, J.  1973. Temperature Regulation in Mammals
  and Other Vertebrates. In: Frontiers of Biology, Vol.
  5. North American Holland Research Monograph.
Bligh, J., and K.G.Johnson. 1973. Glossary of Terms
  for Thermal Physiology. J. Appl. Physiol., 35:941 -
  961.

Boggs, R.F., A.P. Sheppard, and A.J. Clark. 1972.
  Effects of 2450  MHz Microwave Radiation on
  Human  Blood Coagulation  Processes.  Health
  Phys., 22:217-224.
Bowler, K. 1972. The Effect of Repeated Applications
  of Heat on Spermatogenesis in the Rat: A
  Histological Study. J.  Reprod.  Fertility, 28:325-
  333.
                       R-4

-------
Bowman, R.R. 1970. Quantifying Hazardous Electro-
  magnetic Fields:  Practical Considerations. NBS
  Technical Note 389, U.S. Department of Commerce,
  National Bureau of Standards, Boulder, Colorado.
  15pp.
Bowman, R.R.  1976. A  Probe  for  Measuring
  Temperature in Radio-Frequency-Heated Material.
  IEEE Trans. Microwave Theory Techniques, MTT-
  24:43-45.
Bradley, L.M.,  and R.I. Michell. 1981. Differential
  Effects of Glucocorticoids on the Functions of
  Helper and Suppressor T Lymphocytes. Proc. Natl.
  Acad. Sci. USA, 78:3155-3159.

Brodeur, P. 1977. The Zapping of America. W.W.
  Norton and Co., New York, New York. 343 pp.

Brown, J.L 1975. The Evolution of Behavior. W.W.
  Norton and Co., New York, New York, 761  pp.

Bullard, R.W.  1971. Temperature Regulation.  In:
  Physiology, 3rd ed., E.E. Seklurt, ed. Little, Brown &
  Co., Boston, Massachusetts, pp. 651-667.

Burton, A.C. 1939. The Range and Variability of the
  Blood  Flow in the  Human Fingers and  the
  Vasomotor Regulation of Body Temperature. Am. J.
  Physiol., 127:437-453.

Buss,  M.E., and  S.A.  Henderson.  1971. Induced
  Bivalent Interlocking  and  the Course of  Meiotic
  Chromosome Synapsis. Nature  New Biol., 234:
  243-246.

Cabanac, M., and B. Dib. 1983. Behavioral Responses
  to Hypothalamic Cooling and Heating in the Rat.
  Brain Res., 264:79-87.

Cahill, D.F. and  J.A.  Elder,  eds.  1983.  Biological
  Effects of Radiofrequency Radiation. External
  review draft. EPA-600/8-83-026A. Available from
  National Technical Information Service (NTIS No.
  PB83 161550), Springfield, Virginia  22161. 582
  PP
Cain, C.A.,  and W.J. Rissmann. 1978.  Mammalian
  Auditory Responses to 3.0  GHz Microwave Pulses.
  IEEE Trans. Biomed. Eng.,  BME-25:288-293.
Cairnie, A.B., and K.E. Leach.  1980. Quantitative
  Studies  of Cytological Damage in Mouse Testis
  Produced by Exposure to  Heat. Can.  J.  Genet.
  Cytol., 22:93-102.

Cairnie, A.B., D.A. Hill, and H.M. Assenheim. 1980a.
  Dosimetry for  a  Study of Effects of  2.45-GHz
  Microwaves on Mouse Testes.  Bioelectromagnetics,
  1:325-336.

Cairnie,  A.B.,  L.F.  Prud'homme-Lalonde, R.K.
  Harding, and M. Zuker. 1980b. The Measurement
  of Rectal and  Testis Temperature in Conscious
  Mice, with  Observations on  the Effect of Direct
  Heating. Phys.  Med. Biol.,  25(2):317-322.
Carlisle, H.J., and D.L Ingram. 1973. The Effects of
  Heating and Cooling the Spinal Cord and Hypo-
  thalamus on Thermoregulatory Behavior in the Pig.
  J. Physiol., 231:353-364.
Carlson, N.R. 1980. Physiology of Behavior, 2nd ed.
  Allyn and Bacon, Inc., Boston, Massachusetts. 748
  PP-
Carpenter, R.L.  1979. Ocular  Effects of Microwave
  Radiation. Bull. N.Y. Acad. Med., 55:1048-1057.
Carpenter, R.L, and C.A. Van Ummersen. 1968. The
  Action of Microwave Radiation on  the Eye.  J.
  Microwave Power,  3:3-19.
Carpenter, R.L., D.K. Biddle, and C.A. Van Ummersen.
  1960a. Biological Effects of Microwave Radiation,
  with Particular Reference to the Eye. In: Proc. Third
  International Conference on Medical Electronics,
  London. Medical Electronics, Part III. pp. 401 -408.
Carpenter, R.L., D.K. Biddle, and C.A. Van Ummersen.
  1960b. Opacities in the Lens of the Eye Experiment-
  ally Induced by Exposure to Microwave Radiation.
  IRE Trans. Med. Electronics, 7:152-157.
Carpenter, R.L, G.J. Hagan, and G.L Donovan. 1977.
  Are  Microwave Cataracts Thermally Caused? In:
  Symposium on Biological Effects and Measurement
  of Radio Frequency/Microwaves, D.G. Hazzard, ed.
  HEW Publication (FDA) 77-8026, Rockville,
  Maryland, pp. 352-379.

Carroll, D.R., D.M. Levinson, D.R. Justesen, andR.L.
  Clarke, 1980.  Failure of  Rats to Escape from a
  Potentially Lethal  Microwave Field. Bioelectro-
  magnetics, 1:101-115.
Carslaw, H.S., and J.C. Jaeger. 1959. Conduction of
  Heat in Solids. Clarendon Press, Oxford,  England.
  pp. 230-231.
Catania, A.C., ed. 1968. Contemporary Research in
  Operant Behavior. Scott, Foresman, Glenview,
  Illinois. 358 pp.
Catravas, G.N. 1976.  Styrofoam Cages for Rats used
  in Microwave Research: Coating with  Quinine.
  Health Phys.,  31:68-69.
Catravas, G.N., J.B. Katz, J. Takenaja, and J.R. Abbott.
  1976. Biochemical Changes in the Brain of Rats
  Exposed to Microwaves of Low Power Density. J.
  Microwave Power,  11:147-148.
Chamness,  A.F., H.R. Scholes, S.W. Sexauer, and
  J.W. Frazer. 1976. Metal Ion Content of Specific
  Areas  of  the  Rat  Brain after  1600 MHz Radio
  Frequency Irradiation. J.  Microwave Power,
  11:333-338.
Chang, B.K., AT. Huang, and W.T. Joines.  1981.
  Microwave Treatment of Intracerebral L1210
  Leukemia:  Schedule-Dependent Partial Reversal of
  the Effects of  Methotrexate.  Bioelectromagnetics,
  2:77-80.
                                                                      R-S

-------
Chappuis, P., P. Pittet, and E. Jequier.  1976. Heat
  Storage Regulation in Exercise During Thermal
  Transients. J. Appl. Physiol., 40:384-392.
Chatterjee, I., M.J. Hagmann, and O.P. Gandhi. 1980.
  Electromagnetic Energy Deposition in  an Inhomo-
  geneous Block Model of Man for  Near-Field
  Irradiation Conditions. IEEE Trans.  Microwave
  Theory Techniques, MTT-28:1452-1459.
Chen, K.C., and C.J. Lin. 1978. A System for Studying
  Effects  of Microwaves on Cells in  Culture. J.
  Microwave Power, 13:251-256.
Chen, K.M., and B.S. Guru. 1977. Induced EM Fields
  Inside Human Bodies Irradiated by EM Waves up to
  500 MHz. J. Microwave Power, 12(2): 173-183.

Chernovetz, M.E., D.R. Justesen, N.W. King, and J.E.
  Wagner. 1975. Teratology, Survival, and Reversal
  Learning after Fetal Irradiation of Mice by 2450-
  MHz Microwave  Energy.  J. Microwave Power,
  10(4):391-409.
Chernovetz, M.E., D.R. Justesen, and A.F. Oke. 1977.
  A Teratological Study  of the Rat: Microwave and
  Infrared Radiations Compared. Radio Sci., 12(6S):
  191-197.
Chernovetz, M.E., D.R. Justesen, and D.M. Levinson,
  1979. Acceleration and  Deceleration of Fetal
  Growth of Rats by 2450-MHz Microwave Radiation.
  In: Electromagnetic Fields in Biological Systems,
  S.S. Stuchly, ed. Ottawa, Canada, pp.  175-193.

Chiabrera, A., M. Hinsenkamp, A.A. Pilla,  J. Ryaby, D.
  Ponta, A. Belmont, F. Beltrame, M. Grattarola, and
  C. Nicolini. 1979.  Cytofluorometry  of Electromag-
  netically Controlled Cell Dedifferentiation. J.
  Histochem. Cytochem., 27:375-381.
Chou, C.K.,  and  R.  Galambos. 1979.  Middle-Ear
  Structures Contribute Little to Auditory Perception
  of Microwaves. J. Microwave Power, 14:321-326.

Chou, C.K., and A.W.  Guy. 1979. Microwave-Induced
  Auditory Responses in Guinea Pigs: Relationship of
  Threshold  and Microwave-Pulse Duration. Radio
  Sci., 14(6S): 193-197.
Chou, C.K., R. Galambos, A.W. Guy, and  R.H.  Lovely.
  1975.  Cochlear   Microphonics Generated by
  Microwave Pulses. J. Microwave Power, 10:361-
  367.
Chou, C.K.,  A.W. Guy,  and R. Galambos  1976.
  Microwave-Induced Auditory Response: Cochlear
  Microphonics.  In: Biological Effects  of Electro-
  magentic Waves,  Vol.  I, C.C. Johnson and M.L.
  Shore,  eds.  HEW Publication  (FDA) 77-8010,
  Rockville, Maryland, pp. 89-103.

Chou, C.K.,  A.W. Guy,  and R. Galambos.  1977.
  Characteristics of Microwave-Induced Cochlear
  Microphonics. Radio Sci., 12(6):221-227.
Chou, C.K., A.W. Guy, K.R. Foster, R. Galambos, and
  D.R. Justesen. 1980a. Holographic Assessment of
  Microwave Hearing. Science, 209:1143-1144.
Chou, C.K.,  L.F. Han,  and  A.W.  Guy. 1980b.
  Microwave Radiation and  Heart-Beat Rate of
  Rabbits. J. Microwave Power, 15:87-93.
Chou, C.K., A.W. Guy,  and R. Galambos.  1982.
  Auditory Perception  of Radio-Frequency Electro-
  magnetic Fields. J. Acoust. Soc. Am., 71(6):1321-
  1334.

Christensen,  D.A. 1977. A New Nonperturbing
  Temperature Probe Using Semiconductor Band
  Edge Shift. J. Bioeng., 1:541-545.
Christman, C.L., H.S. Ho, and S. Yarrow. 1974. A
  Microwave Dosimetry System for Measured
  Sampled Integral-Dose Rate. IEEE  Trans.  Micro-
  wave Theory Techniques, MTT-22(12):1267-1272.

Clapman,  R.M., and C.A. Cain.  1975. Absence of
  Heart Rate Effects in Isolated Frog Heart Irradiated
  with Pulsed Modulated Microwave Energy. J.
  Microwave Power, 10:411-419.
Cleary, S.F. 1980. Microwave  Cataractogenesis.
  Proc. IEEE, 68:49-55.
     «
Cleary, S.F., and B.S. Pasternack. 1966. Lenticular
  Changes in  Microwave  Workers: A Statistical
  Study. Arch. Environ. Health, 12:23-29.
Cleary, S.F., B.S. Pasternack, and G.W. Beebe. 1965.
  Cataract Incidence in Radar Workers.  Arch.
  Environ. Health, 11:179-182.
Cleary, S.F., and R.T. Wangemann. 1976. Effect of
  Microwave Radiation on  Pentobarbital-lnduced
  Sleeping Time.  In: Biological  Effects of Electro-
  magnetic Waves, Vol.  I, C.C. Johnson  and M.L
  Shore, eds. HEW Publication (FDA) 77-8010,
  Rockville, Maryland, pp. 311-322.

Cogan, D.G., S.J. Fricker, M. Lubin, D.D. Donaldson,
  and H.  Hardy. 1958. Cataracts and Ultra-High-
  Frequency Radiation. A.M.A.  Arch. Ind. Health,
  18:299-302.
Cohen, B.H., A.M. Lilienfeld, S. Kramer, and  LC.
  Hyman. 1977. Parental Factors in Down's
  Syndrome: Results of the Second Baltimore Case-
  Control Study.  In: Population  Cytogenetics—
  Studies  in Humans, E.B. Hook and I.H. Porter, eds.
.  Academic Press, New  York, New York. pp. 301-
  352.
Constant, P.C., Jr.  1967. Hearing EM Waves. Digest
  of the Seventh International  Conference  on
  Medical and Biological Engineering, B. Jacobson,
  ed. Department of Medical Engineering, Karolinska
  Institute, Stockholm, Sweden, p. 349.

Cook, H.F. 1952. The Pain Threshold for Microwave
  and Infra-Red Radiations. J. Physiol., 118:1-11.
                      R-6

-------
Cooper, J.R., F.E. Bloom, and R.H. Roth. 1982. The
  Biochemical Basis of Neuropharmacology, 4th ed.
  Oxford University Press, New York, New York. 367
  pp.
Corelli, J.C., R. J. Gutmann, S. Kohazi, and J.  Levy.
  1977. Effects of 2.6-4.0 GHz Microwave Radiation
  on E. co//B. J. Microwave Power, 12:141-144.

Crawford, M.L  1974. Generation of Standard FM
  Fields Using TEM Transmission Cells. IEEE Trans.
  Electromagnetic Compatibility, EMC-16:189-95.

Crosbie,  R.J., J.D. Hardy, and E. Fessender. 1963.
  Electrical Analog Simulation  of  Temperature
  Regulation in  Man. In: Temperature, Its Measure-
  ment and Control in Science and Industry, Part III,
  J.H. Hardy, ed. Reinhold Publ., New York, New York.
  Ch. 55, p. 627.
Crosby,  E.G., T.  Humphrey, and E.W. Lauer. 1962.
  Correlative Anatomy of the Nervous System. The
  Macmillan Company, New York, New York.

Czerski, P.  1975. Microwave Effects on  the Blood-
  Form ing System with Particular Reference to the
  Lymphocyte. Ann. N.Y. Acad. Sci., 247:232-242.

Czerski, P., M. Sierkierzynski, and A. Gidynski. 1974.
  Health  Surveillance of Personnel Occupationally
  Exposed to Microwaves. I. Theoretical Considera-
  tions and Practical  Aspects. Aerospace Med.,
  45:1137-1142.
Daily, L,  Jr., K.G. Wakim, J.F. Herrick, E.M. Parkhill,
  and W.L. Benedict. 1950a. The Effects of Microwave
  Diathermy on the Eye of the Rabbit.  Am. J.
  Ophthalmol., 35:1001-1017.
Daily, L,  Jr., K.G. Wakim, J.F. Herrick, E.M. Parkhill,
  and W.L. Benedict. 1950b. The Effects of Microwave
  Diathermy on the Eye.  Am. J.  Ophthalmol.,
  33:1241-1254.
D'Andrea, J.A., O.P. Gandhi, and R.P. Kesner, 1976.
  Behavioral Effects of Resonant Electromagnetic
  Power Absorption in Rats.  In: Biological Effects of
  Electromagnetic Waves, Vol. I, C.C. Johnson and
  M.L. Shore, eds. HEW Publication (FDA) 77-8010,
  Rockville, Maryland, pp. 257-273.
D'Andrea, J.A., O.P. Gandhi, and J.L. Lords. 1977.
  Behavioral and Thermal  Effects of Microwave
  Radiation at  Resonant and Nonresonant  Wave-
  lengths. Radio Sci., 12:251-256.
D'Andrea, J.A., O.P. Gandhi, J.L Lords, C.H. Durney,
  C.C. Johnson, and L Astle. 1979. Physiological and
  Behavioral Effects of Chronic Exposure to 2450-
  MHz  Microwaves.  J. Microwave Power, 14:351-
  362.
D'Andrea, J.A., O.P. Gandhi, J.L. Lords, C.H. Durney,
  L.  Astle,  L.J.  Stensaas, and A.A. Schoenberg.
  1980   Physiological and  Behavioral  Effects of
  Prolonged Exposure to 915 MHz Microwaves. J.
  Microwave Power, 15(2):123-135.
Dardalhon, M., D. Averbeck, and AJ. Berteaud. 1979.
  Determination of Thermal Equilavent of Millimeter
  Microwaves in Living Cells. J. Microwave Power,
  14:307-312.
Dardalhon, M., D. Averbeck, and A.J. Berteaud. 1980.
  Action des Ondes Centimetriques Seules ou
  Combinees avec les Rayons Ultra Violets sur les
  Cellules  Eucaryotiques. In: URSI International
  Symposium Proceedings, Ondes Electromagne-
  tiques et Biologie, AJ. Berteaud and B. Servantie,
  eds. Paris, France, pp. 17-24.
Dardalhon, M., D. Averbeck, and AJ. Berteaud. 1981.
  Studies on Possible Genetic Effects of Microwaves
  in Procaryotic  and  Eucaryotic  Cells. Radiat.
  Environ. Biophys. 20:37-51.
Davidson,  J.A.,  P.A.  Kondra,  and M.A.K.  Hamid.
  1976. Effects of Microwave Radiation on Eggs,
  Embryos and Chickens. Can. J. Anim. Sci., 56:709-
  713.
Deficis, A., J.C. Dumas, S. Laurens, and G. Plurien.
  1979. Microwave Irradiation and Lipid Metabolism
  injvlice. Radio Sci., 14(6S):99-101.
Deichmann,  W.B., E. Bernal, F.  Stephens, and K.
  Landeen. 1963. Effects on Dogs of Chronic
  Exposure to Microwave Radiation. J. Occupational
  Med., 5:418-425.
Deichmann,  W.B., J. Miale, and K. Landeen. 1964.
  Effect of Microwave Radiation on the Hemopoietic
  System of the Rat. Toxicol. Appl. Pharmacol., 6:71 -
  77.
Delgado, J.M.R., J. Leal, J.L. Monteagudo, and M.
  Garcia-Gracia.  1982. Embryological Changes
  Induced  by Weak,  Extremely  Low Frequency
  Electromagnetic Fields. J. Anat., 134:533-551.

de  Lorge, J.O.  1976. The  Effects of Microwave
  Radiation on Behavior and Temperature in Rhesus
  Monkeys. In: Biological Effects of Electromagnetic
  Waves, Vol. I, C.C. Johnson and M.L. Shore, eds.
  HEW Publication  (FDA) 77-8010,  Rockville,
  Maryland, pp. 158-174.
de  Lorge,  J. 1979. Disruption of Behavior in
  Mammals of  Three Different  Sizes Exposed to
  Microwaves: Extrapolation to Larger Mammals. In:
  Electromagnetic Fields in Biological Systems, S.S.
  Stuchly, ed. Ottawa, Canada, pp. 215-228.
de  Lorge,  J. 1979.  Operant Behavior and Rectal
  Temperature of Squirrel Monkeys During 2.45-
  GHz Microwave Irradiation. Radio Sci., 14(6S):217-
  225.

de  Lorge,  J.O.,  and  C.S. Ezell.  1980. Observing-
  Responses of Rats Exposed to 1.28- and 5.62-GHz
  Microwaves. Bioelectromagnetics, 1:183-198.
                                                                     R-7

-------
Dietzel, F. 1975. Effects of Electromagnetic Radiation
  on Implantation and Intrauterine Development of
  the Rat. Ann. N.Y. Acad. Sci., 247:367-376.

Dietzel, F., and W. Kern. 1970. Abortion Following
  Ultra-Short-Wave Hyperthermia: Animal Experi-
  ments. Arch. Gynak., 209:237-255.
Dietzel, F., W. Kern, and R. Steckenmesser. 1972.
  Deformity and Intrauterine Death after Short-Wave
  Therapy in  Early Pregnancy in Experimental
  Animals. Munch. Med. Wchschr., 114:228-230.

Diffrient,  IM., A.R. Tilley, and J.C. Bardagjy. 1974.
  Humanscale 1/2/3. The MIT Press, Boston,
  Massachusetts.
Dill, D.B., E.F. Adolph, and C.G. Wilber, eds. 1964.
  Adaptation to the Environment. In: Handbook of
  Physiology,  Williams and Wilkins Co., Baltimore,
  Maryland.
Dixey, R., and G. Rein.  1982.  3H-noradrenaline
  Release Potentiated in a Clonal Nerve Cell Line by
  Low-Intensity Pulsed Magnetic Fields. Nature,
  296:253-256.
Djordjevic, Z., and A.  Kolak. 1973. Changes in  the
  Peripheral Blood of the Rat Exposed to Microwave
  Radiation (2400 MHz) in Conditions of Chronic
  Exposure. Aerospace Med., 44:1051-1054.

Djordjevic, Z., N. Lazarevic, and V. Djokovic. 1977.
  Studies on the Hematologic Effects of Long-Term,
  Low-Dose  Microwave Exposure. Aviat.  Space
  Environ. Med., 48:516-518.
Djordjevic, Z., A. Kolak, M. Stojkovic, N. Rankovic, and
  P. Ristic. 1979. A Study of the Health Status of
  Radar Workers. Aviat.  Space Environ. Med.,
  50:396-398.
Drost-Hansen,  W., and J.S. Clegg, eds. 1979. Cell-
  Associated Water, Academic Press, New York, New
  York. 440 pp.
Dumansky, J.D., and  M.G. Shandala. 1974. The
  Biologic Action  and Hygienic Significance of
  Electromagnetic Fields of Superhigh and Ultrahigh
  Frequencies  in Densely Populated Areas.  In:
  Biologic Effects and Health Hazards of Microwave
  Radiation. P. Czerski, K. Ostrowski, M.L Shore, C.
  Silverman, M.J. Suess, and B. Waldeskog, eds.
  Polish Medical  Publishers,  Warsaw, Poland,  pp.
  289-293.
Durney,  C.H.  1980. Electromagnetic Dosimetry for
  Models of Humans  and Animals: A Review of
  Theoretical and Numerical Techniques. Proc. IEEE,
  68(1):33-40.
Durney,  C.H.,  C.C.  Johnson,  P.W.  Barber,  H.
  Massoudi,  M.F. Iskander, J.L.  Lords, D.K. Ryser,
  S.J. Allen,  and J.C. Mitchell.  1978. Radio-
  frequency Radiation Dosimetry Handbook, 2nd ed.

                      R-8
  Report SAM-TR-78-22, USAF School of Aerospace
  Medicine, Brooks Air Force Base, Texas. 141 pp.
Durney, C.H., M.F. Iskander, H. Massoudi, and C.C.
  Johnson. 1979. An Empirical Formula for Broad-
  Band SAR Calculations  of Prolate  Spheroidal
  Models  of Humans and  Animals. IEEE Trans.
  Microwave Theory Techniques. MTT-27{8):758-
  763.
Durney, C.H., M. F. Iskander, H. Massoudi, S.J. Allen,
  and J.C. Mitchell. 1980. Radiofrequency Radiation
  Dosimetry Handbook, 3rd ed. Report SAM-TR-80-
  32,  USAF  School of Aerospace Medicine, Brooks
  Air Force Base, Texas.  136 pp.

Dutta, S.K., W.H. Nelson, C.F. Blackman, and D.J.
  Brusick. 1979a. Lack of Microbial  Genetic
  Response  to 2.45-GHz CW and 8.5  to 9.6-GHz
  Pulsed Microwaves. J. Microwave Power, 14:275-
  280.
Dutta, S.K., M.A.  Hossain, H.S. Ho, and C.F.
  Blackman.  1979b. Effects of 8.6-GHz  Pulsed
  Electromagnetic Radiation on an Escherichia coli
  Repair-Deficient Mutant. In: Electromagnetic
  Fields in Biological  Systems, S.S  Stuchly, ed.
  Edmonton, Canada, pp. 76-95.
Dutta, S.K., W.H. Nelson, C.F. Blackman, and D.J.
  Brusick. 1980. Cellular Effects in Microbial Tester
  Strains  Caused by Exposure to Microwaves or
  Elevated  Temperatures.  J.  Environ.  Pathol.
  Toxicol., 3:195-206.
Dutta,  S.K.  A. Subramoniam,  B. Ghosh,  and R.
  Parshad. 1984.  Microwave  Radiation-Induced
  Calcium Ion Efflux From Human Neuroblastoma
  Cells in Culture. Bioelectromagnetics, 5:71-78.
Dwyer,  M.J., and  D.B.  Leeper, 1978. A Current
  Literature Report on the Carcinogenic Properties of
  Ionizing  and Non-Ionizing Radiation. II. Microwave
  and Radiofrequency Radiation. DHEW Publication
  (NIOSH) No. 78-134, Cincinnati, Ohio. 28 pp.
Edelwejn,  Z.  1968.  Attempt at  Evaluation of the
  Functional State of Brain Synapses in Rabbits
  Exposed Chronically to the Action of Microwaves.
  Acta Physiol. Polon., 19:791-799.
Edwards, G.S., M.L Swicord, and C.C. Davis. 1983.
  Microwave Absorption Characteristics of Highly
  Purified  £. coli DNA. Abstract #A-5,  5th Annual
..Scientific Session, The Bioelectromagnetics
  Society, 12-17 July, Boulder, Colorado (available
  from The Bioelectromagnetics Society, One Bank
  Street Suite 307, Gaithersburg, Maryland 20878).
  p. 3.
Edwards, M.J. 1974. The Effects of Hyperthermia on
  Pregnancy and Prenatal Development. In: Experi-
  mental Embryology and Teratology,  Vol. 1, D.H.M.
  Woollam and G.M. Morriss, eds. Paul Elek, London,
  England, pp. 90-133.

-------
Edwards,  W.P.,  and H.S. Ho.  1975. RF  Cavity
  Irradiation  Dosimetry.  IEEE Trans. Microwave
  Theory Techniques, MTT-23(3):311-313.
Ehrenberg,  L, B. Anderstam, S. Hussain, and Y.
  Hamnerius. 1983. Statistical Aspects of the Design
  of Biological Tests for the  Detection  of  Low
  Genotoxic Activity. Hereditas, 98:34-41.
Elder, J.A.,  and J.S. AM. 1975. The Effect of Micro-
  waves (2450 MHz) on  Isolated  Rat Liver  Mito-
  chondria. Ann.  N.Y. Acad. Sci.,  247:251-262.
Elder, J.A.,  J.S. AM, M.D. Long, and G.E. Anderson,
  1976. A  Coaxial Air  Line  Microwave Exposure
  System:  Respiratory Activity of Mitochondria
  Irradiated  at 2-4  GHz.  In:  Biological Effects of
  Electromagnetic Waves, Vol. I, C.C. Johnson and
  M.L Shore, eds. HEW Publication (FDA) 77-8010,
  Rockville, Maryland, pp.  352-365.
Elizondo, R.S. 1973.  Local Control of Eccrine Sweat
  Gland Function. Fed. Proc. 32:1583-1587.
Ellis, F.P., H.M. Ferres, A.R. Lind, and P.S.B. Newling.
  1960. The Upper Limits of Tolerance of Environ-
  mental Stress.  In: Physiological Responses to Hot
  Environments. Spec. Report, Series No. 298. Med.
  Res. Council, London, pp. 158-179.
Ely, T.S., and D.E. Goldman. 1956. Heat Exchange
  Characteristics of Animals Exposed to  10-cm
  Microwaves. IRE Trans-Med. Electronics, February,
  pp. 38-43.
Ely, T.S.,  D.E. Goldman,  and J.Z. Hearon. 1964.
  Heating  Characteristics of Laboratory Animals
  Exposed  to Ten-Centimeter Microwaves. IEEE
  Trans. Biomed. Eng., 11:123-137.
Emery, A.F., R.E. Short, A.W. Guy, and K.K. Kraning.
  1976. The Numerical Thermal'Simulation of the
  Human  Body  When Undergoing  Exercise or
  Nonionizing Electromagnetic  Irradiation. Trans.
  ASME, J. Heat Transfer, 98:284-291.
Epidemiology Work Group. 1981. Guidelines for
  Documentation  of Epidemiologic Studies. Am. J.
  Epidemiol., 114:609-613.
Fahim, M.S., Z.  Fahim, R. Der,  D.G. Hall,  and J.
  Harman.  1975.  Heat in Male Contraception (Hot
  Water 60°C, Infrared, Microwave, and Ultrasound).
  Contraception,  11(5):549-562.
Federal Register.  1983a. Health Assessment
  Document on the Biological Effects, of Radiofre-
  quency Radiation. Fed. Reg., 48 (No. 141, July 21,
  1983):33345.
Federal  Register.  1983b.  Science  Advisory Board,
  Subcommittee on  the Biological Effects of
  Radiofrequency Radiation; Open Meeting. Fed.
  Reg., 48 (No. 171, September 1):39688.
Federal  Register.  1984.  Science Advisory Board,
  Biological Effects of Radiofrequency Radiation
  Subcommittee; Open Meeting. Fed. Reg., 49 (No. 3,
  January 5):662-663.
Fermi, E., J.R. Pasta, and S. Ulam. 1965. Studies of
  Non Linear Problems. In: Collected Works of Enrico
  Fermi, Vol. II. University of Chicago Press, Chicago,
  Illinois, pp. 978-988.
Ferri, E.S., and G.J. Hagan. 1976. Chronic Low-Level
  Exposure of Rabbits to Microwaves. In: Biological
  Effects of Electromagnetic  Waves, Vol.  I, C.C.
  Johnson  and M.L  Shore, eds.  HEW Publication
  (FDA) 77-8010,  Rockville, Maryland, pp. 129-142.

Flynn, R.J.  1968.  Exencephalia:  Its Occurrence in
  Untreated Mice. Science, 160:898-899.
Folk,  G.E.,  Jr. 1974. Textbook of Environmental
  Physiology, 2nd  ed. Lea and Febiger, Philadelphia,
  Pennsylvania. 466 pp.
Follenius, M., G. Brandenberger, S. Oyono, and V.
  Candas. 1982.  Cortisol  as a Sensitive Index of
  Heat-Intolerance. Physiol. Behav., 29:509-513.

Foster, K.R., and E.D. Finch. 1974. Microwave
  Hearing:  Evidence for Thermoacoustic Auditory
  Stimulation by Pulsed Microwaves. Science,
  1,85:256-258.
Frankel, J.M. 1959. Effects of Restraint on  Rats
  Exposed to High Temperature. J. Appl. Physiol.,
  14:997-999.
Frey, A.H. 1961. Auditory System Response to Radio
  Frequency Energy. Aerospace Med., 32:1140-
  1142.
Frey, A.H. 1962. Human Auditory System Response
  to  Modulated Electromagnetic Energy. J. Appl.
  Physiol., 17:689-692.
Frey, A.H. 1963. Some Effects on Human Subjects of
  Ultra-High-Frequency Radiation. Am. J. Med.
  Electron., 2:28-31.
Frey,  A.H.  1967.  Brain Stem Evoked Responses
  Associated with  Low-Intensity Pulsed UHF Energy.
  J. Appl. Physiol., 23:984-988.
Frey, A.W., and  E.  Coren. 1979. Holographic
  Assessment of Hypothesized Microwave Hearing
  Mechanism. Science, 206:232-234.

Frey, A.H., and S.R. Feld. 1975. Avoidance by Rats of
  Illumination with Low Power Nonionizing Electro-
  magnetic Energy. J. Comp. Physiol.  Psychol.,
  89:183-188.
Frey, A.H., and  R.  Messenger. 1973. Human
  Perception of Illumination with Pulsed Ultrahigh-
  Frequency Electromagnetic Energy.  Science,
  181:356-358.
Frey, A.H., and E. Seifert. 1968. Pulse Modulated UHF
  Illumination of the Heart Associated with Change
  in  Heart Rate. Life Sci., 7:505-512.
                                                                     R-9

-------
Frey,  A.M., S.R.  Feld, and B. Frey. 1975. Neural
  Function and Behavior: Defining the Relationship.
  Ann. N.Y. Acad. Sci., 247:433-439.
Friedman, H.L.  1981.  Are Chronic Exposure to
  Microwaves and  Polycythemia Associated? New
  England J. Med,  304:357-358.
Friedman, H., P.O. Becker, andC.H. Bachman. 1967.
  Effect of  Magnetic  Fields on Reaction Time
  Performance. Nature,  213:949-950.
Froehlich, J.P. 1969. Information Transmittal and
  Communicating  Systems.  Holt,  Rinehart,  and
  Winston, New York, New York. 274 pp.
Frohlich,  H. 1968. Long  Range Coherence  and
  Energy Storage in Biological Systems. Int. J. Quant.
  Chem., 2:641-649.
Frohlich,  H. 1975. The Extraordinary  Dielectric
  Properties of Biological Materials and the Action of
  Enzymes. Proc. Natl.  Acad.  Sci.  USA, 72:4211-
  4215.

Gage, M.I. 1979a. Behavior in Rats after Exposures to
  Various Power Densities of 2450 MHz Microwaves.
  Neurobehav. Toxicol., 1:137-143.
Gage, M.I. 1979b. Microwave Irradiation and
  Ambient Temperature  Interact to  Alter  Rat
  Behavior Following  Overnight  Exposure. J.
  Microwave Power, 14:389-398.
Gage, M.I.,  E.  Berman, and J.B. Kinn. 1979.
  Videotape Observation of Rats and Mice During an
  Exposure to 2450 MHz Microwave Radiation. Radio
  Sci., 14(6S):227-232.
Gale, C.C. 1973.  Neuroendocrine Aspects of
  Thermoregulation. Annu. Rev. Physiol., 35:391-
  430.
Gale,  C.C., M.  Mathews,  and J.  Young. 1970.
  Behavioral Thermoregulatory  Responses to
  Hypothalamic Cooling and Warming in Baboons.
  Physiol. Behav., 5:1-6.
Gandhi, O.P.  1980. State  of  the  Knowledge for
  Electromagnetic Absorbed Dose in  Man and
  Animals. Proc. IEEE, 68(1):24-32.
Gandhi, O.P., E.L Hunt, and J.A. D'Andrea. 1977.
  Deposition of Electromagnetic Energy in Animals
  and in Models of Man with and without Grounding
  and Reflector Effects. Radio Sci., 12(6S):39-47.
Gandhi, O.P.,  M.J.  Hagmann,  and  J.A. D'Andrea.
  1979. Part-Body and Multi-Body Effects on Absorp-
  tion of Radio-Frequency Electromagnetic Energy by
  Animals and  by Models of Man. Radio  Sci.,
  14(6S): 15-21.
Gavalas, R.J., D.O. Walter, J. Hamer, and W.R. Adey.
  1970. Effect of Low-Level, Low-Frequency Electric
  Fields on EEG and Behavior in Macaca Nemestrina.
  Brain Res., 18:491-501.
Gavalas-Medici, R., and S.R. Day-Magdaleno. 1976.
  Extremely Low  Frequency, Weak Electric  Fields
  Affect Schedule-Controlled Behavior of Monkeys.
  Nature, 261:256-258.
Ginoza, W., and R.C. Miller, 1965. Kinetics of X-Ray
  and Heat Inactivation of 4>X174 RF-DNA. Proc. Natl.
  Acad. Sci. USA,  54:551-558.
Ginoza, W., C.J. Hoelle, K.B. Vessey, and C. Carmack.
  1964. Mechanisms  of Inactivation of  Single-
  Stranded Virus  Nucleic Acids by Heat.  Nature,
  203(4945):606-609.
Ginzburg, V.L 1968. The Problem of High Temperature
  Superconductivity. Contemp. Phys., 9:355-374.

Goldblith, S.A., and  D.I.C. Wang.  1967. Effect of
  Microwaves on  Escherichia coli and  Bacillus
  subtilus. Appl. Microbiol., 15:1371-1375.
Goldstein, L.,  and Z. Sisko. 1974. A  Quantitative
  Electroencephalographic Study of the Acute
  Effects of X-Band Microwaves in  Rabbits. In:
  Biologic Effects and Health Hazards of Microwave
  Radiation, P. Czerski, K. Ostrowski, M.L. Shore, C.
  Silverman,  M.J. Suess, and B. Waldeskog, eds.
  Polish Medical Publishers, Warsaw, Poland, pp.
  128-133.

Goodman, E.M., B. Greenebaum, and M.T. Marron.
  1979. Bioeffects of  Extremely Low-Frequency
  Electromagnetic  Fields. Radiat. Res., 78:485-501.
Goodman, R., C.A. L.  Bassett, and A.S. Henderson.
  1983. Pulsing Electromagnetic Fields Induce
  Cellular Transcription. Science, 220:1283-1285.
Gordon, C.J. 1982a. Effects of Ambient Temperature
  and Exposure to 2450-MHz Microwave Radiation
  on Evaporative  Heat Loss in the Mouse.  J.
  Microwave Power, 17:145-150.
Gordon, C.J. 1982b. Open-Loop Gain of Evaporative
  Heat Loss During Radiant Heat Exposure in the
  Mouse. Am. J. Physiol. 242:R275-R279.
Gordon, C.J.  1982c.  Effect of Heating  Rate  of
  Evaporative Heat Loss in the Microwave-Exposed
  Mouse. J. Appl. Physiol., 53(2):316-323.

Gordon, C.J. 1982d. Rewarming Mice from Hypo-
  thermia  by Exposure to 2450-MHz Microwave
  Radiation. Cryobiology, 19:428-434.
Gordon, C.J.  1983a. Behavioral and Autonomic
  Thermoregulation in Mice  Exposed to Microwave
  Radiation. J. Appl. Physiol., 55:1242-1248.

Gordon, C.J. 1983b. Influence of Heating Rate on
  Control of Heat Loss from the Tail in Mice. Am. J.
  Physiol., 244.-R778-R784.
Gordon, C.J. 1983c. Effect of 2450 MHz Microwave
  Exposure on Behavioral Thermoregulation in Mice.
  J. Thermal Biol.,  8:315-319.
                      R-10

-------
Gordon, C.J. 1983d. A Review of Terms for Regulated
  vs.  Forced Neurochemical-lnduced Changes in
  Body Temperature. Life Sci., 32:1285-1295.

Gordon, C.J. 1983e. Note: Further Evidence for an
  Inverse Relation Between Body Mass and Sensiti-
  vity to Radio-frequency Electromagnetic Radiation.
  J. Microwave Power, 18:377-383.
Gordon, C.J., and E.C. White. 1982. Distinction
  Between Heating Rate andTotal Heat Absorption in
  the Microwave-Exposed Mouse. Physiol. Zool.,
  55(3):300-308.
Gordon, R.G., R.B. Roemer, and S.M. Horvath. 1976.
  A Mathematical Model of the Human Temperature
  Regulatory  System  - Transient  Cold Exposure
  Response. IEEE Trans. Biomed. Eng., BME-23:434-
  444.
Gordon, Z.V. 1970. Biological Effect of Microwaves in
  Occupational Hygiene. Israel Program for Scientific
  Translations, Jerusalem, Israel. NASATTF-633,TT
  70-50087; NTIS N71-14632. 101 pp.
Gournay, L.S.  1966. Conversion of Electromagnetic
  to Acoustic Energy by Surface Heating. J. Acous.
  Soc. Amer., 40:1322-1330.
Grant, E.H., S.E. Keefe,  and S. Takashima. 1968. The
  Dielectric Behavior of Aqueous Solutions of Bovine
  Serum  Albumin from Radiowave  to  Microwave
  Frequencies. J. Phys. Chem., 72:4373-4380.

Grant, E.H., RJ.  Sheppard, and G.P. South. 1978.
  Dielectric Behaviour of Biological Molecules in
  Solution. Oxford University Press, Oxford,  England.
  237 pp.

Greene, F.M.  1975a. Development  of Electric  and
  Magnetic Near-Field  Probes.  NBS Technical Note
  658 (COM-75-50161), U.S.  Department  of Com-
  merce,  National Bureau  of  Standards,  Boulder,
  Colorado. 47 pp.
Greene, F.M. 1975b. Development of Magnetic Near-
  Field Probes. HEW Publication (NIOSH) 75-127,
  U.S. Department of Health, Education,  and
  Welfare, Public Health Service, Cincinnati, Ohio.
  28pp.
Greene, F.M. 1976. Development of an RF Near-Field
  Exposure Synthesizer  (10 to 40 MHz). HEW
  Publication (NIOSH) 76-160. U.S. Department of
  Health, Education, and Welfare,  Public Health
  Service, Cincinnati, Ohio. 36 pp.
Greengard, P.,  W.W.  Douglas, A.C.  Nairn,  E.J.
  Nestler,  and  J.M.  Ritchie. 1982.  Effects of
  Electromagnetic Radiation on Calcium in the Brain.
  USAF School  of Aerospace Medicine, Report
  Number SAM-TR-82-15. 113 pp.
Grell, R.F.  1971. Heat-Induced Exchange in  the
  Fourth Chromosome of Diploid  Females of
  Drosophila Melanogaster. Genetics, 69:523-527.
Grundler,  W.,  and F. Keilmann. 1980. Frequency
  Fine-Tuning Studies of Microwave  Influenced
  Yeast Growth (Abstract). Presented at the
  International Symposium on Electromagnetic
  Waves and Biology, Jouy-en-Josas, France, June
  30-July4, 1980. p. 4.

Grundler, W., and F. Keilmann. 1983. Sharp Reso-
  nances in Yeast Growth Prove Nonthermal Sensi-
  tivity to  Microwaves.  Phys. Rev. Lett, 51:1214-
  1216.
Grundler, W., F. Keilmann, and H. Frbhlich. 1977. Res-
  onant Growth  Rate  Response  of Yeast  Cells
  Irradiated by Weak  Microwaves. Phys. Lett.,
  62A:463-466.
Guillet, R., and S.M. Michaelson. 1977. The Effect of
  Repeated Microwave Exposure on Neonatal Rats.
  Radio Sci., 12(6S):125-129.

Gunn, S.A., T.C. Gould, and W.A.D. Anderson. 1961.
  The Effect of Microwave Radiation on Morphology
  and Function of Rat Testis. Lab. Invest., 10(2): 301 -
  314.
Gutman,  R., and  B.K.  Chang. 1982. Effect  of
  Moderate Hyperthermia on the Mitogen Response
  of Mouse and Hamster Lymphocytes In Vitro. In:
  Biomedical Thermology, M. Gautherie  and  E.
  Albert, eds. Alan R. Liss, Inc., New York. pp. 75-84.
Guy, A.W. 1971. Analyses of Electromagnetic Fields
  Induced in Biological  Tissues by Thermographic
  Studies on Equivalent Phantom Models. IEEE
  Trans. Microwave Theory Techniques, MTT-
  19(2):205-214.

Guy, A.W. 1975. Correspondence on D.R. Justesen's
  "Prescriptive Grammar for the Radiobiology  of
  Nonionizing  Radiation." J. Microwave  Power,
  10(4):358-359.
Guy, A.W. 1977. A Method for Exposing Cell Cultures
  to Electromagnetic  Fields under Controlled
  Conditions of Temperature and  Field Strength.
  Radio Sci., 12(6S):87-96.
Guy, A.W. 1979. Miniature Anechoic Chamber for
  Chronic Exposure of Small Animals to Plane-Wave
  Microwave Fields. J. Microwave Power, 14(4):327-
  338.

Guy, A.W.,  and C.K. Chou. 1976. System for
  Quantitative Chronic Exposure of a Population  of
  Rodents to UHF Fields. In: Biological Effects  of
  Electromagnetic Waves, Vol. II, C.C. Johnson and
  M.L  Shore, eds. HEW Publication (FDA) 77-8011,
  Rockville, Maryland, pp. 389-410.

Guy, A.W., and S.F. Korbel. 1972. Dosimetry Studies
  of a  UHF Cavity Exposure  Chamber for Rodents.
  Summaries of Presented Papers for 1972 Micro-
  wave Power Symposium,  Ottawa, Canada. Int.
                                                                   R-11

-------
  Microwave Power Institute, Edmonton, Alberta, pp.
  180-193.
Guy, A.W., J.F. Lehmann, J.A. McDougall, and C.C.
  Sorensen.  1968. Studies on Therapeutic Heating
  by Electromagnetic Energy. In: Thermal Problems
  in Biotechnology, American Society of Mechanical
  Engineers, United Engineering Centers, New York,
  New York. pp. 26-45.
Guy, A.W., C.C. Johnson, J.C. Lin, A.F. Emery, and
  K.K. Kraning.  1973. Electromagnetic  Power
  Deposition in Man  Exposed to High  Frequency
  Fields and the Associated Thermal and Physiologic
  Consequences. Doc. No. SAM-TR-73-13(NTIS AD-
  776 821),  USAF School of Aerospace Medicine,
  Brooks Air Force Base, Texas. 71 pp.
Guy, A.W., J.C. Lin, P.O. Kramar, and A.F.  Emery.
  1974. Quantitation of Microwave Radiation Effects
  on the Eyes of Rabbits at 2450 MHz and 918 MHz.
  Office of Naval Research, Arlington, Virginia. (NTIS
  AD-A007521). 39pp.
Guy, A.W., J.C. Lin, P.O. Kramar, and A.F.  Emery.
  1975a. Effect of 2450-MHz Radiation on the Rabbit
  Eye. IEEE Trans.  Microwave Theory Techniques,
  MTT-23:492-498.

Guy, A.W., C.K. Chou, J.C. Lin, and D. Christensen.
  1975b. Microwave-Induced Acoustic Effects in
  Mammalian  Auditory  Systems and Physical
  Materials. Ann. N.Y. Acad. Sci., 247:194-215.
Guy, A.W., M.D. Webb, and J.A. McDougall. 1975c. A
  New Technique for Measuring  Power Deposition
  Patterns in Phantoms Exposed to EM Fields of
  Arbitrary Polarization: Example the  Microwave
  Oven. Microwave Power Symposium  Proc.,
  Waterloo, Ontario, Canada, pp. 36-47.
Guy, A.W., M.D. Webb, and C.C. Sorenson. 1976.
  Determination of Power Absorption in Man
  Exposed to High Frequency Electromagnetic Fields
  by Thermographic  Measurements  on  Scale
  Models. IEEE  Trans. Biomed. Eng., BME-23(5):361 -
  371.
Guy, A.W., M.D. Webb, and J.A. McDougall. 1977. RF
  Radiation Absorption Patterns: Human and Animal
  Modeling Data. HEW Publication (NIOSH) 77-183,
  U.S. Department of Health, Education, and
  Welfare, Public Health Service, Cincinnati, Ohio.
  67pp.
Guy, A.W., M.D. Webb, A.F. Emery, and C.K. Chou.
  1978. Determination of the Average SAR and SAR
  Patterns in Man and Simplified Models of Man and
  Animals Exposed to Radiation Fields from 50-2450
  MHz and the Thermal Consequences (Abstract).
  Symposium on the  Biological Effects of Electro-
  magnetic Waves,  XIX General Assembly, Interna-
  tional Union of Radio Science, Helsinki, Finland, p.
  13.
Guy, A.W., J. Wallace, and J.A. McDougall. 1979.
  Circularly Polarized 2450-MHz Waveguide Systems
  for  Chronic  Exposure of  Small Animals to
  Microwaves. Radio Sci., 14(6S):63-74.
Guy, A.W., C.K. Chou, R.B. Johnson, and L.L Kunz.
  1980a. Study of Effects of Long-Term, Low-Level
  RF Exposure on Rats: A Plan. Proc. IEEE, 68(1 ):92-
  97.
Guy, A.W., P.O. Kramar, C.A. Harris, and C.K. Chou.
  1980b.  Long-Term 2450-MHz CW Microwave
  Irradiation of Rabbits: Methodology and Evaluation
  of Ocular and Physiologic Effects. J. Microwave
  Power, 15:37-44.
Hagan,  G.J.,  and R.L.  Carpenter. 1976. Relative
  Cataractogenic Potencies  of Two Microwave
  Frequencies (2.45 and 10  GHz). In:  Biological
  Effects of Electromagnetic  Waves, Vol. I, C.C.
  Johnson and M.L. Shore, eds.  HEW Publication
  (FDA) 77-8010, Rockville, Maryland, pp. 143-155.

Hagmann, M.J., and O.P. Gandhi. 1979.  Numerical
  Calculation of Electromagnetic Energy Deposition
  in Models of Man with Grounding and Reflector
  Effects. Radio Sci., 14(6S):23-29.

Hagmann, M.J., O.P.  Gandhi, and C.H. Durney.
  1979a. Numerical Calculation of Electromagnetic
  Energy Deposition for a Realistic Model of Man.
  IEEE Trans. Microwave Theory Techniques,  MTT-
  27(9):804-809.
Hagmann, M.J., O.P. Gandhi, J.A. D'Andrea, and I.
  Chatterjee.  1979b. Head Resonance: Numerical
  Solutions and Experimental  Results. IEEE Trans.
  Microwave  Theory and Techniques,  MTT-27(9):
  809-813.

Hamburger, S.. J.N. Logue, and P.M. Silverman,
  1983. Occupational Exposure to Nonionizing Radi-
  ation and an Association with Heart Disease: An
  Exploratory Study. J. Chronic Disease, 36:791 -802.

Hamer, J. 1968. Effects of Low Level, Low Frequency
  Electric Fields on Human Reaction Time. Commun.
  Behav. Biol., 2(5) Part A:217-222.

Hammel, H.T.  1968. Regulation of Internal  Body
  Temperature. Annu. Rev. Physiol., 30:641-710.

Hamnerius, Y. 1983. Exposure Systems for Studies of
  the Effects of Electromagnetic Fields on Biological
  Systems. Hereditas, 98:48-59.
Hamnerius, Y., H. Olofsson, A. Rasmuson, and B.
  Rasmuson.  1979. A Negative Test for Mutagenic
  Action of Microwave Radiation in Drosophila
  melanogaster. Mutat. Res., 68:217-223.

Hamrick, P.E.  1973. Thermal Denaturation of DNA
  Exposed to 2450 MHz  CW Microwave  Radiation.
  Radiat. Res., 56:400-404.
                      R-12

-------
Hamrick,  P.E., and B.T. Butler.  1973.  Exposure of
  Bacteria to 2450 MHz Microwave Radiation. J.
  Microwave Power, 8:227-233.
Hamrick, P.E., and S.S. Fox. 1977. Rat Lymphocytes
  in  Cell Culture Exposed to  2450 MHz  (CW)
  Microwave Radiation. J. Microwave  Power,
  12:125-132.
Hamrick, P.E., and D.I. McRee. 1975. Exposure of the
  Japanese Quail Embryo to 2.45-GHz Microwave
  Radiation During the Second Day of Development.
  J. Microwave Power, 10:211-221.

Hamrick, P., and D.I. McRee. 1980. The Effect of 2450
  MHz Microwave Irradiation on the Heart Rate of
  Embryonic Quail. Health Phys., 38:261 -268.
Hamrick,  P.E., and J.G.  Zinkl. 1975.  Exposure of
  Rabbit Erythrocytes to Microwave Radiation.
  Radiat. Res., 62:164-168.
Hamrick,  P.E., D.I. McRee,  P. Thaxton, and  C.R.
  Parkhurst. 1977. Humoral  Immunity  of Japanese
  Quail Subjected to  Microwave Radiation During
  Embryogeny. Health Phys., 33:23-33.
Hardy, J.D.  1949. Heat Transfer. In: Physiology of
  Heat Regulation,  L.H.  Newburgh,  ed. W.B.
  Saunders, Philadelphia, Pennsylvania, pp. 78-108.
Hardy, J.D., and P. Bard. 1974. Body Temperature
  Regulation. In: Medical Physiology, 13th ed., V.B.
  Mountcastle, ed. C.V. Mosby Co., St. Louis,
  Missouri,  pp. 1305-1342.

Hardy, J.D., AT.  Milhorat, and E.F.  Dubois. 1941.
  Basal Metabolism and Heat Loss of Young Women
  at  Temperatures from  22°C to 35°C. J. Nutr.,
  21:383-404.
Hardy, J.D., H.G. Wolff, and H.  Goodell. 1967. Pain
  Sensations and Reactions.  Hafner Publishing Co.,f
  New York, New York. Chapter X.

Harrison, G.H., J.E. Robinson, D. McCulloch, andA.Y.
  Cheung.  1980. Comparison  of Hyperthermal
  Cellular Survival in the Presence or Absence of
  2.45 GHz Microwave Radiation. In:  Ondes
  Electromagnetiques et Biologie, A.J. Berteaud and
  B. Servantie, eds. Paris, France, pp. 41-45.
Hart, J.S. 1971. Rodents. In: Comparative Physiology
  of  Thermoregulation, Vol.  II,  G.C. Whitlow, ed.
  Academic Press, New York, New York. pp. 1-149.
Heller,  J.H.  1970. Cellular  Effects  of Microwave
  Radiation.  In:  Biological Effects and  Health
  Implications of  Microwave  Radiation, S.F. Clean/,
  ed. HEW  Publication BRH/DBE 70-2. Bureau of
  Radiological Health, Rockville, Maryland, pp. 116-
  121.

Henderson,  H.M., K. Hergenroeder, and S.S. Stuchly.
  1975. Effect of 2450 MHz Microwave Radiation on
  Horseradish Peroxidase. J.  Microwave  Power,
  10:27-35.
Hendler, E. 1968. Cutaneous Receptor Response to
  Microwave  Irradiation. In: Thermal  Problems in
  Aerospace Medicine, J.D. Hardy, ed.  Technivision
  Services, Maidenhead, England, pp. 149-161.

Hendler, E., J.D. Hardy, and D. Murgatroyd. 1963.
  Skin Heating and Temperature Sensation Produced
  by Infra Red and Microwave Irradiation. In:
  Temperature: Its Measurement  and Control in
  Science and Industry. Part 3, Biology and Medicine,
  C.M. Herzfeld, ed. Reinhold, New York, New York.
  pp. 211-230.
Hensel, H.  1973. Neural Processes in Thermoregula-
  tion. Physiol. Rev., 53:948-1017.
Heynick, L.N.,  P.  Poison, and  A.  Karp.  1977. A
  Microwave Exposure System for Primates. Radio
  Sci., 12:103-110.
Hinde, R.A. 1970. Animal Behavior: A Synthesis of
  Ethology  and Comparative  Psychology, 2nd ed.
  McGraw-Hill, New York, New York. 876 pp.

Hizal,  A.,  and Y.K.  Baykal. 1978. Heat Potential
  Distribution in an Inhomogeneous Spherical Model
  of a Cranial Structure Exposed to Microwaves Due
  to Loop or Dipole Antennae. IEEE Trans. M icrowave
  Theory Techniques, MTT-26(8):607-612.
Hjeresen, D.L., S.R. Doctor, and R.L Sheldon. 1979.
  Shuttlebox Side Preference as Mediated by Pulsed
  Microwave and Conventional Auditory Cues. In:
  Electromagnetic Fields in Biological Systems, S.S.
  Stuchly,  ed. Ottawa, Canada, pp. 194-214.

Ho, H.S.  1975. Dose Rate Distribution  in Triple-
  Layered  Dielectric Cylinder with  Irregular Cross
  Section  Irradiated  by Plane  Wave  Sources. J.
  Microwave Power, 10(4):421-432.

Ho, H.S. 1978. Effect of Plexiglas Animal Holders on
  Microwave  Energy  Absorption.  IEEE Trans.
  Biomed.  Eng., BME-25(5):474-476.
Ho, H.S., and W.P. Edwards. 1977a. Dose Rate and
  Oxygen Consumption  Rate in  Mice Confined in a
  Small Holder During Exposure to 2450  MHz
  Radiation. Radiat. Environ. Biophys.,  14:251-256.
Ho, H.S.,  and W.P. Edwards.  1977b. Oxygen-
  Consumption Rate of  Mice under Differing  Dose
  Rates of Microwave Radiation.  Radio  Sci.,
  12(6S):131-138.
Ho, H.S., and W.P.  Edwards. 1979. The Effect of
  Environmental  Temperature  and Average  Dose
  Rate of  Microwave Radiation on the Oxygen-
  Consumption Rate of Mice. Radiat. Environ.
  Biophys. 16:325-338.
Ho, H.S., andM. McManaway. 1977. Heat-Dissipation
  Rate of   Mice after  Microwave Irradiation.  J.
  Microwave Power, 12:93-100.
Ho, H.S.,  E.I. Ginns, and  C.L  Christman.  1973.
  Environmentally Controlled Waveguide Irradiation

                   R-13

-------
  Facility. IEEE Trans. Microwave Theory Techniques,
  MTT-21:837-840.
Ho. H.S.,  M.R.  Foster, and M.L  Swicord.  1976.
  Microwave Irradiation Apparatus  Design and
  Dosimetry. In: Biological Effects of Electromagnetic
  Waves, Vol. II, C.C. Johnson and M.L Shore, eds.
  HEW Publication (FDA)  77-8011,  Rockville,
  Maryland, pp. 423-434.
Hollaender, A. 1971. Chemical Mutagens, Principles
  and Methods for Their Detection. Vols. 1, 2, and 3.
  Plenum Press, New York, New York.
Honig,  W.K.,  and  J.E.R.  Staddon,  eds.  1977.
  Handbook of Operant Behavior. Prentice-Hall,
  Englewood Cliffs, New Jersey. 689 pp.
Hossain, M., and S.K. Dutta. 1982. Comparison of
  Bacterial Growth to High-Intensity Microwave
  Exposure and Conventional Heating. Bioelectro-
  magnetics, 3:471-474.
Huang, AT., and N.G. Mold. 1980. Immunologic and
  Hematopoietic Alterations by 2450-MHz Electro-
  magnetic Radiation. Bioelectromagnetics, 1:77-
  87.
Huang, AT., M.E. Engle, J.A. Elder, J.B. Kinn, andT.R.
  Ward. 1977.  The  Effect of Microwave Radiation
  (2450 MHz) on the Morphology and Chromosomes
  of Lymphocytes. Radio Sci., 12(6S):173-177.

Hunt, E.L, and R.D. Phillips. 1972. Absolute Physical
  Dosimetry for Whole Animal Experiments. Digest
  of Papers of the Microwave Dosimetry Workshop,
  Atlanta, Georgia, pp. 74-77.

Hunt, E.L, N.W.  King, and R.D.  Phillips.  1975.
  Behavioral Effects of Pulsed Microwave Radiation.
  Ann. N.Y. Acad. Sci., 247:440-453.

Hunter, W.S., K.R. Holmes, and R.S. Elizondo. 1981.
  Thermal  Balance in Ketamine-Anesthetized
  Rhesus Monkey Macaca mulata. Am. J. Physiol.,
  241 :R301 -R306.

IEEE Microwave Theory and Techniques  Society.
  1980. Symposium on Electromagnetic Dosimetric
  Imagery, Washington, D.C. Institute of Electrical
  and Electronic Engineers.
Illinger, K.H. 1970. Molecular Mechanisms for
  Microwave Absorption in  Biological  Systems.  In:
  Biological Effects and Health Implications  of
  Microwave Radiation, S.F.  Clean/, ed. BRH/DBE
  70-2. Bureau of  Radiological Health, Rockville,
  Maryland, pp. 112-115.

Illinger, K.H. 1982.  Spectroscopic Properties of In
  Vivo Biological Systems:  Boson Radiative Equili-
  brium with Steady-State Nonequilibrium Molecular
  Systems. Bioelectromagnetics, 3:9-16.
Iskander, M.F., C.H.  Durney, H. Massoudi, and C.C.
  Johnson. 1979. Approximate Calculation of SAR

                      R-14
  for  Planewave Irradiation of Man  Model Near a
  Ground Plane. In: Electromagnetic  Fields in
  Biological Systems, S.S. Stuchly, ed. International
  Microwave Power  Institute,  Edmonton, Alberta,
  Canada, pp. 304-323.

Iskander, M.F.,  P.W.  Barber, C.H. Durney, and H.
  Massoudi. 1980. Irradiation of Prolate Spheroidal
  Models of Humans in the Near Field of a Short
  Electric  Dipole. IEEE Trans. Microwave Theory
  Techinques, MTT-28{7):801 -807.

Ismailov, E. Sh.  1971.  Mechanism of Effects of
  Microwaves  on Erythrocyte  Permeability for
  Potassium and Sodium Ions. Biol. Nauki, 3:58-60
  (English trans.: JPRS 72606, Jan. 12,1979, pp. 38-
  41).

Ismailov, E. Sh. 1977. Infrared Spectra of Erythrocyte
  Ghosts in the Region of the Amide I and Amide II
  Bands on Microwave Irradiation. Biophysics,
  21:961 -963 (trans, of Biofizika 21:940-942,1976).
Ismailov, E. Sh.  1978. Effect of Ultrahigh Frequency
  Electromagnetic Radiation on the Electrophoretic
  Mobility of Erythrocytes.  Biophysics,  22:510-516
  (trans, of Biofizika 22:493-498,  1977).
Jensh,  R.P. 1979. Biological Effects of 6 GHz
  Microwave Irradiation. Final Report, GTEL Grant
  08000-1106.  131 pp.
Jensh, R.P. 1980. Behavioural Teratology: Application
  in Low Dose Chronic Microwave Irradiation
  Studies. In: Neural and Behavioural Teratology
  (Vol. 4, Advances in the Study of Birth  Defects),
  T.V.N. Persaud, ed.  University Park Press,
  Baltimore, Maryland, pp. 135-162.

Jensh, R.P., J.  Ludlow, W.H. Vogel, T. McHugh, I.
  Weinberg, and R.L. Brent. 1979. Studies Concern-
  ing  the Effects of Non-Thermal Protracted Prenatal
  915 MHz Microwave Radiation on Prenatal and
  Postnatal Development in the Rat. XIV International
  Symposium on the Applications of Microwave
  Energy, IMPI,  Paris, France, pp.  99-101.
Johansen, K. 1963. Cardiovascular Dynamics in the
  Amphibian, Amphiuma  tridactylum Cuvier.  Acta
  Physiol. Scand., 60:1-82.
Johnson, C.C. 1975. Recommendations for Specifying
  EM Wave Irradiation  Conditions in Bioeffects
  Research. J. Microwave Power,  10(3):249-250.

Johnson, C.C.,  and A.W. Guy. 1972.  Nonionizing
  Electromagnetic Wave Effects  in Biological
  Materials and Systems. Proc. IEEE, 60:692-718.
Johnson, R.B., D.E. Meyers, A.W. Guy, R.H. Lovely,
  and R. Galambos. 1976. Discriminative Control of
  Appetitive Behavior by Pulsed Microwave Radiation
  in Rats. In: Biological Effects of Electromagnetic
  Waves, Vol. I, C.C. Johnson and M.L. Shore, eds.
  HEW  Publication  (FDA)  77-8010, Rockville,
  Maryland, pp.  238-247.

-------
Johnson, R.B., S. Mizumori, and R.H. Lovely. 1978.
  Adult Behavioral Deficit in Rats Exposed Prenatally
  to 918-MHz Microwaves.  In: Developmental
  Toxicology of Energy-Related Pollutants. D.D.
  Mahlum, M.R. Sikov,  P.L. Hackett, and F.D.
  Andrew, eds. DOE Symposium Series 47, Wash-
  ington, D.C. pp. 281-299.

Joines, W.T., and  C.F. Blackman. 1980. Power
  Density,  Field Intensity, and Carrier Frequency
  Determinants of RF-Energy-lnduced  Calcium  Ion
  Efflux from Brain Tissue. Bioelectromagnetics
  1:271-275.
Joines, W.T., and C.F. Blackman. 1981. Equalizing
  the  Electric Field  Intensity  Within  Chick Brain
  Immersed in  Buffer Solution at Different Carrier
  Frequencies. Bioelectromagnetics, 2:411-413.

Joines, W.T., C.F. Blackman, and M.A. Hollis. 1981.
  Broadening of the RF Power-Density Window  for
  Calcium-Ion Efflux from Brain Tissue. IEEE Trans.
  Biomed. Eng., BME-28:568-573.

Justesen, D.R. 1975. Toward a Prescriptive Grammar
  for the  Radiobiology of  Non-Ionising Radiations:
  Quantities, Definitions, and Units of Absorbed
  Electromagnetic Energy—An Essay. J. Microwave
  Power, 10(4):343-356.
Justesen, D.R. 1980. Microwave Irradiation and
  Blood-Brain Barrier. Proc. IEEE, 68:60-67.

Justesen, D.R., D.M. Levinson, R.L. Clarke, andN.W.
  King. 1971. A Microwave Oven for Behavioural and
  Biological  Research:  Electrical and Structural
  Modifications, Calorimetric,  Dosimetry, and
  Functional Evaluation. J. Microwave  Power,
  6:237-258.
Justesen,  D.R.,  D.M. Levinson, and  L.R.  Justesen.
  1974. Psychogenic Stressors are Potent Mediators
  of the Thermal Response to Microwave Irradiation.
  In: Biologic  Effects  and  Health Hazards  of
  Microwave Irradiation, P. Czerski, K. Ostrowski,
  M.L.  Shore, C. Silverman,  M.J.  Suess,  and  B.
  Waldeskog,  eds.  Polish  Medical Publishers,
  Warsaw, Poland, pp. 134-140.

Justesen, D.R.,  H.A. Ragan, L.E. Rogers, A.W. Guy,
  D.L  Hjeresen, W.T. Hinds, and R.D. Phillips. 1978.
  Compilation  and  Assessment of  Microwave
  Bioeffects. A Selective Review of the Literature on
  Biological Effects of Microwaves in Relation to the
  Satellite  Power System (SPS). PNL-2634. U.S.
  Department of Energy, Washington, D.C. 65 pp.

Justesen,  D.R., E.R.  Adair, J.C.  Stevens,  and  V.
  Bruce-Wolfe. 1982. A  Comparative  Study of
  Human Sensory Thresholds: 2450-MHz Micro-
  waves vs Far-Infrared Radiation. Bioelectromagne-
  tics,  3:117-125.
Kaczmarek, L.K., and W.R. Adey. 1973. The Efflux of
  45Ca2* and [3H]x-Aminobutyric Acid  from Cat
  Cerebral Cortex. Brain Res., 63:331-342.
Kaczmarek, L.K., and W.R. Adey. 1974. Weak Electric
  Gradients Change Ionic and Transmitter Fluxes in
  Cortex. Brain  Res., 66:537-540.
Kallen,  B., G.  Malmquist, and  U.  Moritz. 1982.
  Delivery  Outcome Among Physiotherapists  in
  Sweden: Is Non-Ionizing Radiation a Fetal Hazard?
  Arch.  Environ. Health, 37:81-85.

Kandel, E.R., and J.H. Schwartz, eds. 1981. Principles
  of Neural  Science. Elsevier North Holland, Inc.,
  New York, New York.  749 pp.

Kantor,  G., and T.C. Cetas.  1977. A Comparative
  Heating-Pattern Study of Direct-Contact Applicators
  in Microwave Diathermy. Radio Sci.,  12(6S):111-
  120.
Kaplan,  J.N. 1981.  Study  of the Lethal  Effects  of
  Microwaves in the Developing Squirrel Monkey.
  Final  Report for  Contract No. 68-02-3210, U.S.
  Environmental Protection Agency,  Health Effects
  Research  Laboratory, Research  Triangle Park,
  North  Carolina. 54 pp.
Kaplan,  IT., W.  Metlay,  M.M. Zaret,  L Birenbaum,
  and S.W. Rosenthal. 1971. Absence of Heart-Rate
  Effects in  Rabbits During Low-Level Microwave
  Irradiation. IEEE Trans. Microwave Theory Tech-
  niques MTT-19:168-173.
Kaplan,  J., P. Poison, C. Rebert, K.  Lunan,  and  M.
  Gage. 1982. Biological and Behavioral Effects  of
  Prenatal and  Postnatal  Exposure  to  2450-MHz
  Electromagnetic Radiation in the Squirrel Monkey.
  Radio  Sci.: 17(5S):135S-144S.

Keilmann, F.  1978. Nonthermal Microwave Reso-
  nances in Living Cells. In: Coherence in Spectro-
  scopy  and Modern  Physics,  F.T. Arecchi,  R.
  Bonifacio, and M.O. Scully, eds. NATO Advanced
  Study Institute Series: Series B, Physics, Vol. 37.
  Plenum Publishing Corp., New York, New York. pp.
  347-360.
Keller, S.E., J.M. Weiss, S.J. Schleifer, N.E. Miller,
  and M. Stein.  1983- Stress-Induced Supression of
  Immunity  in  Adrenalectomized Rats. Science,
  221:1301-1304.

Kerslake, D. McK., and J.L. Waddell. 1958. The Heat
  Exchange of Wet  Skin. J. Physiol., 141:156-163.

Keys, A. 1970.  Coronary Heart Disease in Seven
  Countries. American Heart Association Monograph
  No. 29. pp. 1-198.
King, N.W., D.R.  Justesen, and A.D. Simpson. 1970.
  The Photo-Lickerandum: A Device for Detecting the
  Licking Response with Capability  for Near-
  Instantaneous Programming of Variable Quantum
                                                                    R-15

-------
  Reinforcement.  Behav.  Res.  Meth. Instrument.,
  2:125-129.
King, N.W.,  D.R. Justesen, and R.L Clarke.  1971.
  Behavioral Sensitivity to Microwave Irradiation.
  Science, 172:398-401.
Kinn, J.B. 1977. Whole-Body Dosimetry of Microwave
  Radiation  in Small Animals: The Effect of Body
  Mass and Exposure Geometry. Radio Sci.,
  12(6S):61-64.
Kinn, J.B., G.E. Anderson, and W.M. Kozel. 1984. A
  Microcomputer Controlled Calorimeter.  J. Micro-
  wave Power, In Press.
Kirkwood,  J.G., and J.B. Schumaker. 1952. The
  Influence of Dipole Moment Fluctuations on the
  Dielectric Increment of Proteins in Solution. Proc.
  Natl. Acad. Sci., 38:855-862.

Kitsovskaya,  LA. 1964. The  Effect of Centimeter
  Waves of  Different Intensities on the Blood and
  Hemopoietic Organs of White Rats. Gigiena Truda
  Prof Zabolev, 8:14-25.
Kleiber, M. 1972. A New Newton's Law of Cooling?
  Science, 178:1283-1285.
Kleiber, M. 1975. The Fire of Life. An Introduction to
  Animal Energetics, rev. ed. R.E. Krieger, Huntingdon,
  New York.  453 pp.
Kling, J.W.,  and LA.  Riggs, eds.  1971. Woodworth
  and Schlosberg's Experimental Psychology, 3rd ed.
  Holt Rinehart and Winston, New York, New York.
  1279pp.
Konig, H. 1971. Biological Effects of Extremely Low
  Frequency  Electrical Phenomena in the Atmo-
  sphere. J.  Interdiscipl. Cycle Res., 2:317-323.

Konig,  H., and F. Ankermuller.  1960. Uber den
  Einfluss besonders nieder-frequenter elektrischer
  vorgange in der Atmosphare auf den Menschen.
  Naturwissenschaften, 21:486-490.

Konorski, J.  1967. Integrative Activity of the  Brain.
  The University of Chicago Press,  Chicago, Illinois.
  531 pp.
Kramar, P.O., A.F. Emery, A.W. Guy, and J.C. Lin.
  1975. The Ocular Effects of  Microwaves on
  Hypothermic Rabbits: A Study of  Microwave
  Cataractogenic Mechanisms. Ann. N.Y. Acad. Sci.,
  247:155-165.
Kramar, P.O., C. Harris, A.W.  Guy, and A.F. Emery.
  1976. Mechanism of Microwave Cataractogenesis
  in Rabbits. In: Biological Effects of Electromagnetic
  Waves, Vol. 1, C.C. Johnson and M.L Shore, eds.
  HEW Publication (FDA)  77-8010, Rockville,
  Maryland,  pp. 49-60.
Kramar, P.,  C. Harris, A.F. Emery, and A.W. Guy.
  1978. Acute Microwave Irradiation and Cataract
  Formation in Rabbits and Monkeys. J. Microwave
  Power, 13:239-249.
Kremer, F., Chr. Koschnitzke, L. Santo, P. Quick, and
  A. Poglitsch. 1983. The Non-thermal Influence of
  Millimeter Wave Radiation on the Puffing of Giant
  Chromosomes. Abstract #H-2,5th Annual Scientific
  Session, The Bioelectromagnetics Society, 12-17
  July,  Boulder,  Colorado (available  from The
  Bioelectromagnetics Society,  One Bank Street,
  Suite 307, Gaithersburg, Maryland 20878). p. 56.
Kritikos, H.N., and H.P. Schwan. 1975. The Distribution
  of Heating Potential Inside Lossy Spheres.  IEEE
  Trans. Biomed. Eng., BME-22(6):457-463.
Krupp, J.H.  1977. Thermal Response in Macaca
  Mulatta Exposed to 15- and 20-MHz Radiofrequency
  Radiation. Report No. SAM-TR-77-16, USAF
  School of Aerospace Medicine, Brooks Air Force
  Base, Texas. 10 pp.

Lancranjan, I., M. Maicanescu,  E. Rafaila, I. Klepsch,
  and H.I.  Popescu. 1975. Gonadic Function in
  Workmen with Long-Term Exposure to Microwaves.
  Health  Phys., 29:381-383.
Lappenbusch, W.L., LJ. Gillespie, W.M. Leach, and
  G.E.'Anderson.  1973.  Effect  of 2450-MHz
  Microwaves on the Radiation Response of X-
  Irradiated Chinese Hamsters. Radiat. Res., 54:294-
  303.
Lebovitz,  R.M., and R.L. Seaman. 1977. Microwave
  Hearing: The Response of Single Auditory Neurons
  in the Cat to Pulsed Microwave Radiation. Radio
  Sci. 12(65)229-236.
Lehmann, J.F., A.W. Guy, J.B. Stonebridge, and B.J.
  deLateur. 1978. Evaluation of a Therapeutic Direct-
  Contact  915-MHz Microwave  Applicator for
  Effective  Deep-Tissue Heating in Humans.  IEEE
  Trans. Microwave Theory Techniques,  MTT-
  26:556-563.
Lehninger, A.L 1975. Biochemistry, 2nd ed. Worth
  Publishers, Inc., New York, New York. pp. 143-144,
  875-876.
LeVeen, H.H., S. Wapnick, V. Piccone, G. Falk, and N.
  Ahmed. 1976. Tumor Eradication by Radiofre-
  quency Therapy. Response in 21  Patients. J. Am.
  Med. Ass., 235:2198-2200.
Levinson, D.M., A.M. Grove, LR. Clarke, and D.R.
  Justesen.  1982. Photic Cuing  of Escape by Rats
  from an  Intense  Microwave  Field. Bioelectro-
  magnetics, 3:105-116.
Liboff, A.R., T. Williams, Jr., D.M.  Strong, and R.
  Wistar, Jr. 1984.  Time-Varying Magnetic Fields:
  Effect on  DNA Synthesis. Science, 223:818-820.
Liburdy,  R.P.  1977. Effects  of Radio-Frequency
  Radiation on Inflammation. Radio Sci., 12(6S): 179-
  183.
                      R-16

-------
Liburdy, R.P. 1979. Radiofrequency Radiation Alters
  the Immune System:  Modulation  of T- and B-
  Lymphocyte  Levels and  Cell-Mediated  Immuno-
  competence  by Hyperthermic Radiation.  Radiat.
  Res., 77:34-46.

Liburdy, R.P. 1980. Radiofrequency Radiation Alters
  the Immune System.  II. Modulation of In  Vivo
  Lymphocyte Circulation. Radiat. Res., 83:66-73.

Lilienfeld, A.M.,  J. Tonascia, S. Tonascia, C.A.
  Libauer, and G.M. Cauthen. 1978. Foreign Service
  Health Status Study—Evaluation of Health Status
  of Foreign Service and Other Employees from
  Selected Eastern European Posts.  Final  Report,
  Contract No. 6025-619073  (NTIS  PB-288163),
  Dept. of State, Washington, D.C. 436 pp.

Lin, J.C. 1977. Theoretical Calculation of Frequencies
  and Thresholds of  Microwave-Induced Auditory
  Signals. Radio Sci.,  12:237-242.

Lin,  J.C. 1978.  Microwave Auditory Effects and
  Applications. Charles C. Thomas, Springfield,
  Illinois. 221 pp.
Lin, J.C., and M.F. Lin. 1980. Studies on Microwave
  and Blood-Brain Barrier Interaction. Bioelectro-
  magnetics, 1:313-323.
Lin, J.C., and M.F. Lin. 1982. Microwave Hyperther-
  mia-lnduced Blood-Brain Barrier Alterations.
  Radiat. Res.,  89:77-87.

Lin, J.C., and W.D. Peterson, Jr.  1977. Cytological
  Effects of 2450 MHz CW Microwave Radiation. J.
  Bioeng., 1:471-478.

Lin, J.C., A.W. Guy, and C.C. Johnson. 1973. Power
  Deposition in a Spherical Model of Man Exposed to
  1-20 MHz Electromagnetic  Fields. IEEE Trans.
  Microwave Theory Techniques, MTT-21(12):791-
  797.

Lin, J.C., H.I. Bassen, and C.L. Wu. 1977. Perturbation
  Effect of Animal Restraining  Materials on Micro-
  wave  Exposure. IEEE Trans. Biomed. Eng., BME-
  24(1):80-83.
Lin,  J.C.,  J.C.  Nelson, and M.E. Ekstrom.  1979a.
  Effects of Repeated Exposure to 148-MHz Radio
  Waves on Growth and  Hematology of Mice. Radio
  Sci., 14:(6S)173-179.

Lin, J.C., M.J.  Ottenbreit, S.  Wang, S. Inoue, R.O.
  Bellinger, and M. Fracassa.  1979b. Microwave
  Effects on Granulocyte  and Macrophage Precursor
  Cells of Mice In Vitro. Radiat.  Res., 80:292-302.

Lindahl,  T.,  and  B. Nyberg. 1974.  Heat-Induced
  Deamination of Cytosine Residues in Deoxyribo-
  nucleic Acid. Biochemistry,  13(16):3405-3410.

Lin-Liu,  S., and W.R. Adey. 1982. Low Frequency
  Amplitude Modulated  Microwave Fields Change
  Calcium Efflux Rates from Synaptosomes. Bioelec-
  tromagnetics, 3:309-322.
Liu, L.M., FJ. Rosenbaum, and W.F. Pickard. 1976.
  The Insensitivity of  Frog Heart Rate to Pulse
  Modulated Microwave Energy. J.  Microwave
  Power, 11:225-232.
Liu, L.M., F.G. Nickless, and S.F. Clean/. 1979. Effects
  of Microwave Radiation on Erythrocyte Membranes.
  Radio Sci., 14(6S):109-115.

Livingston, G.K., C.C. Johnson, and L.A. Dethlefsen.
  1979. Comparative Effects  of  Water-Bath-  and
  Microwave-Induced Hyperthermia on Survival of
  Chinese Hamster Ovary (CHO)  Cells. Radio Sci.,
  14(S):117-123.
Lobanova,  E.A.  1974.  The Use of  Conditioned
  Reflexes to Study Microwave Effects on the Central
  Nervous System. In:  Biologic Effects and Health
  Hazards of Microwave Radiation, P. Czerski, K.
  Ostrowski, M.L Shore, C. Silverman, M.J. Suess,
  and B. Waldeskog, eds. Polish Medical Publishers,
  Warsaw, Poland, pp. 110-118.

Lotz, W.G. 1982. Hyperthermia in Rhesus Monkeys
  Exposed to a Frequency (225 MHz) near Whole-Body
  Resonance. Naval Med. Res. Devel. Com. MF58.
  524.02C-009, Naval Aerospace  Medical Research
  Laboratory, Pensacola, Florida.

Lotz, W.G., and S.M. Michaelson. 1978. Temperature
  and Corticosterone Relationships in  Microwave-
  Exposed Rats. J. Appl. Physiol., 44:438-445.

Lotz, W.G., and S.M. Michaelson. 1979. Effects of
  Hypophysectomy and  Dexamethasone  on  Rat
  Adrenal Response to Microwaves. J. Appl. Physiol:
  Respirat. Environ. Exercise Physiol., 47:1284-
  1288.
Lotz, W.G. and  R.P. Podgorski. 1982.  Temperature
  and Adrenocortical Responses in Rhesus Monkeys
  Exposed to Microwaves. J. Appl. Physiol., 53(6):
  1565-1571.
Lovely,  R.H., D.E. Myers, and A.W.  Guy.  1977.
  Irradiation of Rats by 918-MHz Microwaves at 2.5
  mW/cm2:  Delineating the Dose-Response Rela-
  tionship. Radio Sci., 12(6S): 139-146.

Lu,  S., N. Lebda, S.M. Michaelson, S. Pettit, and D.
  Rivera.  1977. Thermal and Endocrinological
  Effects of Protracted Irradiation of Rats by 2450-
  MHz Microwaves. Radio Sci.,  12:147-156.

Lu, S., N. Lebda, S. Pettit, and S.M. Michaelson. 1981.
  Microwave-Induced Temperature, Corticosterone,
  and Thyrotropin  Interrelationships. J. Appl.
  Physiol.:  Respirat. Environ. Exercise Physiol.,
  50:399-405.
Luben, R.A., C.D. Cain, M.C.-Y. Chen, D.M. Rosen,
  and W.R. Adey. 1982. Effects of Electromagnetic
  Stimuli on Bone and Bone Cells In Vitro: Inhibition
  of Responses to Parathyroid Hormone by Low-
                    R-17

-------
  Energy Low-Frequency Fields. Proc.  Natl.  Acad.
  Sci. USA, 79:4180-4184.
Lyle, D.B., P. Schechter, W.R. Adey, and R.L Lundak.
  1983. Suppression of  T-Lymphocyte  Cytotoxicity
  Following Exposure to Sinusoidally  Amplitude-
  Modulated Fields.  Bioelectromagnetics, 4:281-
  292.

Machle, W.,  and T.F. Hatch. 1947. Heat: Man's
  Exchanges and Physiological Responses. Physiol.
  Rev., 27:200-227.
MacLeod, J., and R.S. Hotchkiss. 1941. The Effect of
  Hyperpyrexia Upon Spermatozoa Counts in Men.
  Endocrinology, 28:780-784.
Magin, R.L, S. Lu, and  S.M. Michaelson. 1977a.
  Stimulation of Dog Thyroid by Local Application of
  High Intensity Microwaves.  Am. J. Physiol.,
  233.E363-E368.
Magin, R.L, S. Lu, and  S.M. Michaelson. 1977b.
  Microwave  Heating Effect on the Dog Thyroid
  Gland.  IEEE Trans. Biomed. Eng.,  BME-24:522-
  529.
Majewska, K. 1968. Investigations on the Effect of
  Microwaves on the Eye. Pol. Med. J., 7:989-994.
Manikowska, E.,  J.M. Luciani, B.  Servantie,  P.
  Czerski, J. Obrenovitch, and A. Stahl. 1979. Effects
  of 9.4 GHz Microwave Exposure  on Meiosis in
  Mice. Experientia, 35:388-390.
Manikowska-Czerska, E., P. Czerski, and W.M. Leach.
  1983a.  Effects  of 0.915  and  9.4 GHz CW
  Microwaves on Meiosis in Male Mice. Abstract #H-
  3, 5th Annual Scientific Session, The Bioelectro-
  magnetics Society, 12-17 July, Boulder, Colorado
  (available from The Bioelectromagnetics Society,
  One Bank Street,  Suite  307, Gaithersburg,
  Maryland 20878). p. 57.
Manikowska-Czerska, E., P. Czerski, and W.M. Leach.
  1983b. Dominant Lethal Testing after Exposure of
  Mice to 0.915 GHz Microwaves. Abstract #H-4,5th
  Annual Scientific Session, The Bioelectromagnetics
  Society, 12-17 July, Boulder, Colorado (available
  from The  Bioelectromagnetics Society, One Bank
  Street, Suite 307, Gaithersburg, Maryland 20878).
  p. 57.
Marmor, J.B., N. Hahn, and G.M. Hahn. 1977. Tumor
  Cure and Cell Survival after Localized Radiofre-
  quency Heating. Cancer Res., 37:879-883.
Mathur, D.S., M.A. Aman, and K.R.  Sarkar. 1980.
  Induction  of Maternal Haploids in Maize Through
  Heat Treatment of  Pollen. Current Sci., 49:744-
  746.
Mayers, C.F., and J.A. Habeshaw. 1973. Depression
  of Phagocytosis: A Non-Thermal Effect of Micro-
  wave Radiation as a Potential Hazard to Health. Int.
  J. Radiat.  Biol., 24:449-461
McAfee, R.D., R. Braus, Jr., and J. Fleming, Jr. 1973.
  The Effect of 2450 MHz Microwave Irradiation on
  the Growth of Mice. J. Microwave Power, 8:111-
  116.
McAfee,  R.D., A.  Longacre, Jr., R.R. Bishop, ST.
  Elder, J.G. May, M.G. Holland,  and R. Gordon.
  1979. Absence of Ocular Pathology after Repeated
  Exposure of Unanesthetized Monkeys to 9.3 GHz
  Microwaves. J. Microwave Power, 14:41-44.

McLaughlin, J.R. 1953. A Survey of Possible Health
  Hazards from Exposure to Microwave  Radiation.
  Hughes Aircraft Corp., Culver City, California.

McLees, B.D., E.D. Finch, and M.L Albright. 1972. An
  Examination of Regenerating  Hepatic Tissue
  Subjected to Radio-Frequency Irradiation. J. Appl.
  Physiol., 32:78-85.
McNiven, D.R., and  D.J.  Wyner. 1976. Microwave
  Therapy and Muscle  Blood  Flow in Man. J.
  Microwave Power, 11:168-170.
McRee,  D.I., and P.E.  Hamrick. 1977. Exposure of
  Japanese Quail  Embryos to 2.45-GHz Microwave
  Radiation During Development. Radiat. Res.,
  71:355-366.
McRee,  D.I., G.K.  Livingston,  and G. MacNichols.
  1978. Incidence of Sister Chromatid Exchange in
  Bone  Marrow Cells of the Mouse  Following
  Microwave Exposure (Abstract). In: Symposium on
  Electromagnetic Fields  in Biological  Systems,
  IEEE/IMP), Ottawa, Canada, pp. 15-16.

McRee, D.I., R. Faith, E.E. McConnell, and A.W. Guy.
  1980a. Long-Term  2450-MHz  CW Microwave
  Irradiation  of Rabbits: Evaluation of Hematological
  and Immunological Effects. J. Microwave Power,
  15:45-52.
McRee, D.I.,  M. Galvin, C. Hall, and M. Lieberman.
  1980b. Microwave Effects on Embryonic Cardiac
  Tissue of Japanese Quail. ln:OndesElectromagne-
  tiques et Biologie, A.J. Berteaud and B. Servantie,
  eds. Paris, France, pp. 79-84.
McRee,  D.I., G. MacNichols, and  G.K. Livingston.
  1981. Incidence  of Sister Chromatid Exchange in
  Bone  Marrow Cells of the Mouse  Following
  Microwave Exposure. Radiat. Res., 85:340-348.

Merritt, J.H., R.H. Hartzell, and J.W. Frazer. 1976. The
  Effect of 1.6 GHz Radiation on Neurotransmitters in
  Discrete Areas  of the Rat Brain. In: Biological
  Effects of Electromagnetic  Waves, Vol. I, C.C.
  Johnson and M.L. Shore, eds. HEW Publication
  (FDA) 77-8010, Rockville, Maryland, pp. 290-298.

Merritt, J.H., A.F. Chamness, R.H. Hartzell, and S.J.
  Allen. 1977. Orientation Effects on Microwave-
  Induced Hyperthermia and Neurochemical Corre-
  lates. J. Microwave Power, 12:167-172.
                      R-18

-------
Merritt, J.H., A.F. Chamness, and S.J. Allen.  1978.
  Studies on Blood-Brain Barrier Permeability after
  Microwave-Radiation. Radiat. Environ.  Biophys.,
  15:367-377.
Merritt, J.H.,  W.W.  Shelton, and A.F.  Chamness.
  1982. Attempts to Alter  45Ca2+ Binding to Brain
  Tissue with Pulse-Modulated Microwave Energy.
  Bioelectromagnetics, 3:475-478.
Michaelson, S.M., and H.P. Schwan. 1973. Compara-
  tive Aspects of Radiofrequency and  Microwave
  Biomedical Research. IEEE Int. Microwave Symp.,
  Boulder, Colorado, pp. 330-332.

Michaelson,  S.M., R.A.E. Thomson, and J.W.
  Howland. 1961. Physiologic Aspects of Microwave
  Irradiation of Mammals. Am. J. Physiol., 201:351-
  356.
Michaelson, S.M., R.A.E. Thomson, LT. Odland, and
  J.W. Howland. 1963. The  Influence of Microwaves
  on Ionizing  Radiation Exposure. Aerospace  Med.,
  34:111-115.
Michaelson, S.M., R.A.E. Thompson, M.Y. EITamami,
  H.S. Seth,  and  J. W.  Howland. 1964. The
  Hematologic Effects of  Microwave Exposure.
  Aerospace Med., 35:824-829.
Michaelson, S.M., R. Guillet, and  F.W. Heggeness.
  1978.  Influence  of Microwave Exposure on
  Functional Maturation of the Rat. In: Developmental
  Toxicology  of Energy-Related  Pollutants, D.D.
  Mahlum, M.R.  Sikov, P.L.  Hackett, and F.D.
  Andrew, eds. DOE Symposium Series 47, Wash-
  ington, D.C. pp. 300-316.

Mickey, G.H., and L Koerting. 1970. Chromosome
  Breakage  in Cultured Chinese Hamster  Cells
  Induced  by Radiofrequency Treatment.  Environ.
  Mutagen Soc., 3:25-26.

Mikolajczyk, H. 1976. Microwave-Induced  Shifts of
  Gonadotropic Activity in Anterior Pituitary Gland of
  Rats.  In:  Biological  Effects of  Electromagnetic
  Waves, C.C.  Johnson and M.L Shore, eds. U.S.
  DHEW  (FDA) 77-8010, Rockville, Maryland,  pp.
  377-383.

Milham, S., Jr.  1982.  Mortality from Leukemia  in
  Workers Exposed to Electrical and Magnetic Fields.
  New England J. Med., 307:249.

Milroy, W.C., and S.M. Michaelson. 1972. Thyroid
  Pathophysiology of Microwave  Radiation. Aero-
  space Med., 43:1126-1131.

Miro, L., R. Loubiere, and A  Pfister. 1974. Effects of
  Microwaves on the Cell Metabolism of theReticulo-
  Histocytic System. In: Biological Effects and Health
  Hazards of  Microwave Radiation, P. Czerski,  K.
  Ostrowski, M.L. Shore, C. Silverman, M.T. Suess,
  and B. Waldeskog, eds. Polish Medical Publication,
  Warsaw, Poland, pp. 89-97.
Mitchell,  D.S., W.G.  Switzer,  and E.L Bronaugh.
  1977. Hyperactivity and Disruption of Operant
  Behavior  in Rats  after Multiple Exposure  to
  Microwave Radiation. Radio Sci., 12(6S):263-271.

Mittler, S. 1976. Failure of 2- and 10-Meter Radio
  Waves  to  Induce Genetic Damage in Drosophila
  melanogaster. Environ. Res., 11:326-330.

Mittler, S.  1977. Failure of  Chronic Exposure  to
  Nonthermal FM Radio Waves to MulaleDrosophila.
  J. Heredity, 68:257-258.

Mittler, S.  1979.  Hyperthermia and Radiation-
  Induced Dominant Lethals and Chromosome Loss
  in  Female Drosophila melanogaster. J.  Heredity,
  70:81-82.

Moe, K.E., R.H. Lovely, D.E. Meyers, and A.W. Guy.
  1976. Physiological and Behavioral Effects  of
  Chronic Low Level Microwave Radiation in Rats. In:
  Biological Effects of Electromagnetic Waves, Vol. I.,
  C.C. Johnson and M.L Shore, eds. HEW Publication
  (FDA) 77-8010, Rockville, Maryland, pp. 248-256.

Monahan, J.C.,  and  H.S. Ho. 1976. Microwave
  Induced Avoidance  Behavior in the Mouse. In:
  Biqlogical Effects of Electromagnetic Waves, Vol. I.,
  C.C. Johnson and M.L Shore, eds. HEW Publication
  (FDA) 77-8010, Rockville, Maryland, pp. 274-283.

Monahan, J.C., and H.S. Ho.  1977. The  Effect  of
  Ambient Temperature  on the  Reduction  of
  Microwave Energy Absorption by Mice. Radio Sci.,
  12(6S):257-262.

Monahan, J.C., and W.W. Henton. 1979. The Effect of
  Psychoactive Drugs  on Operant Behavior Induced
  by Microwave Radiation. Radio Sci., 14(6S):233-
  238.
Monjan,  A.A.  1981. Stress and Immunologic
  Competence: Studies in Animals. In: Psychoneuro-
  immunology, R. Ader, ed. Academic Press,  New
  York, New York. pp.  185-228.
Montgomery,  L.D. 1972.  A  Simulation  of Heat
  Transfer in  Man Under Immersed Conditions.
  Doctoral Thesis, UCLA, Los Angeles, California.
Montgomery, L.D. 1974a. A Model of Heat Transfer in
  Immersed Man. Ann. Biomed. Eng., 2:19-46.
Montgomery,  L.D. 1974b. Analytic Model for
  Assessing the Thermal  Performance  of  Scuba
  Divers. J. Hydronautics, 8:108-115.

Montgomery, L.D. 1975. Biothermal  Simulation  of
  Scuba  Divers. Aviat.  Space  Environ.  Med.,
  46(6):814-818.
Moore, H.A., R. Raymond, M. Fox,  and A.G. Galsky.
  1979. Low-Intensity Microwave Radiation and the
  Virulence  of Agrobacterium tumefaciens  Strain
  B6. Appl. Environ. Microbiol., 37:127-130.
                                                                   R-19

-------
Morishima, H.O., B. Glaser, W.H. Niemann, and L.S.
  James. 1975. Increased Uterine Activity and Fetal
  Deterioration During Maternal Hyperthermia. Am. J.
  Obstet. Gyn., 121:531-538.
Morris, W., ed. 1976. The American  Heritage
  Dictionary of the English  Language. Houghton
  Mifflin Company, Boston, Massachusetts. 1550pp.
Muller,  H.J., and E. Altenburg. 1919. The Rate of
  Change of Hereditary Factors in Drosophila.  Proc.
  Soc. Exp. Biol. Med., 17:10-14.
Mumford, W.W. 1961. Some Technical Aspects of
  Microwave Radiation Hazards. Proc. IRE, 49:427-
  447.
Mumford, W.W. 1969. Heat  Stress  Due to  RF
  Radiation. Proc. IEEE, 57:171-178.
Muraca, G.J., E.S. Ferri, and F.L Buchta. 1976. A
  Study of the Effects of Microwave Irradiation of the
  RatTestes. In: Biological Effects of Electromagnetic
  Waves, Vol. I., C.C. Johnson and M.L Shore, eds.
  DHEW Publication (FDA) 77-8010, Rockville,
  Maryland, pp. 484-494.
Myers, R.D., and D.H.  Ross. 1981.  Radiation and
  Brain Calcium: A  Review and Critique.  Neurosci.
  Biobehav. Rev., 5:503-543.

NCRP. 1981. Radiofrequency Electromagnetic Fields,
  Properties, Quantities and  Units, Biophysical
  Interaction and Measurements. NCRP Report No.
  67, March 1, 1981. National Council on  Radiation
  Protection and Measurements  Publications,
  Washington,  D.C. 134pp.
Nelson,  A.J.M., and J.A.G.  Holt.  1978. Combined
  Microwave Therapy. Med. J. Aust.,  2:88-90.
Nieset, R.T., R. Baus,  Jr., R.D. McAfee, J.J. Friedman,
  A.S. Hyde, and J.D. Fleming, Jr. 1958. Review of
  the Work Conducted at Tulane University. In: Proc.
  Second Tri-Service Conference  on Biological
  Effects of Microwave Energy. (NTIS AD 131 477).
  pp. 202-214.
O'Connor, M.E. 1980. Mammalian Teratogenesis and
  Radio-Frequency Fields. Proc. IEEE., 68:56-60.
Odland, L.T. 1973. Radio-Frequency Energy:  A
  Hazard to Workers? Ind. Med. Surg. 42:23-26.
Ogilvie, D.M., and R.H. Stinson. 1966. The Effect of
  Age on Temperature Se!3Ction by Laboratory Mice
  (Mus musculus). Can. J. Zool., 44:511-517.
Olcerst, R.B., S. Belman, M. Eisenbud, W.W.
  Mumford, and J.R. Rabinowitz. 1980. The Increased
  Passive Efflux of Sodium and Rubidium from Rabbit
  Erythrocytes by Microwave Radiation. Radial. Res.,
  82:244-256.

Oldendorf, W.H. 1970. Measurement of Brain Uptake
  of Radiolabeled Substances Using a Tritiated Water
  Internal Standard.  Brain Res., 24:372-376.
                       R-20
Oliva, S.A., and G.N. Catravas.  1977. A Multiple-
  Animal Array for Equal Power Density Microwave
  Irradiation. IEEE Trans. Microwave Theory Tech-
  niques, MTT-25(5):433-436.
Olsen, R.G., and W.C. Hammer.  1980. Microwave-
  Induced Pressure  Waves  in a Model  of Muscle
  Tissue. Bioelectromagnetics, 1:45-54.
Olsen,  R.G., J.L  Lords,  and C.H.  Durney.  1977.
  Microwave-Induced Chronotropic Effects in the
  Isolated Rat Heart. Ann. Biomed. Eng., 5:395-409.
Oscar,  K.J., and T.D. Hawkins.  1977. Microwave
  Alteration of the Blood-Brain  Barrier  System of
  Rats. Brain Res., 126:281-293.
Oscar,  K.J., S.P.  Gruenau,  M.T. Folker, and S.I.
  Rapoport. 1981. Local Cerebral Blood  Flow after
  Microwave Exposure.  Brain Res., 204:220-225.
Paff, G.H., R.J. Boucek, R.E. Nieman, and  W.B.
  Deichmann. 1963. The Embryonic Heart Subjected
  to Radar. Anat. Rec., 147:379-385.
Palmbald, J. 1981. Stress and Immunologic Com-
  petence:  Studies in Man. In: Psychoneuroimmu-
  nology, R. Ader, ed. Academic Press, New York. pp.
  228-257.
Parker, L.N. 1973. Thyroid Suppression and Adreno-
  medullary Activation by Low-Intensity Microwave
  Radiation. Am. J. Physiol., 224:1388-1390.
Paulsson, L-E., Y. Hamnerius, and W.G. McLean.
  1977.  The  Effects of Microwave Radiation on
  Microtubules and Axonal Transport. Radiat. Res.,
  70:212-223.
Pavlov, I.P. 1960. Conditioned Reflexes (Translated by
  G.V. Anrep). Dover Publications, New  York, New
  York. 430 pp.
Pay, T.L, E.C. Beyer and C.F. Reichelderfer. 1972.
  Microwave Effects on Reproductive Capacity and
  Genetic Transmission in Drosophila melanogaster.
  J. Microwave Power, 7:75-82.
Pay, T.L, F.A. Andersen, and G.L. Jessup, Jr. 1978. A
  Comparative Study of the Effects of Microwave
  Radiation and  Conventional Heating on the
  Reproductive Capacity of Drosophila melanogaster.
  Radiat. Res., 76:271-282.
Pazderova,  J., R. Fisher, J.  Formanek, J. John, E.
  Lucas, and V. Styblova 1969.  Health State of
  Workers  Exposed  to  Long-Term Electromagnetic
  Radiation of Order of Meter Waves. (In  Czech.)
  Pracov. Lek., 21:346-361.

Pazderova-Vejlupkova, J., and  M.  Josifko.  1979.
  Changes in  the Blood Count of Growing Rats
  Irradiated with  a  Microwave  Pulse Field. Arch.
  Environ. Health, 34:44-50.
Pellon, J.R., K.M. Ulmer, and R.F. Gomez. 1980. Heat
  Damage to the Folded Chromosome of Escherichia
  coliK-12. Appl. Environ. Microbiol., 40(2):358-364.

-------
Pennes, H.H.  1948. Analysis of Tissue and Arterial
  Blood Temperatures in the Resting  Human
  Forearm. J. Appl. Physiol., 1:93-122.
Peterson, D.J., LM. Partlow, and O.P. Gandhi. 1979.
  An  Investigation of the  Thermal and Athermal
  Effects of Microwave Irradiation on Erythrocytes.
  IEEE Trans. Biomed. Eng., BME-26:428-436.

Peto, R., M.C. Pike, P. Armitage, N.E. Breslow, D.R.
  Cox, S.V. Howard,  N. Mantel, K.  McPherson, J.
  Peto, and P.G. Smith. 1976. Design and Analysis of
  Randomised Clinical Trails Requiring Prolonged
  Observation of Each Patient. I.  Introduction and
  Design. Br. J. Cancer, 34:585-612.

Peto, R., M.C. Pike, P. Armitage, N.E. Breslow, D.R.
  Cox, S.V. Howard,  N. Mantel, K.  McPherson, J.
  Peto, and P.G. Smith. 1977. Design and Analysis of
  Randomized Clinical Trials Requiring Prolonged
  Observation of Each Patient. II. Analysis and
  Examples. Br. J. Cancer, 35:1-39.

Phillips, R.D., N.W.  King, and E.L Hunt.  1973.
  Thermoregulatory, Cardiovascular and Metabolic
  Response of Rats to Single or Repeated Exposures
  to 2450 MHz Microwaves. 1973 Microwave Power
  Symp., Int. Microwave  Power  Inst, Edmonton,
  Canada, pp. 3B35/1-3B35/4.

Phillips, R.D.,  E.L. Hunt, and N.W. King. 1975a. Field
  Measurements, Absorbed Dose, and Biologic
  Dosimetry of Microwaves. Ann. N.Y. Acad. Sci.,
  247:499-509.

Phillips, R.D.,  E.L. Hunt, R.D. Castro, and N.W. King.
  1975b. Thermoregulatory, Metabolic,  and Cardio-
  vascular Response of Rats to Microwaves. J. Appl.
  Physiol., 38:630-635.
Pickard, W.F., and Y.H. Barsoum. 1981.  Radio-
  Frequency  Bioeffects at the Membrane Level:
  Separation of Thermal and Athermal Contributions
  in the Characeae. J. Membrane Biol. 61:39-54.

Pickard, W.F., and F.J. Rosenbaum. 1978. Biological
  Effects of Microwaves at the Membrane Level: Two
  Possible Athermal  Electrophysiological Mecha-
  nisms and  a Proposed Experimental  Test.  Math.
  Biosci., 39:235-253.
Pollack, H. 1979. Epidemiologic Data on American
  Personnel in the Moscow Embassy. Bull. N.Y. Acad.
  Med., 55:1182-1186.
Pollak, M. 1965.  On the  Dielectric Dispersion of
  Polyelectrolytes with Application to DNA. J. Chem.
  Phys., 43:908-909.

Prausnitz,  S., and C. Siisskind. 1962. Effects of
  Chronic Microwave  Irradiation on Mice. IRE Trans.
  Biomed. Electron., 9:104-108.
Preskorn, S.H., W.D.  Edwards, and  D.R. Justesen.
  1978. Retarded Tumor Growth and Greater
  Longevity in Mice after Fetal Irradiation by 2450-
  MHz Microwaves. J. Surg. Oncol., 10:483-492.
Presman, A.S., and N.A. Levitina. 1962. Nonthermal
  Action of Microwaves  on  Cardiac  Rhythm.
  Communication I. A Study  of  the Action of
  Continuous Microwaves.  Bull. Exp. Biol.  Med.,
  53:36-39.
Preston, E., and G. Prefontaine.  1980. Cerebrovas-
  cular Permeability to Sucrose in the Rat Exposed to
  2,450-MHz Microwaves. J. Appl. Physiol., 49:218-
  223.
Preston, E.,  E.J. Vavasour,  and H.M. Assenheim.
  1979. Permeability of the Blood-Brain Barrier to
  Mannitol in the Rat Following 2450 MHz Microwave
  Irradiation. Brain Res., 174:109-117.
Prince, J.E., L.H. Mori, J.W. Frazer, and J.C. Mitchell.
  1972. Cytologic Aspect of RF Radiation in the
  Monkey. Aerospace Med,, 43:759-761.
Prohofsky, E.W.,  K.C. Lu,  LL Van Zandt, and B.F.
  Putnam.  1979. Breathing Modes and Induced
  Resonant Melting of the Double Helix. Phys.  Lett.,
  70A:492-494.

Pucak, G.J., C. S. Lee, and A.S.Zaino. 1977. Effects of
 'Prolonged High Temperature  on Testicular
  Development and Fertility in the  Male  Rat. Lab.
  Anim. Sci., 27(1 ):76-77.

Puska, P., J. Tuomilehto, J. Salonen, A. Nissinen, J.
  Virtamo, S. Bjbrkqvist, K. Koskela, L. Neittaanmaki,
  T.  Takalo, T.  Kottke,  J. Ma'ki, P. Sipila, and P.
  Varvikko.  1978. The  North Karelia  Project:
  Evaluation of  a Comprehensive  Community
  Programme for Control of Caradiovascular Diseases
  in 1972-1977 in North Karelia, Finland.  Research
  Institute  for Community  Health, University of
  Kuopio, Kuopio, Finland. 449 pp.

Rabinowitz, J.R.  1973. Possible Mechanisms for the
  Biomolecular Absorption of Microwave  Radiation
  with Functional Implications.  IEEE Trans. Micro-
  wave Theory Techniques, MTT-21:850-851.
Rama Rao, G., C. A. Cain, J. Lockwood, and W.A.F.
  Tompkins. 1983. Effects of Microwave Exposure on
  the Hamster Immune System. II. Peritoneal
  Macrophage  Function.  Bioelectromagnetics,
  4:141-155.
Ramirez, E., J.L. Monteagudo, M.  Garcia-Garcia, and
  J.M.R. Delgado. 1983. Oviposition and Develop-
  ment of Drosophila Modified by Magnetic Fields.
  Bioelectromagnetics, 4:315-326.

Rand, R.P., A.C. Burton, and T. Ing. 1965. The Tail of
  the Rat, in Temperature Regulation and Acclimatiza-
  tion. Can. J. Physiol. Pharmacol., 43:257-267.

Rapoport, S.I. 1976. Blood-Brain Barrier in Physiology
  and Medicine. Raven Press, New York, New  York.
  316 pp.
                                                                    R-21

-------
Reaves, T.A., Jr.  1977. Gain of Thermosensitive
  Neurons in the Preoptic  Area of the Rabbit,
  Oryctolagus cuniculus. J. Thermal Biol., 2:31-33.
Reno, V.R. 1974. Microwave Reflection, Diffraction,
  and Transmission Studies of Man. Report NAMRL-
  1199. Naval Aerospace Medical Research Labora-
  tory,. Pensacola, Florida. 39  pp.
Richardson, A.W., T.D. Duane, and H.M. Mines. 1948.
  Experimental Lenticular Opacities Produced by
  Microwave Irradiations. Arch. Phys. Med., 29:765-
  769.
Riddle, M.M., R. J.  Smialowicz, and R.R. Rogers.
  1982. Microwave Radiation (2450-MHz) Potentiates
  the Lethal  Effect  of  Endotoxin in  Mice.  Health
  Phys., 42:335-340.
Riley, V. 1981. Psychoneuroendocrine Influences on
  Immunocompetence and  Neoplasia.  Science,
  212:1100-1109.
Rissmann, W.J., and C.A.  Cain. 1975. Microwave
  Hearing in Mammals. Proc. Natl. Elec.  Cong.,
  30:239-244.
Roberti, B., G.H. Heebels, J.C.  M. Hendricx, A.H.A.M.
  de Greet  and O.L Wolthuis.  1975.  Preliminary
  Investigations of the Effects of Low-Level Micro-
  wave Radiation and Spontaneous Motor Activity in
  Rats. Ann. N.Y. Acad. Sci., 247:417-424.
Roberts,  N.J., Jr. 1979.  Temperature and Host
  Defense. Microbiol. Rev.,43:241-259.
Roberts,  N.J.,  Jr., and R.T.  Steigbigel.  1977.
  Hyperthermia and  Human  Leukocyte Functions:
  Effects on Response of Lymphocytes to Mitogen
  and Antigen and Bactericidal Capacity of Monocytes
  and Neutrophils. Infect. Immun., 18:673-679.

Robinette, C.D.  and C. Silverman. 1977. Causes of
  Death Following Occupational  Exposure to
  Microwave Radiation (Radar) 1950-1974. In:
  Symposium on  Biological Effects and Measure-
  ments of Radiofrequency/Microwaves, D.G.
  Hazzard, ed.  HEW Publication (FDA) 77-8026,
  Rockville, Maryland, pp. 338-344.
Robinette, C.D., C. Silverman, and S. Jablon. 1980.
  Effects upon Health of Occupational Exposure to
  Microwave Radiation  (Radar). Am. J. Epidemiol.,
  112:39-53.

Rogers,  P., and  A. Matossian-Rogers.  1982.
  Differential Sensitivity of  Lymphocyte Subsets to
  Corticosteroid Treatment. Immunology, 46:841-
  848.
Rosenthal, S.W.,  L. Birenbaum, I.T. Kaplan, W.
  Metlay, W.Z.  Snyder, and M.M. Zaret.  1976.
  Effects of 35 and 107 GHz CW Microwaves on the
  Rabbit Eye. In: Biological Effects of Electromagnetic
  Waves, Vol. I, C.C. Johnson and M.L. Shore, eds.
  HEW Publication (FDA) 77-8010,  Rockville,
  Maryland, pp. 110-128.
Roszkowski, W., J.K. Wrembel, K. Roszkowski, M.
  Janiak, and S. Szmigielski. 1980. The Search for an
  Influence of Whole-Body Microwave Hyperthermia
  on Anti-Tumor  Immunity. J. Cancer Res.  Clin.
  Oncol., 96:311-317.
Rotkovska,  D.,  and A. Vacek. 1972. Effect of High-
  Frequency Electromagnetic  Field Upon Haemo-
  poietic Stem Cells in Mice. Folia Biologica (Praha),
  18:292-297.
Rotkovska,  D.,  and A. Vacek.  1975. The Effect of
  Electromagnetic Radiation on the Hematopoietic
  Stem Cells of Mice. Ann. N.Y. Acad. Sci., 247:243-
  250.
Rotkovska,  D.,  and A. Vacek. 1977. Modification of
  Repair of X-lrradiation Damage of Hemopoietic
  System  of Mice by Microwaves. J. Microwave
  Power, 12:119-123.
Rozzell, T.C., C.C. Johnson, C.H. Durney, J.L Lords,
  and R.G.  Olsen. 1974. A Nonperturbing Tempera-
  ture Sensor for Measurements in Electromagnetic
  Fields. J. Microwave Power,  9(3):241 -249.

Rudnev, M., A. Bokina, N. Eksler, and M. Navakatikyan.
  1978. The Use of Evoked Potential and Behavioral
  Measures in the Assessment of Environmental
  Insult. In: Multidisciplinary Perspectives in Event-
  Related Brain Potential Research, D.A. Otto, ed.
  EPA-600/9-77-043, U.S. Environmental Protection
  Agency, Research Triangle Park, North Carolina.
  pp. 444-447.
Rudge, A.W. 1970.  An Electromagnetic  Radiation
  Probe for Near-Field Measurements at Microwave
  Frequencies. J. Microwave Power, 5(3):155-174.

Ruggera, P.S. 1976. E- and H-Field Instrumentation
  and  Calibration below 500 MHz.  In:  Biological
  Effects of Electromagnetic  Waves.  Vol.  II, C.C.
  Johnson  and M.L. Shore, eds.  HEW Publication
  (FDA) 77-8011, Rockville, Maryland, pp. 281-296.

Rugh, R. 1976a. Are Mouse Fetuses Which Survive
  Microwave  Radiation Permanently  Affected
  Thereby? Health Phys.,  31:33-39.

Rugh, R. 1976b. The Relation of Sex, Age, and Weight
  of  Mice to Microwave Radiation Sensitivity. J.
  Microwave Power, 11 (2): 127-132.
Rugh, R., and M. McManaway.  1976. Anesthesia as
  an Effective Agent Against the Production of
  Congenital Anomalies in Mouse Fetuses Exposed
  to  Electromagnetic Radiation. J. Exp. Zool.,
  197:363-368.
Rugh, R., and M. McManaway. 1977. Mouse Fetal
  Sensitivity to Microwave Radiation. Cong. Anom.,
  17:39-45.
                      R-22

-------
Rugh, R., E.I. Ginns, H.S. Ho, and W.M. Leach. 1974.
  Are Microwaves Teratogenic? In: Biologic Effects
  and Health Hazards of Microwave Radiation, P.
  Czerski, K. Ostrowski, M.L. Shore, C. Silverman,
  M.J. Suess, and B. Waldeskog, eds. Polish Medical
  Publishers, Warsaw, Poland, pp. 98-107.
Rugh,.R., E.I. Ginns, H.S. Ho, and W.M. Leach. 1975.
  Responses of the Mouse to Microwave  Radiation
  During Estrous Cycle and Pregnancy. Radiat. Res.,
  62:225-241.
Rukspollmuang, S., and K.M. Chen. 1979. Heating of
  Spherical Versus Realistic Models of Human and
  Infrahuman  Heads by  Electromagnetic Waves.
  Radio Sci., 14(6S):51-62.
Ruppin, R. 1979. Electromagnetic Power Deposition
  in a Dielectric Cylinder  in the  Presence of a
  Reflecting Surface. IEEE Trans. Microwave Theory
  Techniques, MTT-27(11):910-914.

Sagan, P.M., and R.G. Medici. 1979. Behavior of
  Chicks Exposed to Low-Power 450-MHz  Fields
  Sinusoidally Modulated at EEG Frequencies. Radio
  Sci.. 14(6S):239-245.

Sams & Co.  1981. Reference  Data for Radio
  Engineers. Howard W. Sams & Co., Inc., New York,
  New York. Chapter 1.

Sanders, A.P., D.J. Schaefer, and W.T. Joines. 1980.
  Microwave Effects on Energy Metabolism of Rat
  Brain. Bioelectromagnetics, 1:171-181.
Sanza, J.N., and J. de Lorge.  1977. Fixed Interval
  Behavior of Rats Exposed to Microwaves at Low
  Power Densities. Radio Sci., 12(6S):273-277.
Satinoff, E., and R. Hendersen. 1977. Thermoregula-
  tory Behavior. In:  Handbook of Operant Behavior,
  W.K. Honig and J.E. Staddon, eds. Prentice Hall,
  Engiewood Cliffs,  New Jersey, pp.  153-173.

Saunders, R.D., S.C. Darby, and C.I. Kowalczuk. 1983.
  Dominant Lethal  Studies in Male Mice after
  Exposure to 2.45 GHz Microwave Radiation. Mutat.
  Res., 117:345-356.
Schlagel, C.J.,  K. Sulek,  H.S. Ho, W.M. Leach, A.
  Ahmed, and J.N. Woody. 1980. Biological  Effects of
  Microwave Exposure. II. Studies on the Mechanisms
  Controlling Susceptibility to Microwave-Induced
  Increases in Complement Receptor-Positive
  Spleen Cells. Bioelectromagnetics, 1:405-414.

Schmidt-Nielsen, K. 1964. Desert Animals. Physio-
  logical Problems  of Heat and Water. Oxford
  University Press, New Y.ork. 277 pp.

Schmidt-Nielsen, K.  1972. How Animals Work.
  Cambridge University Press, New York, New York.
  114pp.

Schmidt-Nielsen, K.  1975. Scaling in Biology: The
  Consequences of Size. J. Exp. Zool., 194:287-308.
 Schmidt-Nielsen, K. 1979. Animal Physiology:
  Adaptation and Environment, 2nd ed. Cambridge
  University Press, Cambridge, Massachusetts. 560
  PP-
 Schmidt-Nielsen, K.,  B. Schmidt-Nielsen, S.A.
  Jarnum, andT.R. Houpt. 1957. Body Temperature of
  the Camel and Its Relation to Water Economy. Am.
  J. Physiol., 188:103-112.
 Scholander, P.F., R. Hock, V. Walters, F. Johnson, and
  C. Irving. 1950. Heat Regulation in Some Arctic and
  Tropical Mammals and Birds. Biol. Bull., 99:237-
  258.
 Scholl, D.M., and S.J. Allen. 1979. Skilled Visual-
  Motor  Performance  by Monkeys in a  1.2-GHz
  Microwave Field. Radio Sci., 14(6S):247-252.

 Schrot, J., J.R. Thomas, and  R.A. Banvard. 1980.
  Modification  of the Repeated Acquisition of
  Response  Sequences in Rats by Low-Level
  Microwave Exposure. Bioelectromagnetics, 1:89-
  99.

 Schwan, H.P.  1957.  Electrical Properties of Tissue
  and Cell Suspensions. Adv. Biol.  Med. Phys.,
  5:147-209.

 Schwan, H.P.  1965.  Electrical Properties of Bound
  Water. Ann. N.Y. Acad. Sci., 125:344-354.
 Schwan,  H.P., and  K.R. Foster.  1980.  RF-Field
  Interactions with Biological Systems: Electrical
  Properties and Biophysical Mechanisms. Proc.
  IEEE, 68:104-113.
 Schwan, H.P., and K.Li. 1956. Hazards Due to Total
  Body Irradiation by Radar.  Proc, IRE,  44:1572-
  1581.
 Schwan, H.P., A. Anne, and L. Sher. 1966. Heating of
  Living  Tissues.  Report  NAEC-ACEL-534, U.S.
  Naval  Air Engineering Center, Philadelphia,
  Pennsylvania. 30 pp.
 Schwarz, G. 1967. On Dielectric Relaxation Due to
  Chemical  Rate Processes. J. Phys. Chem.,
  71:4021-4030.
 Schwa rz, G. 1972. Dielectric Relaxation of Biopoly-
  mers in Solution. Adv. Mol. Relaxation Processes,
  3:281-295.
 Schwa rz, G., and J. Seelig. 1968. Kinetic Properties
  and the  Electric  Field  Effect of the Helix-Coil
  Transition of Poly (/-Benzyl L-Glutamate) Deter-
  mined from Dielectric Relaxation Measurements.
  Biopolymers, 6:1263-1277.

Scientific American. 1979. The Brain. Sci. Am.,
  241(3):1-252.

Seaman, R.L, and H. Wachtel. 1978. Slow and Rapid
  Responses to CW and Pulsed Microwave Radiation
  by Individual Aplysia Pacemakers. J. Microwave
  Power,  13:77-86.
                                                                   R-23

-------
Segal, A.S., and R.L Magin. 1982. Microwaves and
  the Blood-Brain Barrier: A Review. J. Bioelectricity,
  1:351-398.
Servantie, B., G. Bertharion, R. Joly, A. Servantie, J.
  Etienne, P.  Dreyfus,  and  P. Escoubet. 1974.
  Pharmacologic  Effects of a Pulsed Microwave
  Field.  In: Biologic Effects  and Health Hazards of
  Microwave Radiation,  P.  Czerski, K. Ostrowski,
  M.L.  Shore,  C.  Silverman,  M.J. Suess, and  B.
  Waldeskog,  eds. Polish Medical  Publishers,
  Warsaw, Poland, pp. 36-45.
Shacklett, D.E., T.J. Tredici, and D.L Epstein. 1975.
  Evalution  of Possible Microwave-Induced Lens
  Changes in the  United States Air Force. Aviat.
  Space Environ. Med., 46:1403-1406.
Shah, S.A., and J.A. Dickson, 1978a. Effect  of
  Hyperthermia on the Immune Response of Normal
  Rabbits. Cancer Res., 38:3518-3522.
Shah, S.A., and J.A. Dickson. 1978b. Effect  of
  Hyperthermia on the Immuno-competence of VX2
  Tumor-Bearing Rabbits.  Cancer  Res., 38:3523-
  3531.
Shandala, M.G., M.I. Rudnev, and M.A. Navakatian.
  1977. Patterns of Change in  Behavioral Reactions
  to Low Power Densities of Microwaves (Abstract).
  International Symposium on  the Biological Effects
  of Electromagnetic Waves (URSI), Airlie, Virginia.
  p. 88.
Shelton, W.W., Jr., and J.H. Merritt. 1981. In Vitro
  Study of Microwave Effects  on Calcium Efflux in
  Rat Brain Tissue. Bioelectromagnetics, 2:161 -167.
Sheppard, A.R., S.M. Bawin, and W.R. Adey. 1979.
  Models of Long-Range Order in Cerebral Macro-
  molecules: Effect of Sub-ELF  and of Modulated
  VHF and UHF Fields. Radio Sci., 14(6S):141-145.
Sher, L.D. 1970. Interaction of Microwave and RF
  Energy on Biological Material. In: Electronic
  Product Radiation and the Health Physicist. Report
  BRH/DEP  70-26, HEW,  Bureau of Radiological
  Health, Rockville, Maryland, pp. 431-462.
Sheridan,  J.P., B.P. Gaber,  F. Cavatorta, and P.E.
  Schoen. 1979. Molecular  Level Effects of Micro-
  waves on Natural  and  Model Membranes:  A
  Raman  Spectroscopic Investigation (Abstract).
  Joint Meeting of  USNC/URSI and the Bioelectro-
  magnetics Society, Seattle, Washington, p. 468.
Sherins,  R.J.,  D.  Brightwell,  and P.M. Sternthal.
  1977. Longitudinal Analysis of Semen in Fertile
  and Infertile  Men. In: The Testes in Normal and
  Infertile Men, P. Troen and H.R. Nankin, eds. Raven
  Press, New York, New York. pp. 473-488.
Shore, M.L, R.P. Felten, and A. Lamanna. 1977. The
  Effect of Repetitive Prenatal Low-Level Microwave
  Exposure on Development in the Rat. In: Symposium
  on Biological Effects and Measurement of Radio

                      R-24
  Frequency/Microwaves,  D.G.  Hazzard, ed. HEW
  Publication (FDA) 77-8026, Rockville, Maryland.
  pp. 280-289.
Siekierzynski,  M., P. Czerski,  H. Milczsret, A.
  Gidynski, C. Czarnecki, E. Dziuk, and W. Jedrzejczak.
  1974a. Health Surveillance of Personnel Occupa-
  tionally Exposed to Microwaves. II.  Functional
  Disturbances. Aerospace Med. 45:1143-1145.

Siekierzynski, M., P. Czerski, A. Gidynski, S. Zydecki,
  C. Czarnecki, E. Dziuk, andW. Jedrezejczak. 1974b.
  Health Surveillance of Personnel Occupationally
  Exposed to Microwaves.  III. Lens Translucency.
  Aerospace Med., 45:116-1148.

Siems, LL, A.J. Kosman, and S.L. Osborne. 1948. A
  Comparative Study of Short Wave and Microwave
  Diathermy  on Blood Flow. Arch. Phys.  Med.,
  29:759-764.
Sigler,  AT.,  A.M. Lilienfeld, B.H. Cohen,  and J.E.
  Westlake. 1965. Radiation Exposure in Parents of
  Children with Mongolism (Down's Syndrome).
  Bull. J. Hopkins Hosp., 117:374-399.
Silverman, C. 1979. Epidemiologic Approach to the
  Study of Microwave Effects. Bull. N.Y. Acad. Med.,
  55:1166-1181.
Silverman,  C.  1980.  Epidemiologic Studies of
  Microwave Effects. Proc. IEEE, 68:78-84.

Skinner, B.F. 1953. Science and Human Behavior.
  The Free Press, New York, New York, 461 pp.
Smialowicz, R.J. 1976. The Effect  of Microwaves
  (2450 MHz) on Lymphocyte  Blast Transformation
  In  Vitro. In: Biological Effects of Electromagnetic
  Waves, Vol. I, C.C. Johnson and M.L. Shore,  eds.
  HEW Publication (FDA) 77-8010,  Rockville,
  Maryland, pp. 472-483.
Smialowicz, R.J., J.B. Kinn, and J.A. Elder. 1979a.
  Perinatal Exposure of  Rats to 2450-MHz  CW
  Microwave Radiation: Effects on  Lymphocytes.
  Radio Sci., 14(6S): 147-153.
Smialowicz, R.J., M.M. Riddle, P.L Brugnolotti, J.M.
  Sperrazza,  and J.B.  Kinn.  1979b. Evaluation of
  Lymphocyte Function in Mice Exposed  to 2450
  MHz (CW) Microwaves. In: Elecfromagnetic Fields
  in  Biological  Systems,  S.S.  Stuchly,  ed.  The
  International Microwave Power Institute, Edmon-
  ton, Canada, pp. 122-152.

Smialowicz, R.J., J.S. AN, E. Berman, S.J.  Bursian,
  J.B. Kinn, C.G. Liddle, LW. Reiter, and C.M. Weil.
  1981 a. Chronic Exposure of Rats to 100-MHz (CW)
  Radiofrequency Radiation: Assessment of Biologi-
  cal Effects. Radiat. Res., 86:488-505.

Smialowicz, R.J., M.M. Riddle, P.L. Brugnolotti,  R.R.
  Rogers, and K.L Compton.  1981b. Detection of
  Microwave Heating in 5-Hydroxytryptamine-

-------
  Induced Hypothermic Mice. Radiat. Res., 88:108-
  117.
Smialowicz, R.J., P.L Brugnolotti, and M.M. Riddle.
  1981c. Complement  Receptor  Positive  Spleen
  Cells in Microwave (2450-MHz) Irradiated Mice. J.
  Microwave Power, 16:73-77.
Smialowicz, R.J., C.M. Weil, J.B. Kinn, and J.A. Elder.
  1982.  Exposure of  Rats to 425-MHz (CW)
  Radiofrequency Radiation: Effects on Lymphocytes.
  J. Microwave Power, 17:211-221.
Smialowicz, R.J., R.R. Rogers,  R.J.  Garner, M.M.
  Riddle, R.W.  Luebke, and  D.G.  Rowe.  1983.
  Microwaves (2450-MHz) Suppress Murine Natural
  Killer  Cell Activity. Bioelectromagnetics,  4:371-
  381.
Smith, J.B., R.P. Knowlton, andS.S. Agarwal. 1978.
  Human Lymphocyte Responses are Enhanced by
  Culture at40°C. J. Immunol., 121:691-694.
Smith,  P.E.,  and  E.W.  James. 1964. Human
  Responses to Heat Stress. Arch. Environ.  Health,
  9:332-342.
Smolyanskaya, A.Z., and R.L. Vilenskaya. 1973.
  Effects of Millimeter-Band Electromagnetic
  Radiation on the  Functional Activity of  Certain
  Genetic Elements of Bacterial Cells. Usp Fiz. Nauk,
  110:571-572.  (Trans,  in Soviet Physics Uspekhi,
  16(4):571-572, 1974.)
Snyder,  S.H.  1971. The Effect of  Microwave
  Irradiation on the Turnover Rate of Serotonin and
  Norepinephrine and the Effect of Monoamine
  Metabolizing Enzymes.  Final Report, Contract No.
  DADA 17-69-C-9144, U.S. Army Medical Research
  and Development Command, Washington, D.C. (NTIS
  AD-729 161). 26pp.
Sommer, H.C., and  H.E.  von Gierke. 1964. Hearing
  Sensations  in  Electric Fields.  Aerospace Med.,
  35:834-839.
Spalding, J.F., R.W. Freyman,  and  L.M. Holland.
  1971.  Effects of 800-MHz Electromagnetic
  Radiation on Body Weight, Activity, Hematopoiesis
  and Life Span  in Mice. Health Phys., 20:421-424.
Spiegel, R.J.,  D.M. Deffenbaugh,  and J.E. Mann.
  1979. Modeling Heat Transfer in Man Exposed to
  an Electromagnetic Field. Final Tech.  Report No.
  14-9239, Southwest  Research Institute, San
  Antonio, Texas. 106 pp.
Spiegel, R.J.,  D.M.  Deffenbaugh,  and J.E. Mann.
  1980a. A Thermal Model of the Human Body
  Exposed to an  Electromagnetic Field.  Bioelectro-
  magnetics, 1(3):253-270.
Spiegel, R.J., W.E. Oakey, and E.L Bronaugh. 1980b.
  A Variable-Volume Cavity Electromagnetic Near-
  Field Simulator.  IEEE  Trans. Electromagnetic
  Compatibility, EMC-22(4):289-297.
Stavinoha, W.B., A. Modak, M.A. Medina, and A.E.
  Gass. 1975. Growth and Development of Neonatal
  Mice Exposed to High-Frequency Electromagnetic
  Fields (NTIS AD-A022 765). 12 pp.
Stern,  S.,  L Margolin, B. Weiss, S. Lu, and S.M.
  Michaelson.  1979.  Microwaves: Effect  on Ther-
  moregulatory Behavior in Rats. Science, 206:1198-
  1201.
Stitt, J.T.  1979. Fever Versus Hyperthermia.  Fed.
  Proc., 38:39-43.
Stodolnik-Baranska, W. 1967. Lymphoblastoid
  Transformation  of  Lymphoctyes  In Vitro after
  Microwave Irradiation. Nature, 214:102-103.
Stolwijk, J.AJ. 1969.  Expansion of a Mathematical
  Model  of Thermoregulation  to  Include  High
  Metabolic Rates.  NASA CR-102192 (NTIS  N70-
  19831), Washington, D.C. 120 pp.
Stolwijk, J.A.J. 1971. A Mathematical  Model of
  Physiological Temperature Regulation in Man.
  NASA CR-1855  (NTIS  N71-33401), Washington,
  D.C. 76 pp.
Stolwijk,  J.AJ. 1980. Mathematical  Methods of
  Thermal Regulation. Ann. N.Y. Acad. Sci., 33:309-
  325.
Stolwijk, J.A.J., and D.J. Cunningham.  1968.
  Expansion of a Mathematical Model  of Ther-
  moregulation  to Include High Metabolic  Rates.
  NASA CR-92443 (NTIS N69-16568), Washington,
  DC. 133pp.
Stolwijk, J.A.J., and J.D.  Hardy. 1966. Temperature
  Regulation in Man — A Theoretical Study. Pflugers
  Arch., 291:129-162.
Stolwijk, J.A. J., and J.D. Hardy. 1977. Control of Body
  Temperature.  In:  Handbook  of  Physiology -
  Reactions to Environmental Agents, Douglas H. K.
  Lee, ed. Williams and Wilkins, Baltimore, Maryland.
  Ch. 4, pp. 45-68.
Stratton,  J.A.  1941.  Electromagnetic Theory.
  McGraw-Hill, New York, New York. pp. 414-420.
Strieker, E.M., and F.R. Hainsworth. 1971. Evaporative
  Cooling in the Rat: Interaction with Heat Loss from
  the Tail.  Q.J. Exp. Physiol., 56:231-241.
Stuchly, M.A.,  and S.S. Stuchly. 1980.  Dielectric
  Properties of Biological Substances - Tabulated. J.
  Microwave Power, 15(1):19-26.
Sugano, Y. 1981. Seasonal Changes in Heat Balance
  of  Dogs Acclimatized to Outdoor Climate. Jap. J.
  Physiol., 31:465-475.
Sulek, K., CJ. Schlagel, W. Wiktor-Jedrzecjzak, H.S.
  Ho, W.M. Leach,  A. Ahmed, J.N.  Woody. 1980.
  Biologic  Effects of  Microwave Exposure.  I.
  Threshold Conditions  for the Induction of the
  Increase in Complement Receptor Positive (CR+)
                                                                   R-25

-------
  Mouse Spleen Cells Following Exposure to 2450-
  MHz Microwaves. Radiat. Res., 83:127-137.
Susskind, C. 1962. Nonthermal Effects of Microwave
  Radiation. Report  RADC-TDR-62-624,  Annual
  Scientific Report (1961 -62) on Contract No. NONR-
  222(92) and Final Report on Contract AF41 (657)-
  114, University of California, Electronics Research
  Laboratory, Berkeley, California.  25 pp.
Susskind, C. 1975. Correspondence on D.R. Juste-
  sen's "Prescriptive Grammar for the Radiobiology
  of Nonionizing Radiation."  J. Microwave Power,
  10(4):357.
Sutton, C.H., and F.B. Carroll.  1979. Effects of
  Microwave-Induced Hyperthermia on the Blood-
  Brain Barrier of the Rat. Radio Sci., 14:329-334.
Swicord, M.L 1971. Microwave Measurements and
  New Types of Detectors for Evaluation of Health
  Hazards.  BRH/DEP Publication No. 71-1.  U.S.
  Department of  Health,  Education and Welfare,
  Public Health Service, Rockville,  Maryland. 33 pp.
Switzer, W.G., and D.S. Mitchell.  1977. Long-Term
  Effects of 2.45-GHz Radiation on  the Ultrastructure
  of the Cerebral Cortex and on Hematologic Profiles
  of Rats. Radio Sci., 12:287-293.

Szmigielski, S. 1975. Effect of 10-cm  (3  GHz)
  Electromagnetic  Radiation (Microwaves)  on
  Granulocytes In Vitro.  Ann.  N.Y. Acad.  Sci.,
  247:275-281.
Szmigielski, S., J. Jeljaszewicz, and M. Wiranowska.
  1975. Acute Staphylococcal Infections in Rabbits
  Irradiated with 3-GHz Microwaves. Ann. N.Y. Acad.
  Sci., 247:305-311.
Szmigielski, S.,  M. Kobus,  and M. Janiak. 1976.
  Enhanced Cytotoxic Effect of Hyperthermia (43°C)
  on Colcemide-Treated Normal  and S\/4o-Trans-
  formed Cells Grown In Vitro. Z. Geschwulstkrankh.,
  47:396-399.
Szmigielski, S., G. Pulverer, W. Hryniewicz, and M.
  Janiak. 1977.  Inhibition of Tumor Growth in  Mice
  by Microwave Hyperthermia, Streptolysin S and
  Colcemide. Radio Sci., 12(6S):185-189.
Szmigielski,  S.,  M.  Janiak, W. Hryniewicz, J.
  Jeljaszewicz, and  G.  Pulverer. 1978. Local
  Microwave Hyperthermia (43°C) and Stimulation
  of the Macrophage and T-Lymphocyte Systems in
  Treatment of Guerin Epithelioma in  Rats. Z.
  Krebsforsch., 91:35-48.
Szmigielski, S., A.  Szydzinski, A. Pietraszek, and M.
  Bielec. 1980. Acceleration of Cancer Development
  in Mice  by Long-Term  Exposition to 2450-MHz
  Microwave  Fields.  In: URSI International Sym-
  poisum Proceedings, Ondes Electromagnetiqueset
  Biologie, A.J. Berteaud and B. Servantie, eds. Paris,
  France, pp. 165-169.
Szmigielski,  S., A. Szymdzinski, A. Pietraszek, M.
  Bielec,  M. Janiak, and  J.K. Wrembel.  1982.
  Accelerated Development of Spontaneous and
  Benzopyrene-lnduced Skin Cancer in Mice Exposed
  to 2450-MHz Microwave Radiation. Bioelectro-
  magnetics, 3:179-191.
Taber,  C.W. 1953. Taber's  Vocabulary of Medical
  Terms. F.A. Davis Company, Philadelphia.
Takashima, S. 1963. Dielectric Dispersion of DNA. J.
  Mol. Biol., 7:455-467.
Takashima,  S., and A. Minikata.  1975.  Dielectric
  Behavior of Biological Macromolecules. Digest of
  Literature on Dielectrics, 37:602-653  (NAS,
  Washington, D.C.).
Takashima, S., B.  Onaral, and H.P. Schwan. 1979.
  Effects of Modulated RF  Energy on the EEC  of
  Mammalian Brains. Radiat. Environ.  Biophys.,
  16:15-27.
Taylor, C.R. 1977.  Exercise and Environmental Heat
  Loads: Different Mechanisms for Solving Different
  Problems? Inter. Rev. Physiol., Environ. Physiol. II,
  Vol. 14, D. Robertshaw, ed. University Park Press,
  Baltimore, Maryland, pp. 119-146.
Taylor, E.M., and B.T. Ashleman. 1974. Analysis of
  Central Nervous System  Involvement in the
  Microwave Auditory Effect. Brain Res., 74:201-
  208.
Taylor, E.M., and B.T. Ashleman. 1975. Some Effects
  of Electromagnetic Radiation  on the Brain and
  Spinal Cord of Cats. Ann. N.Y. Acad. Sci., 247:63-
  73.
Tell, R. 1972. Microwave  Energy Absorption  in
  Tissue. EPA Report PB 208-233, U.S. Environmental
  Protection Agency, Washington, D.C. 53 pp.
Tell, R.A., and F. Harlen. 1979. A Review of Selected
  Biological  Effects and Dosimetric Data Useful for
  Development of Radiofrequency Safety Standards
  for Human Exposure. J. Microwave Power, 14:405-
  424.
Tennent,  D.M. 1946. A  Study of Water Losses
  Through the  Skin  in the  Rat. Am. J.  Physiol.,
  145:436-440.

Thauer, R. 1963. Circulatory Adjustments to Climatic
  Requirements.  Handbook  of Physiology, Vol. III.
  American Physiological Society, Washington, D.C.
  p. 1921.
Thomas, J.R.,  and G.  Maitland. 1979. Microwave
  Radiation  and Dextroamphetamine: Evidence  of
  Combined Effects on Behavior of Rats. Radio Sci.,
  14:253-258.
Thomas, J.R., E.D. Finch, D.W. Fulk, and L.S. Burch.
  1975. Effects of Low-Level Microwave Radiation
  on Behavioral Baselines.  Ann.  N.Y. Acad. Sci.,
  247:425-432.
                       R-26

-------
Thomas, J.R., S.S. Yeandle, and LS. Burch. 1976.
  Modification of Internal Discriminative Stimulus
  Control of Behavior by Low Levels of  Pulsed
  Microwave Radiation. In: Biological Effects of
  Electromagnetic Waves, Vol. I., C.C. Johnson and
  M.L Shore, eds. HEW Publication (FDA) 77-8010,
  Rockville, Maryland, pp. 201-214.
Thomas, J.R., L S. Burch, and S.S. Yeandle. 1979.
  Microwave  Radiation and  Chlordiazepoxide:
  Synergistic Effects on Fixed-Interval Behavior.
  Science, 203:1357-1358.
Thomas, J.R., J. Schrot, and R.A. Banvard.  1980.
  Behavioral Effects of Chlorpromazine and Diazepam
  Combined with Low-Level Microwaves.  Neuro-
  behav. Toxicol., 2:131-135.

Thomson,  R.A.E., S.M.  Michaelson, and J.W.
  Howland.  1965. Modification of X-lrradiation
  Lethality in Mice by Microwaves (Radar).  Radiat.
  Res., 24: 631-635.
Tinney, C.E., J.L. Lords, and C.H. Durney. 1976. Rate
  Effects in Isolated Turtle Hearts  Induced by
  Microwave Irradiation. IEEE Trans.  Microwave
  Theory Techniques, MTT-24:18-24.
Tolgskaya, M.S., andZ.V. Gordon. 1973. Pathological
  Effects of Radio Waves. (Trans, from Russian by B.
  Haigh.) LC Cat. Card 72-94825. Consultants
  Bureau, New York, New York.  pp. 63-106.
Tyazhelov, V.V., S.I. Alekseyev,  and P.A. Grigor'ev.
  1979a. Change in the Conductivity of Phospholipid
  Membranes Modified by Alamethicin on Exposure
  to a High Frequency Electromagnetic Field.
  Biophysics,  23:750-751. (Trans, of Biofizika,
  23:732-733,  1978.)

Tyazhelov, V.V., R.E. Tigranian, E.O. Khizhniak, and
  I.G. Akoev. 1979b. Some Peculiarities of Auditory
  Sensations Evoked  by Pulsed Microwave  Fields.
  Radio Sci., 14(6S):259-263.
Ubeda, A., J. Leal, M.A.  Trillo,  M.A. Jimenez, and
  J.M.R.  Delgado. 1983.  Pulse  Shape of  Magnetic
  Fields Influences Chick Embryogenesis. J. Anat.,
  137:513-536.

United States  Senate Committee on  Commerce,
  Science, and Transportation. 1979.  Microwave
  Irradiation  of the U.S. Embassy in Moscow. U.S.
  Government Printing Office, Washington, D.C. 26
  PP-
Van Demark, N.L, and M.J. Free. 1970. Temperature
  Effects. In: The Testis, Vol.  Ill, A.D. Johnson, W.R.
  Gomes, and N.L. Van Demark, eds. Academic Press,
  New York, New York, Chapter 7, pp. 233-312.

Varma,  M.M.,  and  E.A. Traboulay,  Jr.  1976.
  Evaluation of Dominant Lethal Test and DNA
  Studies in Measuring  Mutagenicity  Caused by
  Non-Ionizing  Radiation. In: Biological  Effects of
  Electromagnetic Waves, Vol. I, C.C. Johnson and
  M.L. Shore, eds. HEW Publication (FDA) 77-8010,
  Rockville, Maryland, pp. 386-396.
Varma, M.M., and E.A. Traboulay. 1977. Comparison
  of Native and Microwave Irradiated DNA. Experien-
  tia,  33:1649-1650.
Varma,  M.M., E.L Dage,  and S.R.  Joshi. 1976.
  Mutagenicity Induced by Non-Ionizing Radiation in
  Swiss Male Mice. In:  Biological  Effects  of
  Electromagnetic Waves, Vol. I, C.C. Johnson and
  M.L. Shore, eds. HEW Publication (FDA) 77-8010,
  Rockville, Maryland, pp. 397-405.
Vendrik, A.J. H., and J.J. Vos. 1958. Comparison of
  the  Stimulation of the Warmth Sense Organ  by
  Microwave and Infrared. J. Appl. Physiol., 13:435-
  444.
Wachtel, H., R. Seaman, and W.Joines. 1975. Effects
  of Low-Intensity Microwaves on Isolated Neurons.
  Ann. N.Y. Acad. Sci., 247:46-62.
Wakim,  K.G., J.W. Gersten, J.F. Herrick, E.G. Elins,
  andF.H. Krusen. 1948. The Effects of Diathermy on
  the  Flow of Blood in the Extremities. Arch. Phys.
  Med.,  29:583-593.
Wangemann, R.T., and S.F. Clean/. 1976. The/n Vivo
  Effects of 2.45 GHz Microwave Radiation on Rabbit
  Serum Components and Sleeping Times. Radiat.
  Environ. Biophys., 13:89-103.
Ward, T.R., J.W. Allis, and J.A. Elder. 1975. Measure
  of Enzymatic Activity Coincident with 2450 MHz
  Microwave  Exposure. J.  Microwave  Power,
  10:315-320.
Weil,  C.M.  1974. Propagation of Plane Waves
  Through  Two  Parallel Dielectric Sheets. IEEE
  Trans. Biomed. Eng.,BME-21(2):165-168. (Adden-
  dum and Corrections, BME-24(1):78-80, 1977.)

Weil,  C.M. 1975.  Absorption  Characteristics  of
  Multilayered Sphere  Models  Exposed to UHF/
  Microwave  Radiation. IEEE  Trans. Biomed. Eng.,
  BME-22(6):468-476.
Weil, C.M. 1977. Review of Exposure Techniques and
  Dosimetric Methods Employed in Microwave
  Bioeffects Research. Proceedings of IEEE Region III
  Conference (Southeastern), Williamsburg, Virginia.
  Cat. No. 77 CHO 1233-6 Region III.  pp. 507-510.
Weil,  C.M.,  W.T. Joines,  and  J.B.  Kinn. 1981.
  Frequency Range of Large-Scale TEM Mode
  Rectangular Strip Lines. Microwave J., 24(11):93-
  100.
Werner, J.  1980. The Concept  of Regulation for
  Human Body Temperature. J. Thermal Biol., 5:75-
  82.
Wever, R. 1973.  Human Circadian Rhythms Under
  the  Influence of Weak Electric Fields and the
  Different  Aspects of These  Studies.  Int.  J.
  Biometeorol., 17:227-232.
                                                                   R-27

-------
Wickersheim, K.A., and R.V. Alves. 1982. Fluoroptic
  Thermometry:  A New RF-lmmune Technology.
  Biomedical Thermology (Prog. Clin. Biol. Res., Vol.
  107, M. Gauthrie and E. Albert, eds.). Alan R. Liss,
  New York, New York. pp. 547-554.
Wiktor-Jedrzejczak, W., A. Ahmed, P. Czerski, W.M.
  Leach, and K.W. Sell. 1977a. Immune Response of
  Mice at 2450-MHz Microwave Radiation: Overview
  of Immunology and Empirical Studies of Lymphoid
  Splenic Cells. Radio Sci., 12(6S):209-219.
Wiktor-Jedrzejczak, W., A. Ahmed, P. Czerski, W.M.
  Leach,  and K.W. Sell.  19775. Increase  in the
  Frequency of Fc Receptor (FcR) Bearing Cells in The
  Mouse Spleen Following a Single Exposure of Mice
  to 2450 MHz  Microwaves.  Biomedicine, 27:250-
  252.
Wiktor-Jedrzejczak, W., A. Ahmed, K.W. Sell,  P.
  Czerski, and W.M. Leach. 1977c. Microwaves
  Induce an Increase  in the Frequency  of Comple-
  ment Receptor-Bearing Lymphoid Spleen Cells in
  Mice. J.  Immunol., 118:1499-1502.
Williams, D.B., J.P. Monalen, W.J. Nicholson, and
  J.J. Aldrich. 1955. Biologic Effects  Studies on
  Microwave Radiation. AMA Arch. Ophth., 54:863-
  874.
Wilson,  B.S., J.M. Zook, W.T.  Joines, and J.H.
  Casseday. 1980. Alterations in Activity at Auditory
  Nuclei  of the Rat Induced by Exposure  to
  Microwave Radiation: Autoradiographic Evidence
  Using [14C]2-Deoxy-D-Glucose. Brain  Res., 187:
  291-306.
Wilson, J.G. 1973. Environment and Birth Defects.
  Academic Press, New York, New York. 305 pp.
Wissler, E.H. 1961.  Steady-State Temperature
  Distribution in Man. J. Appl. Physiol., 16:734-740.
Wissler, E.H.  1964. A Mathematical  Model  of the
  Human  Thermal  System.  Bull. Math. Biophys.,
  26:147-166.
Wyndham, C.H.,  and A.R.Atkins. 1960. An Approach
  to the Solution of the Human Biothermal Problem
  with the Aid of an Analog Computer. Proc. 3rd Int.
  Conf. Medical  Electronics, London, England.
Yamaura,  I.,  and S.  Chichibu. 1967.  Super-High
  Frequency Electric Field and Crustacean Ganglionic
  Discharges. Tohoku J. Exp.  Med., 93:249-259.
Yang, H.K., C.A. Cain, J.  Lockwood,  and W.A.
  Tompkins. 1983. Effects of Microwave Exposure on
  the Hamster Immune System. I. Natural Killer Cell
  Activity. Bioelectromagnetics, 4:123-139.
Youmans,  H.D.,  and H.S. Ho.  1975. Development of
  Dosimetry  for RF  and Microwave Radiation-l:
  Dosimetric Quantities for  RF and  Microwave
  Electromagnetic Fields. Health Phys., 29:313-316.
Zamenhof, S., and S. Greer. 1958. Heat as an Agent
  Producing High Frequency of Mutations and
  Unstable Genes in Escherichia coli.  Nature,
  182(4635):611-613.
Zaret, M.M.  1974. Selected  Cases of Microwave
  Cataract in Man Associated with Concomitant
  Annotated Pathologies. In:  Biologic Effects and
  Health Hazards of Microwave Radiation, P. Czerski,
  K. Ostrowski, M.L.  Shore, C. Silverman, M.J.
  Suess, and B.  Waldeskog,  eds.  Polish  Medical
  Publishers, Warsaw, Poland, pp. 294-301.
Zaret, M.M. 1976.  Electronic Smog as a Potentiating
  Factor in Cardiovascular Disease: A Hypothesis of
  Microwaves as an Etiology for Sudden Death from
  Heart Attack in North Karelia. Med.  Res.  Eng.,
  12:13-16.
Zaret, M.M.  1977. Microwave Radiation (R.O.
  Becker), Dr. Zaret's Reply. Letter to the Editor. N.Y.
  State J. Med., 77:2172-2174.
Zaret, M.M., and  M.  Eisenbud.  1961. Preliminary
  Results of Studies of the  Lenticular Effects of
  Microwaves Among  Exposed Personnel. In:
  Biological Effects of  Microwave Radiation, Vol. I,
  Proceedings of the 4th Tri-Service Conference, M.F.
  Peyton, ed. Plenum Press, New York, New York. pp.
  293-308.
Zaret, M.M., IT. Kaplan, and A.M. Kay. 1970. Clinical
  Microwave  Cataracts.  In: Biological Effects and
  Health Implications of  Microwave Radiation, S.F.
  Cleary, ed. HEW Publication  BRH/DBE 70-2,
  Rockville, Maryland, pp. 82-84.
Zeman,  G.H., R.L Chaput, Z.R.  Glazer, and L.C.
  Gershman. 1973. Gamma-Aminobutyric Acid
  Metabolism in Rats Following Microwave Exposure.
  J. Microwave Power, 8:213-216.
                      R-28

-------
                                           Glossary
ABSORPTION The irreversible conversion of electro-
magnetic energy into other forms of energy as a result
of interaction with matter. See SPECIFIC ABSORPTION
and SPECIFIC ABSORPTION RATE.
ACCLIMATION A physiological change, occurring
within the lifetime of an organism, which reduces the
strain caused by  experimentally induced stressful
changes in particular climatic factors.
a-HELIX A conformation found in protein molecules
where the amino acid chain turns in a helical pattern
with weak chemical bonds (hydrogen bonds) forming
between successive turns in the helix.
AMES TEST A standard test for mutagenic potential
of various agents performed with specialized strains
of bacteria.
ANECHOIC CHAMBER A chamber lined with material
that absorbs RF radiation;  an RF-exposure system
free of scattered and reflected radiation.
ANGULAR MOMENTUM The quantity of motion of a
rotating body - directly  proportional to the angular
velocity of the rotating body.
ANTENNA A device for radiating or receiving  radio
waves.
ANTENNA  REGIONS  The distinction between
electromagnetic fields far from and those near to the
antenna. The regions are usually classified into three
zones: near (static)  zone,  intermediate  (induction)
zone,  and far (radiation) zone. The zones are spatially
located by drawing spheres of different radii around
the antenna. The radii are approximately r < ^ for the
near zone, ^ < r(^) for the intermediate zone, and r >
h for the far zone. Note that -A is the wavelength of the
electromagnetic field produced by the antenna. In the
far zone the field components (E and H) lie transverse
to the direction of the propagation, and the shape of
the field pattern is  independent of the radius at which
it is taken. In the near and intermediate zones the field
patterns are quite complicated, and the shape is, in
general, a function of the radius and-angular position
(azimuth and elevation)  in front of the antenna.

ATHERMAL EFFECT (NONTHERMAL EFFECT) Any
effect  of electromagnetic energy on a body that is not
a heat-related effect.
ATTENUATION A  general term used to denote a
decrease in magnitude of RF transmission from one
point to another.
AUTOLYSIS The decomposition of an organ or tissue
by its own enzymes.

AVERAGE  POWER W The  time-average rate of
energy transfer:
                = 1/(t2-ti)/,t2W(t)dt
For radar calculations, average power W = peak
power x pulse width x pulse repetition frequency.

BASES  Biochemcial  compounds, either purine or
pyrimidines, that are ring structures containing
carbon and nitrogen and which are capable of weak
bonding to each other.
/3-SHEET A conformation found in protein molecules
where two or more portions of amino acid chains line
up side-to-side, with weak hydrogen bonds forming
between adjacent chains.

BIOPOLYMER A  polymeric substance formed in a
biological system, e.g., DNA, proteins.
BLASTOGENIC RESPONSE The transformation of
small  lymphocytes  into  large morphologically
primitive blast-like cells  capable  of  undergoing
mitosis; this phenomenon can be induced in cultured
cells by a variety of agents, including mitogens as well
as antigens,  to  which the  cell donor has been
previously immunized.

BLOOD-BRAIN BARRIER  A functional concept to
explain the observation  that many  substances
transported by blood readily enter other tissues but do
not enter the brain. The barrier functions as if it were
a continuous membrane lining the brain vasculature.
BURSA OF FABRICUS  A lymphoidal organ  in the
hindgut of birds that influences B cell development.

CALCIUM EFFLUX The release of calcium ions from a
sample into a surrounding solution.
CALORIE The amount of heat necessary to  raise the
temperature of one gram of water 1  °C. One calorie
equals 4.184 joules.

CARCINOMA Any of the various types of malignant
neoplasm derived from epithelial tissue, occurring
more  frequently in  the skin, bronchi, stomach, and
prostate gland in men, and in the breast, cervix, and
skin in women.

CATARACTOGENIC Giving rise to the formation of a
cataract, an opacity in the crystalline lens of the eye.
                                              G-1

-------
CHEMOSIS Excessive edema of the ocular conjunc-
tiva.
CHROMOSOMES Large complex biochemical struc-
tures, containing nucleic acid (DNA) and proteins,
which can be visualized in some cells by certain light
microscopy techniques.
CIRCULARLY POLARIZED If the  electric field  is
viewed as a point in space, the locus of the end point
of the vector will rotate and trace out an ellipse once
each cyle.
COLONY-FORMING  UNIT (CPU) Colonies of bone
marrow or blood cells arising from a single progenitor
cell when grown in vitro  or in vivo.
COMPLEX DIELECTRIC PERMITTIVITY The character-
ization of electrical parameters of materials  at the
macroscopic level.
CONFORMATION The spatial distribution of the parts
of a macromolecule in relation to each other, i.e., how
a chain of amino acids folds on itself to form a protein.
CONTINUOUS  WAVE  (CW) Electromagnetic fields
that vary sinusoidally  in time; that is,  those fields
which oscillate at a single frequency.
COUNTER-CURRENT  HEAT EXCHANGE  The heat
exchange  between blood  flowing  in opposite
directions at different temperatures, e.g., adjacent
arteries and veins.
CORE TEMPERATURE The temperature near the
center of the body;  usually measured through the
rectum.
CRITICAL TEMPERATURE, LOWER The ambient
temperature below which the rate of metabolic heat
production of  a resting  thermoregulating animal
increases by shivering and/or nonshivering thermo-
genic  processes to maintain thermal balance.
CRITICAL  TEMPERATURE, UPPER 1. The ambient
temperature above which thermoregulatory evapor-
ative heat loss processes of a resting thermoregulating
animal are recruited. 2.  The ambient temperature
above which there is an  increase in metabolic rate
due to a rise in the core temperature  of a resting
thermoregulating animal.
CYTOSIS Increase in number of cells, e.g., leukocy-
tosis,  increase  in the total  number of circulating
leukocytes.
DEBYE A unit for dipole moment equal to the dipole
moment of a charge distribution of one  positive and
one negative charge, each equal in  magnitude to the
charge of an electron separated by 1 A (10"10m).
DECIBEL (dB) A unit expressing the logarithmic ratio
of two powers or voltages. One tenth of a Bel.
DENSITOMETRY The measurement of exposure to an
RF field; usually expressed in units of milliwatts per
square centimeter.
 DEPTH OF PENETRATION For a plane wave electro-
 magnetic  field incident on the boundary of a good
 conductor, the depth of penetration of the wave is that
 depth at which the field strength of the wave has been
 reduced to 1 /e or approximately 37% of its original
 value.
 DIELECTRIC MATERIAL A class of materials that act
 as electric insulators. For this class the conductivity is
 presumed to be zero, or very small. The positive and
 negative charges in dielectrics are tightly bound
 together so that there is no actual transport of charge
 under the influence of a field. Such material alters
 electromagnetic fields  because of induced charges
 formed by the interaction of the dielectric with the
 incident field.
 DIPOLE A molecule (or other structure) having the
 effective centers of positive and negative  charges
 separated.
 DIPOLE MOMENT A quantity describing the strength
 of a particular dipole:
                 "d*=/vp(r)"r*d7
 where the dipole (d) is the integral of the charge
 distribution (p(T)) times the vector distance (rf from a
 designated origin over the entire space containing the
 charge distribution (v).
 DNA Deoxyribonucleic acid;  the chemical which
 makes up  the genes, the basic unit of heredity.
 DNA MELTING CURVE Characteristic loss of helical
 structure and separation  of the strands  of a  DNA
 molecule when temperature is raised.
 DOSIMETRY The measurement of the absorbed dose
 or dose rate by  an object in a radiofrequency field;
 usually expressed as watts per kilogram or joules per
 kilogram.
 DUTY FACTOR  (CYCLE) The  product of  the pulse
 duration and the pulse repetition frequency.
 ELECTRIC FIELD STRENGTH The force on a stationary
 unit positive charge at a point in an electric field. This
 force may  be measured in volts per  meter (V/m).
 ELECTROMAGNETIC RADIATION (EMR) Energy in
 the form of electric and magnetic fields.
 ELECTROMAGNETIC WAVE A wave characterized by
 variations  of electric and magnetic fields.  Electro-
 magnetic waves are categorized as radio waves, light
 rays, etc., depending on the frequency.
 ELECTROPHORETIC MOBILITY The ability of a cell or
 macromolecule to move in response to a constant
 electric field.

 ELECTROSTRICTIVE  FORCE  Force  exerted by an
 electrostatic field that causes the elastic deformation
of a dielectric.
 ELLIPSOID SHAPE A surface, all plane sections of
which are  ellipses or circles.
                       G-2

-------
ENZYME KINETICS (ACTIVITY)  Measure of  how
rapidly an enzyme catalyzes a chemical reaction.
EPITHELIOMA Carcinoma derived from squamous
cells (scale-like cells) or from the basal and adnexal
cells (accessory cells) of the skin.
ERYTHROPOIESIS The production of red blood cells.
EUKARYOTE An organism composed of one or more
complex cells containing chromosomes that are
segregated  from  the rest of the cell  by a nuclear
membrane.  This distinction is in contrast to bacteria,
which have less complex DNA structure distributed
throughout  the cell volume.
EXCITATION The absorption of energy by a molecule
or other structure.
FAR FIELD REGION See antenna regions.
FEVER A pathological condition in which there is an
abnormal rise in core temperature. The extent of the
rise is variable. The temperature rise in an individual
may be considered as fever when it is greater than the
mean standard deviation for the species in basal
conditions.
FIELD-INDUCED MIGRATION The physical movement
of charged bodies under  the influence of an
electromagnetic field.
50TH PERCENTILE For  a large set of measurements
arranged in order of magnitude, the 50% percentile is
the value such that 50% of the measurements are
less than that value and 50% are greater.
FINITE DIFFERENCE TECHNIQUE The approximation
of differentials by their finite difference; e.g., dy/dx ~
Ay/Ax.
FIRST-ORDER DIFFERENTIAL EQUATION A differen-
tial  equation of order one  (only  single-order
derivatives are included in the equation).
FREQUENCY The number of sinusoidal cycles made
by electromagnetic  radiation in one second; usually
expressed in units of hertz.
GIGAHERTZ (GHz) One billion cycles per second.

HERTZ (Hz) One cycle per second.
HISTOPATHOLOGY The department of pathology
concerned with minute structure, composition, and
function of diseased tissues; microscopic pathology.
HOMEOTHERMY The pattern of temperature regula-
tion in a warm-blooded species in which the cyclic
variation in core temperature, either nychthemerally
or seasonally, is maintained within arbitrarily defined
limits (±2°C)  despite  much larger variations in
ambient temperature.
HUMORAL Relating to the extracellular fluids of the
body, i.e., blood and lymph. In immunology, the name
ascribed to immune mechanisms leading to antibody
products.
HYPERTHERMIA  The condition of a temperature-
regulating animal when the core temperature is more
than one standard deviation above the mean core
temperature of the species in resting conditions in a
thermoneutral environment.
HYPOTHERMIA The condition of a temperature-
regulating animal when the core temperature is more
than one standard deviation below the mean core
temperature of the species in resting conditions in a
thermoneutral environment.
INSENSIBLE WATER LOSS The sum of the water lost
by diffusion through the skin and water lost  in
breathing, and excluding any water excreted.
IN UTERO Within the uterus or womb.
IN VITRO Within glass; observable in a  test tube.
IN VIVO Within the living body.
INFINITE SLAB A piece of material that has an infinite
cross section but finite thickness.
IONIZING ELECTROMAGNETIC RADIATION Electro-
magnetic radiation of high frequency,  short wave-
lengths, and high  photon energy which, when it
interacts with matter, causes  the  removal  of
electrons from atoms, e.g., x-rays and gamma rays.

KILOHERTZ (kHz) One thousand cycles  per second.

LATENT HEAT  OF VAPORIZATION The quantity of
heat released (or  absorbed) in the reversible process
of evaporation (or condensation) of unit mass of liquid
(or vapor) under isobaric and isothermal equilibrium
conditions.
LOSSY CAPACITOR A capacitor containing a
dielectric material with a loss tangent above 0.1.

MACROMOLECULE A molecule, such as a protein or
a nucleic acid,  that has a  molecular weight greater
than a  few thousand.

MEGAHERTZ (MHz) One million cycles  per second.
METABOLIC RATE See resting metabolic rate.
M ETASTABLE A state that is not stable but wi II exist for
a long  period of time.
MICROWAVES  A  particular  segment of the RF
radiation spectrum with a frequency range of 300
MHz to 300 GHz.
MITOGEN A substance that stimulates  lymphocytes
to proliferate independently of any specific antigen.

MODULATION The process of varying the amplitude,
frequency, or phase of an  RF carrier wave.
MULTILAMELLAR VESICLES  Phospholipid bilayers
that  form a series of concentric closed  spherical
structures somewhat analogous to the layers of  an
onion;  these structures  are  commonly used  as
models for studying membrane properties.
                                                                    G-3

-------
MYELOPOIESIS The formation of bone marrow and
the cells that arise from it.

NEUROTRANSMITTER A chemical substance that
transmits nerve impulses across a synapse.
NEAR FIELD REGION See antenna regions.
NONIONIZING RADIATION  (NIR) Electromagnetic
radiation of low frequency, long wavelength, and low
photon energy, unable to cause ionization (i.e.,  to
remove an electron from an atom); e.g., RF radiation.

NONTHERMAL Not related to heat.
NUCLEIC ACIDS Biochemical compounds, consisting
of one or more subrings of a base, a sugar, and a
phosphate  group. When subrings are covalently
bonded to each other, forming a chain, the bond
occurs between the sugar and the  phosphate, with
the base off to the side. In DNA (q.v.), two such
chains are  weakly attached to each other through
contact of their bases, which  form the rungs of a
ladder, while the sugar and phosphate groups form
the sides of the ladder. The bases associate with each
other only if their chemical structure  is compatible,
i.e., complementary. There are four such bases
commonly found in  DNA. The sequence of these
bases provides the information necessary to make
other biochemical molecules.

OPERANT CONDITIONING The process of rewards
and reinforcements by which specific  behaviors are
learned.
ORGANOGENESIS The development  or growth  of
organs, especially embryologic.
OSMOTIC  FRAGILITY The  tendency  of  a cell
membrane to break because of a large imbalance of
ion concentration inside and outside the cell.
PARABOLIC REFLECTOR One of the most widely
used microwave antennas, consisting of a metal disk
whose surface forms a circular parabola.
PARAMETER  Any of a set  of physical  properties
whose values  determine  the characteristics  or
behavior of something.
PHAGOCYTOSIS The engulfing of microorganisms,
other cells, or foreign bodies by phagocytes.
PENIA Reduction in the number of cells;  e.g.,
neutropenia, reduction in the number of polymorpho-
nuclear neutrophiles.

PERITONEAL  Relating to the peritoneum, which is
the serous sac lining the abdominal cavity and
covering most of the viscera therein contained.

PHILIA Increase in number of cells; e.g., neutrophilia,
increase in  number of neutrophiles.
PHONON A particle of mechanical vibrational (sound)
energy.

PHOTON  A particle of electromagnetic energy.
PHOSPHOLIPID BILAYER A double-layered sheet of
phospholipid molecules arranged so that the hydrophi-
lic (water-liking) part of the molecules associate with
water and the hydropholic (water-disliking) part of the
molecules  associate  with each other and avoid
contact with water; this structure is the basis for
biological membranes.

PLASMID Short piece of DNA that normally codes for
one of a few proteins and can often be transferred to
another cell; it is usually  separate from the cell's
major  DNA, which  carries the information  for
reproducing the cell.

PLANE WAVE An electromagnetic wave in which the
electric and magnetic field vectors lie in a plane
perpendicular to the direction of wave propagation.

POIKILOTHERM A cold-blooded animal; an ectotherm;
an  animal  with little or  no control  of its body
temperature.
POIKILOTHERMY The pattern of thermoregulation of
a  species  exhibiting  a large  variability of core
temperature as a proportional function of ambient
temperature.

POLARIZABILITY A linear coefficient that quantitates
the change in the magnitude of the dipole moment
(q.v.) of a molecule or any other structure in response
to an electric field, i.e.,
                  "d*=crE*+c£

where or is_the polarizability, E is the applied electric
field, and 87 is the dipole moment when there is no
applied field.
POWER DENSITY Magnitude of the Poynting vector at
a point in space, in power per unit area (watts per
square meter). For plane waves E2  is simply related to
power  density, and it is the quantity measured by a
survey meter when the sensing element is sensitive
to the square of the magnitude of the electric fields;
i.e., P = EV3770, in mW/cm2.

PROLATE SHPEROID  An  approximately spherical
object  that  is elongated in the direction of a line
joining the poles; similar to a football.

PUPA The second stage in the development of an
insect, between the larva and the imago.
Qio The ratio of the rate of a physiological process at a
particular temperature to the rate at a temperature
10°C lower, when the logarithm of the rate is an
approximately linear function of temperature.

RADIATION The transfer of energy from one body to
another through an intervening medium.
RADIOFREQUENCY RADIATION See RF radiation.

RESONANCE A small electrical stimulus at a given
frequency that produces a large amplitude response
in the system at the same frequency.
                      G-4

-------
RESONANT  FREQUENCY That frequency  which
produces resonance  in a system; typically those
frequencies whose wavelengths are integral multiples
of the body's length.

RESPIRATORY CONTROL  RATIO. A  measure of
mitochondria!  activity and  integrity;  the  rate of
utilization of oxygen in the presence of adenosinedi-
phosphate (ADP) divided by the rate of  utilization of
oxygen in the absence of ADP.

RESTING METABOLIC RATE (RMR) The metabolic
rate of an animal which is resting in a thermoneutral
environment but not in the postabsorptive state. The
relationship of RMR (W/kg) to body mass, M (kg), is
RMR = 3.86M"024 Basal metabolic rate (BMR) is the
rate of energy production of an animal in a rested,
awake, fasting, and thermoneutral state.

RF RADIATION Radiofrequency radiation; nonionizing
electromagnetic radiation in the frequency range 0 to
3000 GHz.
ROOT MEAN SQUARE (RMS) Certain electrical
effects are proportional to the square root of the mean
value of the square of a periodic function (over one
period). This value is known as the effective value or
the root-mean-square (RMS)  value since it is derived
by first squaring the function, determining the mean
value of this squared value, and extracting the square
root of the mean value to determine the end result.
SARCOMA A  tumor, usually  highly  malignant,
formed by proliferation of poorly differentiated cells; a
malignant connective  tissue neoplasm.

SCHEDULE  OF  REINFORCEMENT Also called
reinforcer schedule. The specification of the way in
which reinforcers  are assigned to particular
responses within  an operant-class  or classes.
Examples include: the fixed ratio schedule, in which
the last of a constant, specified number of responses
is reinforced; the fixed interval schedule, in which a
constant, specified period of time must elapse before
a response is reinforced; the differential reinforce-
ment of low  rate schedule, in which a response is
reinforced only if at least a  specified period of time
has elapsed since the last response.
"SILK" Polyester material used in the silk-screen
graphic process.

SPECIFIC ABSORPTION The absorbed energy in the
tissue, in  joules per kilogram  (J/kg). See also
SPECIFIC ABSORPTION RATE.

SPECIFIC ABSORPTION RATE (SAR)  The  rate at
which energy is absorbed in  the tissue, in watts per
kilogram: SAR =a£i2/p, where a = tissue conductivity
at irradiation frequency, Et = rms electric field
strength in the tissue, and p = tissue density (kg/m3).

SPECIFIC HEAT (c) The quantity of heat required to
raise the temperature  of unit mass of a substance by
SPONTANEOUS BEHAVIOR Unlearned or natural
responses to a stimulus.
SQUARE LAW The output of a device is proportional
to the square of the input to the device.
SUBSTRATE The molecule upon which an enzyme
catalyzes a chemical reaction.
SURVEY  INSTRUMENT A portable instrument
capable of measuring the strength of electric and
magnetic fields.
SYNGENEIC Individuals of a species that are
genetically identical  at all relevant transplantation
loci.
TEMPERATURE, AMBIENT (Ta) The  average tem-
perature of a gaseous or liquid environment (usually
air or  water) surrounding a  body,  as measured
outside the  thermal and hydrodynamic boundary
layers that overlay the body.
TEMPERATURE, CORE The mean temperature of the
tissues at a depth below that which is affected directly
by a change in the temperature gradient  through
peripheral tissues. Mean core temperature cannot be
measured accurately, and is generally represented by
a specific  core temperature, e.g., that of the rectum
(Synonym: TEMPERATURE, DEEP BODY.)
TEMPERATURE-HUMIDITY INDEX  A means of
estimating the heat  stress caused  by variations in
temperature and humidity.
TEMPERATURE  REGULATION  The  maintenance of
the temperature or temperatures of a body within a
restricted range under conditions involving variable
internal and/or external heat loads. Biologically, the
existence of some  degree of body temperature
regulation by autonomic or behavioral means.
TEMPERATURE REGULATION, AUTONOMIC The
regulation of body temperature by autonomic (i.e.,
involuntary) responses to heat and cold which modify
the rates of  heat production and heat loss (i.e., by
sweating, thermal tachypnea, shivering, and
variations in peripheral vasomotor  tone and  basal
metabolism).
TEMPERATURE REGULATION, BEHAVIORAL The
regulation of body temperature by complex patterns
of responses of the skeletal musculature to heat and
cold which modify the rates of heat production and/or
heat  loss (e.g.,  by exercise,  change  in  body
conformation,  and in the thermal  insulation of
bedding and (in man) of clothing, and by the selection
of an environment which reduces thermal stress).
TERATOLOGY That division  of embryology and
pathology  which deals  with abnormal development
and congenital malformations.
THERMAL EFFECT In the biological tissue or system,
an effect  that is  related to heating  of the tissue
through the application  of electromagnetic fields,
and that can occur through other forms of heating.
                                                                    G-5

-------
THERMOGRAM A recording of temperature distribu-
tion over a surface or in a material as measured at the
surface.
THERMOGENIC  LEVELS Power  densities of RF
radiation which  produce  measurable  temperature
increase in the exposed object.
THERMOGRAPHY A  technique for detecting  and
measuring variations  in  heat emitted by various
regions of the body.
THERMONEUTRAL ZONE (TNZ) The range of ambient
temperature  within  which metabolic  rate is  at a
minimum, and within which temperature regulation
is achieved  by nonevaporative  physical processes
alone.
THERMOREGULATION See temperature regulation.

TRANSDUCE To  convert one form of energy  into
another form, e.g., from heat to electrical current in a
thermocouple.
TWIN-WELL  CALORIMETRY The technique used to
determine the absorbed  RF  dose  or dose  rate by
comparing the excess heat in an RF-exposed system
to an identical unexposed system  in a comparison
well.
VIGILANCE The degree of attentiveness or watchful-
ness. Vigilance is often investigated by measuring the
number of times the occurrence  of an infrequent
event is  correctly detected.  In  some  experiments
objective measurements are  made of the times the
subject looks for occurrence of the  infrequent event.
Called observing responses, these responses may be
required to allow observation of the  infrequent event.

WAVE A disturbance that moves through a medium.
WAVEGUIDE A transmission line comprised  of a
hollow conducting tube within which electromagnetic
waves may be propagated.
WAVE, TRANSVERSE ELECTRIC (TE) In a homogen-
eous, isotropic medium, an electromagnetic wave in
which  the electric   field vector is  everywhere
perpendicular to the direction of the propagation.
WAVE, TRANSVERSE ELECTROMAGNETIC (TEM) In
a homogeneous isotropic medium, an electromagnetic
wave in which the electric and the magnetic  field
vectors are everywhere perpendicular to the direction
of propagation.

WAVE, TRANSVERSE MAGNETIC  (TM) In a homo-
geneous isotropic medium, an electromagnetic wave
in which the magnetic field vector is everywhere per-
pendicular to the  direction of the propagation.
WAVELENGTH The distance between points of
corresponding phase of  a periodic wave of  two
constant cycles. The wavelength A is related to the
phase velocity v and the frequency by A = v/f.
WHOLE-BODY IRRADIATION Pertains to the case in
which the entire body is exposed to the  incident
electromagnetic energy or the case in which the cross
section (physical area) of the body is smaller than the
cross section of the incident radiation beam.
                      G-6

-------
                                       APPENDIX A
                                 Science Advisory Board


Subcommittee on the Biological Effects of Radiofrequency Radiation
Charles Susskind, Ph.D.
Chairman,  Radiofrequency  Health  Effects Sub-
committee
College of Engineering
University of California, Berkeley
Berkeley, CA 94720
Eleanor Adair, Ph.D.
John B. Pierce Foundation Laboratory
290 Congress Avenue
New Haven, CT06519
Barbara Chang, M.D.
Chief, Hematology/Oncology Section
V.A. Medical Center
Augusta,  GA 30910
Stephen Cleary, Ph.D.
Department of Physiology
Medical College of Virginia
Virginia Commonwealth University
Richmond, VA 23298
Carl Durney, Ph.D.
Department of Electrical Engineering
University of Utah
Salt Lake City, UT84112
Arthur Guy, Ph.D.
Bioelectromagnetics Research Laboratory
Department of Rehabilitation Medicine
University of Washington
Seattle, WA 98195
Steven Horvath, Ph.D.
Institute of Environmental Stress
University of California
Santa Barbara, CA 93106
Abraham Lilienfeld, M.D.
The Johns Hopkins University
School of Hygiene and Public Health
615 N.Wolfe Street
Baltimore, MD 21205
Solomon Michaelson, D.V.M.
Department of Radiation Biology and Biophysics
University of Rochester School of Medicine  and
Dentistry
602 Elmwood Avenue
Rochester, NY 14642
Mary Ellen O'Connor, Ph.D.
Department of Psychology
University of Tulsa
600 S. College Avenue
Tulsa, OK 74104
Charlotte Silverman, M.D.
(Committee consultant/liaison)
Food and Drug Administration
National Center for Devices and Radiological Health
12721 Twinbrook Parkway
RockviMe, MD 20852
                                             A-}

-------

-------
                                        APPENDIX B
                                          Authors
Ernest N. Albert1
Joseph S. Ali
JohnW. Allis
Ezra Berman
Carl F. Blackman
Daniel F. Cahill2
Joe A. Elder
Michael I. Gage
Christopher J. Gordon
Doreen Hill3
William T. Joines4
James B. Kinn
William P. Kirk5
Charles G. Liddle
James R. Rabinowitz
Ralph J. Smialowicz
Ronald J. Spiegel
Claude M. Weil6
                             U.S. Environmental Protection Agency
                             Office of Research and Development
                             Office of Health Research
                             Health Effects Research Laboratory
                             Research Triangle Park, NC 27711

                             'Department of Anatomy
                             The George Washington University
                             Medical Center
                             Washington, DC 20037

                             "Also with Department of Electrical Engineering
                             Duke University
                             Durham, NC 27706


                             Current Addresses:

                             2Carolina Power and Light Company
                             Harris Energy and Environmental Center
                             Route 1, Box 327
                             New Hill, NC 27562

                             3Office of Radiation Programs
                             U.S. Environmental Protection Agency
                             Washington, DC 20460

                             5Three Mile Island Field Station
                             U.S. Environmental Protection Agency
                             100 Brown Street
                             Middletown, PA 17057

                             6Boeing Company, BMAC 41 -08
                             P.O. Box 3707
                             Seattle, WA 98124
                                             B-1
                                                                  « U.8.aoV0MMENTFMNTMa OFFICE 1H4 - 759-102/107M

-------

-------

-------
United Elates
Environmental Protection
Agency
Center for Environmental Research
Information
Cincinnati OH 45268
                                                                  Please mike all necessary changes on the above label,
                                                                  detach or copy, and return to the address in the upper
                                                                  lelt-hand corner

                                                                  If you do not wish to receive these reports CHECK HERE a,
                                                                  detach, or copy this cover, and return to the address in the
                                                                  upper left-hand corner
                                                                EPA-600/8-83-026F

-------