Skip to main content

Full text of "Annual report of the director of the Geophysical Laboratory"

See other formats


■ 


CARNEGIE 
INSTITUTION 


Annual  Report  of  the  Director 
Geophysical  Laboratory 

5251  BROAD  BRANCH  ROAD,  NORTHWEST,  WASHINGTON,  D.C.  20015-1305 

1990-1991 


For  the  year  July  1, 1990-June  30,  1991 

Issued  December  1991 

Papers  from  the  Geophysical  Laboratory 

Carnegie  Institution  of  Washington 

NO.  2250 


Digitized  by  the  Internet  Archive 

in  2012  with  funding  from 

LYRASIS  Members  and  Sloan  Foundation 


http://www.archive.org/details/annualreportofd199091carn 


Geophysical  Laboratory 


Washington,  District  of  Columbia 


Charles  T.  Prewitt 
Director 


Published  by:     Geophysical  Laboratory 

5251  Broad  Branch  Rd.,  N.W. 
Washington,  D.C.,  20015-1305 
USA 


ISSN  0576-792X 
December  1991 


When  used  in  bibliographic  citations,  The  Annual  Report  should  be  cited  as  follows: 

Author,  Title,  Annu.  Rep.  Director  Geophys.  Lab.,  Carnegie  Instn.  Washington,  1990-1991,  pagina- 
tion, 1991. 


GEOPHYSICAL  LABORATORY 


Contents 


Introduction 1 

Igneous  and  Metamorphic  Petrology  — - 
Field  Studies 3 

Global  Convection  and  Hawaiian  Upper  Mantle 

Structures.  T.Neil  Irvine 3 

Megacrystalline  Dunites  and  Peridotites:  Hosts 

for  Siberian  Diamonds.  N.  P.  Pokhilenko,  D. 

G.  Pearson ,  F.  R.  Boyd, 

andN.V.Sobolev 11 

Mantle  Metasomatism:  Evidence  from  a  MARID 

-  Harzburgite  Compound  Xenolith. 

F.R.Boyd 18 

Boron  and  Beryllium  Concentrations  in  Subduc- 

tion-Related  Metamorphic  Rocks  of  the 

Catalina  Schist:  Implications  for  Subduction- 

Zone  Recycling.  Gray  E.  Bebout,  Jeffrey  G. 

Ryan,  and  William  P.  Leeman 23 

Laser  Fluorination  of  Sulfide  Minerals  with  F2 

Gas.  D.  Rumble,  J.  M.  Palin, 

andT.  C.  Hoering 30 

Stable  Isotope  and  Trace  Element  Indicators  of 

Devolatilization  History  in  Metashales  and 

Metasandstones.  Gray  E.  Bebout 34 

The  Fa  Content  of  Normative  ol. 

Felix  Chayes 40 

Igneous  and  Metamorphic  Petrology  — 
Experimental  Studies 45 

Raman  Spectra  of  High-Temperature  Silicate 
Melts:  Na20-Si02,  K20-SiC>2,  andLi20-SiC>2 
Binary  Compositions.  John  D.  Frantz  and 
Bjorn  O.  Mysen 45 

Peralkalinity  and  H2O  Solubility  Mechanisms  in 
Silicate  melts,  Bjorn  Mysen 53 

Partitioning  of  fluorine  and  chlorine  between  apa- 
tite and  non-silicate  fluids  at  high  pressure  and 
temperature.  James  Brenan 61 

Investigation  of  Fluid  Immiscibility  in  the  System 
H20-NaCl-CC>2  Using  Mass  Spectrometry  and 
Microthermometry  Techniques  Applied  to 
Synthetic  Fluid  Inclusions. 
Robert  K.  Popp,John  D.  Frantz, 
andThomas  C.  Hoering 68 


Akermanite-Gehlenite-Sodium  Melilite  Join  at 

0 

950  C  and  5  kbar  in  the  Presence  of  CO2  + 

H2O.  H.G.  Huckenholz,  H.S.  Yoder,  Jr.,  T. 

Kunzmann,  and  W.  Seiberl 75 

Merwinite  Stability  and  High-Temperature  Phase 

Relations  in  the  Presence  of  CO2 +H2O.  H.  G. 

Huckenholz,  H.  S.  Yoder,  Jr., 

and  W.  Seiberl 81 

The  System  Mg2Si04-Fe2SiC>4  at  Low  Pressure. 

Hiroko  Nagahara,  Ikuo  Kushiro, 

andBjorn  O.  Mysen 88 

Fe3+,  Mg  order-disorder  in  heated  MgFe204:  a 

powder  xrd  and  57pe  mossbauer  study.  H.  St. 

C.  O'Neill,  H.  Annersten  andD.  Virgo  ....93 

Crystallography  -  Mineral  Physics 101 

Predicted  High-Pressure  Mineral  Structures  with 
Octahedral  Silicon.  Robert  M.  Hazen  and  Larry 
W.  Finger 101 

Simultanous  High  P-T  Diffraction  Measurements 
of  (Fe,Mg)Si03-Perovskite  and  (Fe,Mg)0 
Magnesiowiistite:  Implications  for  Lower 
Mantle  Composition.  YingweiFei,  Ho-Kwang 
Mao,  Russell  J.  Hemley,  and  Jinfu  Shu  ..107 

High-Pressure  Crystal  Chemistry  of  Iron-Free 
Wadsleyite,  p-Mg2SiC>4  Jinmin  Zhang,  Rob- 
ert M.  Hazen,  and  Jaidong  Ko 115 

Phase  Transitions  in  Framework  Minerals. 
David  Palmer 120 

First-prirciples  Studies  of  Elasticity  and  Post- 
Stishovite  Phase  Transitions  in  Si02. 
Ronald  E.  Cohen 126 

Molecular  Dynamics  Simulations  of  Melting  of 
MgO  at  High  Pressures.  Zhaoxin  Gong, 
Ronald  E.  Cohen,  and  Larry  L.  Boyer  ...  129 

Glass  Diffraction  Measurements  with  Polychro- 
matic Synchrotron  Radiation.  Charles  Meade 
and  Russell  J .  Hemley 135 

X-Ray  Diffraction  of  Solid  Nitrogen-Helium  Mix- 
tures. Willem  L.  Vos,  Larry  W.  Finger, 
Russell  J.  Hemley,  Ho-Kwang  Mao, 
Jing  Zhu  Hu,  Jin  Fu  Shu,  Richard  LeSar, 
Andre  de  Kuijper, 
and  Jan  A.  Schouten 138 


CARNEGIE  INSTITUTION 


Evidence  for  Orientational  Ordering  of  Solid  Deu- 
terium at  High  Pressures.  Russell  J.  Hemley 
and  Ho-Kwang  Mao 141 

BlOGEOCHEMlSTRY 147 

Nitrogen  Isotope  Tracers  of  Atmospheric  Deposi- 
tion in  Coastal  Shelf  Waters  off  North  Caro- 
lina. 
Marilyn  L.  Fogel  and  Hans  W.  Paerl 147 

Nitrogen  Diagenesis  in  Anoxic  Marine  Sediments: 
Isotope  Effects.  David  J.  Velinsky,  David  J. 
Burdige,  and  Marilyn  L.  Fogel 154 

The  Isotopic  Ecology  of  Plants  and  Animals  in 
Amboseli  National  Park,  Kenya.  PaulL.  Koch, 
Anna  K.  Behrensmeyer, 
and  Marilyn  L.  Fogel 163 


Rapid  Racemization  of  Aspartic  Acid  in  Mollusk 
and  Ostrich  Eggshells:    A  New  Method  for 
Dating  on  a  Decadal  Time  Scale.  Glenn  A. 
Goodfriend,  David  W.  von  Endt, 
and  P.E.  Hare 172 

A  Burning  Question:  Differences  between  Labo- 
ratory-Induced and  Natural  Diagenesis  in  Os- 
trich Eggshell  Proteins.  A  S.  Brooks,  P.E.  Hare, 
J.E.  Kokis  and  K.  Durana 176 

Publications 181 

Personnel 185 


GEOPHYSICAL  LABORATORY 


Introduction 


Last  year's  introduction  to  the  Annual 
Report  described  the  co-location  of  the 
Geophysical  Laboratory  and  the  Depart- 
ment of  Terrestrial  Magnetism  in  our  new 
and  newly-renovated  building  complex  on 
Broad  Branch  Road.  This  year  the  con- 
struction and  renovation  were  completed 
and  both  departments  are  pursuing  their 
normal  research  objectives  with  almost  all 
of  our  equipment  operating  as  it  should. 
Moving  and  the  relocation  of  laboratories 
has  been  disruptive  for  many  staff  mem- 
bers, but  we  all  believe  that  the  new  envi- 
ronment and  proximity  to  colleagues  at 
DTM  are  worth  the  effort. 

The  principal  new  research  initiative 
for  the  Geophysical  Laboratory  this  year  is 
our  participation  in  the  new  Center  for 
High  Pressure  Research  with  the  State 
University  of  New  York  at  Stony  Brook 
and  Princeton  University.  This  is  one  of  the 
14  Centers  established  in  1991  through 
funding  by  the  NSF  Science  and  Technol- 
ogy Center  Program.  Depending  on  Con- 
gressional budgetary  approvals,  NSF  in- 
tends to  continue  the  Program  for  at  least 
four  more  years,  and  it  is  possible  that 
funding  could  extend  for  a  total  of  eleven 
years.  The  funds  supplied  by  NSF  together 
with  contributions  from  the  three  institu- 
tions and  additional  external  grants  will  be 
used  to  support  a  variety  of  initiatives  re- 
lated to  high-pressure  research. 

The  Center  is  composed  of  staff,  students, 
and  laboratories  at  the  Geophysical  Labo- 
ratory, the  Department  of  Earth  and  Space 
Sciences  at  Stony  Brook,  and  the  Depart- 


ment of  Geological  and  Geophysical  Sci- 
ences at  Princeton.  GL  staff  members 
involved  with  the  Center  are  Francis  Boyd, 
Ronald  Cohen.  Larry  Finger,  Robert  Hazen, 
Ho-kwang  Mao,  Russell  Hemley,  Bjom 
Mysen,  Charles  Prewitt,  and  David  Virgo. 
In  addition  to  the  above  institutions,  col- 
laboration is  being  developed  between  the 
Center  and  other  laboratories  in  universi- 
ties, industry,  and  government.  The  princi- 
pal goal  of  the  Center  is  to  study  fundamen- 
tal questions  about  the  Earth's  evolution, 
structure,  and  dynamic  state,  and  about  the 
nature  of  interiors  of  other  planets.  In  addi- 
tion, we  expect  to  generate  extensive  new 
information  about  material  properties  at 
high  pressures  and  temperatures,  and  to 
synthesize  new  materials  of  interest  to  phys- 
ics, chemistry,  and  materials  science  as 
well  as  to  the  earth  sciences.  Experimental 
work  will  be  complemented  by  theoretical 
computer  simulations  and  by  development 
of  new  equipment  and  techniques  for  high- 
pressure  research,  including  larger-volume 
experimental  apparatus. 

Systematic  high-pressure  work  on  ma- 
terials of  geological  interest  has  been  a 
fundamental  component  of  Geophysical 
Laboratory  activities  since  1904  and  GL 
staff  have  played  the  major  role  in  the 
development  and  utilization  of  many  types 
of  high-pressure  apparatus,  including  pis- 
ton-cylinder, cold-seal,  gas-media,  and  dia- 
mond-anvil devices.  For  example,  in  re- 
cent years  Ho-kwang  Mao  and  his  col- 
leagues have  led  the  development  and  ap- 
plication of  the  diamond-anvil  cell  for  ex- 


CARNEGIE  INSTITUTION 


perimental  studies  at  high  pressure.  Static 
pressures  of  about  5.0  megabars — substan- 
tially greater  than  that  at  the  Earth's  cen- 
ter— have  been  attained,  the  important  pres- 
sure-measuring scale  using  ruby  fluores- 
cence has  been  extended  to  1.8  megabars, 
and  pressure-scale  x-ray  diffraction  studies 
have  been  extended  above  3  megabars. 
Techniques  have  been  developed  for  heat- 
ing samples  within  the  cell  by  laser  and  for 
studying  them  by  means  of  a  variety  of 
spectroscopic  techniques.  The  establish- 
ment of  the  Center  will  allow  us  to  continue 
and  extend  this  kind  of  innovation  to  greater 
extremes  of  pressure  and  temperature,  larger 
sample  volumes,  and  experiments  on  many 
different  kinds  of  materials. 

In  recent  years,  Mao,  Russell  Hemley, 
and  colleagues  have  have  concentrated 
much  of  their  work  on  inert  gases  that 
crystallize  into  solid  forms  at  high  pres- 
sures. Their  experiments  with  solid  hydro- 
gen have  exceeded  2.5  megabars,  where 
they  discovered  new  phase  transitions  and 
the  first  evidence  of  transformations  into 
the  metallic  form.  To  support  this  research, 
a  number  of  staff  members  and  postdoctoral 
fellows  have  been  active  in  developing  and 
using  x-ray  and  infrared  beam  lines  at  the 
National  Synchrotron  Light  Source, 
Brookhaven  National  Laboratory,  for  ex- 
periments that  could  not  be  performed  sat- 
isfactorily without  the  use  of  synchrotron 
radiation.  In  particular,  the  superconduct- 
ing wiggler  beam  line  X17C  provides  a 
high-energy  beam  with  very  high  intensity 
x-rays  for  probing  tiny  samples  in  dia- 
mond-anvil cells.  The  infrared  beam  line 
U4  provides  high-intensity  infrared  radia- 
tion for  spectroscopic  measurements  of 
samples  in  diamond-anvil  cells,  and  is  the 


only  facility  of  its  type  anywhere  in  the 
World. 

Another  new  development  this  year  is 
the  installation  of  a  Dilor  micro-Raman 
system  by  John  Frantz  and  Bj  om  My  sen  for 
examining  the  structures  of  silicate  liquids 
at  high  temperature.  Heretofore,  most 
Raman  studies  of  melts  actually  involved 
measurements  on  glasses  quenched  from 
high  temperatures.  Investigators  were 
forced  to  assume  that  the  glasses  were 
representative  of  melts  at  temperature,  but 
there  were  many  doubts  about  the  validity 
of  this  assumption.  Now,  Frantz  and  Mysen 
have  made  extensive  recordings  of  Raman 
spectra  on  silicate  melts  at  temperatures  as 
high  as  1600°C  and  it  appears  that  they 
have  opened  up  a  new  and  exciting  area  of 
melt  research. 

Douglas  Rumble,  Michael  Palin,  and 
Thomas  Hoering  have  developed  a  method 
for  laser  fluorination  of  sulfide  minerals 
with  fluorine  gas.  This  technique  allows 
fast  and  precise  in-situ  micro-sampling  on 
three  of  the  four  sulfur  isotopes,  32S,  33S, 
and  34S,  and  will  provide  information  on 
mass  transfer  in  sulfide  system  on  a  scale 
that  was  previously  inaccessible. 

In  addition  to  these  initiatives  related  to 
a  new  NSF  Center  and  to  new  instrumenta- 
tion, staff  members,  postdoctoral  fellows 
and  associates,  and  visiting  investigators 
have  been  involved  in  a  wide  range  of 
research,  ranging  from  racemization  dat- 
ing of  ostrich  shells  to  global  convection  of 
the  upper  mantle.  Much  of  this  research  is 
described  in  brief  summaries  in  this  Annual 
Report,  thus  continuing  the  Geophysical 
Laboratory  tradition  of  early  communica- 
tion of  results  before  full  papers  are  pub- 
lished in  scientific  journals. 


GEOPHYSICAL  LABORATORY 


Igneous  and  Metamorphic  Petrology 

Field  studies 


Global  Convection  and  Hawaiian  Upper 
Mantle  Structure 


T.  Neil  Irvine 

This  report  gives  further  development 
of  the  global  convection  system  proposed 
by  Irvine  (1989).  In  this  system,  upper 
mantle  convection  is  stratified  at  both  the 
400-  and  the  670-km  seismic 
discontinuities,  and  the  lower  mantle  fea- 
tures an  orthogonal  framework  of  six  prin- 
cipal convection  centers,  or  axes  where 
upwelling  occurs  beneath  Iceland,  Hawaii, 
the  Balleny  Islands  (near  Antarctica),  and 
the  Okavango  delta  (in  Botswana),  and 
downwelling  beneath  Peru  and  the  eastern 
edge  of  Vietnam.  In  last  year's  Report 
(Irvine,  Annual  Report  1989-1990,  p.  3- 
1 1),  the  concept  of  an  upper  mantle  vortex 
supercell  was  introduced  for  the  Iceland 
center  and  explored  on  the  basis  of  seismic 
data.  Application  of  this  concept  is  now 
extended  to  the  Hawaiian  center. 


Mantle  Tomography  and  Hotspots 

When  the  above  global-convection  sys- 
tem was  proposed,  some  support  was  cited 
from  seismic  tomography,  but  the 
tomographic  maps  then  available  portrayed 
only  broad  features  and  frequently  seemed 
incompatible.   More  recent  maps  are  not 


completely  consistent  either  (e.g.,  see  com- 
parisons made  by  Romano wicz,  1 99 1 ),  but 
a  set  by  Inoue  et  al.  (1990)  is  especially 
interesting  because  the  maps  are  unusually 
detailed.  Much  of  the  detail  correlates 
meaningfully  with  surface  geological  fea- 
tures (particularly  volcanic  hotspots),  so 
the  data  appear  significant.  Some  features 
are  notably  compatible  with  the  mantle 
convection  relationships  favored  here.1 

The  Inoue  et  al.  (1990)  maps  for  nomi- 
nal depths  of  478-629  km  and  1203-1435 
km  are  illustrated  in  Fig.  1 ,  together  with  a 
map  by  Woodhouse  and  Dzie  wonski  ( 1 989) 
for  150  km.  The  relations  for  1203-1435 
km  warrant  particular  attention  because  ( 1 ) 
they  are  dominated  by  two  major  zones  of 
low  velocity,  one  under  Hawaii,  the  other 
beneath  Okavango,  and  (2)  the  extensions 
of  these  two  zones  in  combination  with 
several  small,  low-velocity  anomalies,  en- 
compass most  of  the  world's  hotspots.  The 
two  major  anomalies  are  both  prominent  in 
several  earlier  tomographic  maps  (e.g., 
Giardini  et  al,  1987),  but  the  velocity  data 
in  these  cases  were  smoothed  to  low-order 
spherical  harmonics,  so  the  anomalies  are 
only  broadly  delimited.  The  anomaly  loca- 

1  In  interpretations  of  mantle  tomography  here  it 
is  assumed  simply  that  seismically  slow  regions 
are  relatively  warm  and,  thus,  may  represent  zones 
of  convective  upwelling,  whereas  fast  regions  are 
cooler  and,  therefore,  may  reflect  downwelling. 
Compositional  and  phase  differences  and  seismic 
anisotropy  are  likely  to  be  complicating  factors 
but  cannot  be  considered  here. 


CARNEGIE  INSTITUTION 


150  km 

Woodhouse  and  Dziewonski,  1989 
Model  U84L85/SH    Z  =1-8 


+  +  +  + 
+  +  +  + 
+  +  +  + 
+  +  +  + 

% 

§ 

-3% 


0 


3% 


478-629km 

Inoue  et  al.,1990 


:::::  + 

+  +  +  + 
+  +  +  + 
+  +  +  + 
+  +  +  + 

'//// 
VK 

± 

± 

± 

X 

i 

0 


2% 


GEOPHYSICAL  LABORATORY 


tions  on  the  Inoue  et  al.  maps  appear  much 
better  resolved. 

Neither  the  Iceland  nor  the  Balleny  cen- 
ter is  seismically  slow  at  1203-1435  km, 
but  Iceland  is  flanked  by  small  low-veloc- 
ity anomalies  on  the  northwest  and  south- 
east. Also,  the  Hawaii  and  Okavango 
anomalies  are  beltlike  at  this  depth  and  tend 
to  follow  the  great  circle  that  includes  Ice- 
land and  Balleny.  In  combination,  the  two 
belts  span  almost  half  the  Earth's  circum- 


ference. (A  rather  similar  anomaly  arrange- 
ment is  also  indicated  for  478-629  km  in 
Fig.  IB.) 

Several  small  anomalies  in  Fig.  1C  do 
not  correlate  with  hotspots,  and  a  few 
hotspots  are  not  associated  with  low  ve- 
locities (notably  Tristan  da  Cunha,  Fernando 
de  Norhana,  and  Martin  Vas,  all  in  a  South 
Atlantic  region  where  relatively  high  ve- 
locities are  indicated);  but  at  the  present 
state  of  the  science,  it  seems  more  signifi- 


1203-1435  km 
Inoue  et  al.,  1990 


Volcanic  hotspots 


FIG.  1.  Global  tomography  maps  for  nominal  depths  of  (A)  150  km,  redrawn  from  Woodhouse  and 
Dziewonski  (1989)  and  (B  and  C)  478-629  km  and  1203-1435  km,  redrawn  from  Inoue  etal.  (1990).  The 
maps  also  show  the  global  convection  framework  of  Irvine  (1 989). 

The  principal  correlations  between  low-velocity  anomalies  and  hotspots  in  Fig.  1C  are  as  follows.  The 
Hawaiian  anomaly  spreads  southward  beneath  the  Samoa,  Marquesas,  Tahiti,  and  Austral  McDonald 
hotspots  and  comes  close  to  Caroline  (northeast  of  New  Guinea).  A  strong  small  anomaly  underlies  the 
southeastern  Australia  hotspot,  and  weaker  ones  match  the  Galapagos  and  San  Felix  (south  of  Peru),  and 
possibly  Yellowstone.  In  the  North  Atlantic,  a  medium-sized  anomaly  is  surrounded  by  the  Azores, 
Canary  Islands,  Madeira,  Cape  Verde  Islands,  and  New  England  hotspots;  and  in  the  South  Atlantic,  a 
strong  small  anomaly  underlies  the  Cameroon  hotspot,  and  a  weak  one  matches  St.  Helena.  Where  the 
Okavango  anomaly  extends  into  northern  Africa,  it  and  a  small  satellite  are  associated  with  four 
continental  hotspots;  to  the  east  and  south,  the  Okavango  anomaly  encompasses  the  Comores,  Reunion, 
Marion  Island,  Crozet  Islands,  and  Kerguelen  hotspots. 


CARNEGIE  INSTITUTION 


cant  that  many  hotspots  are  matched.  A 
currently  popular  concept  is  that  hotspots 
are  initiated  by  mantle  plumes  originating 
in  the  D"  seismic  zone  just  above  the  core- 
mantle  boundary,  and  then  continue  to  be 
fed  from  this  zone  by  thin,  stemlike  chan- 
nels of  up  welling  (e.g.,  Olson,  1990).  Im- 
pressive geoid  and  tomographic  evidence 
for  a  source  region  at  the  core -mantle  bound- 
ary has  been  available  for  some  years  (e.g., 
Chase,  1979;  Crough  and  Jurdy,  1980; 
Woodhouse  and  Dziewonski,  1989),  but 
the  map  in  Fig.  1C  contains  what  may  well 
be  the  first  discernible  geophysical  indica- 
tions of  feeder  stems  at  intermediate  depths 
in  the  lower  mantle. 


Stratified  Mantle  Convection 

A  prominent  feature  of  the  three  maps  in 
Fig.  1  is  that  they  are  all  very  different;  in 
fact,  a  general  observation  from  seismic 
tomography  is  that  the  three  mantle  divi- 
sions delimited  by  the  400-  and  670-km 
seismic  discontinuities  tend  to  exhibit  con- 
trasting relations.  Although  rarely  cited, 
this  observation  would  seem  a  rather  strong 
argument  in  favor  of  at  least  some  stratified 
convection. 

In  the  mantle  convection  relationships 
proposed  by  Irvine  (1989),  it  was  assumed 
that  the  400-km  interface  is  everywhere 
mechanically  coupled,  and  a  combination 
of  thermal  and  mechanical  coupling  rela- 
tionships was  then  devised  for  the  670-km 
interface,  designed  to  account  for  the  gen- 
eral tectonic  features  of  the  Earth's  sur- 
face. An  unorthodox  feature  of  the  result- 
ing arrangement  is  that,  beneath  the  ob- 


served zones  of  upwelling  along  the  mid- 
ocean  ridges,  there  are  zones  of 
downwelling  in  the  depth  interval  400-670 
km.  Some  seismic  evidence  for  this  possi- 
bility was  cited  from  the  tomographic  maps 
of  Nataf  et  al.  (1986),  but  it  was  acknowl- 
edged to  be  equivocal.  The  relationships  of 
maps  A  and  B  of  Fig.  1  suggest  stronger 
evidence  for  this  possibility,  although  this 
evidence  too  is  not  beyond  question.  In 
particular,  the  maps  indicate  that,  while  the 
regions  beneath  the  mid-ocean  ridges  at 
150  km  (map  A)  are  generally  seismically 
slow  (as  expected),  those  at  478-629  km 
(map  B)  commonly  exhibit  relatively  high 
velocities.  This  contrast  is  especially  con- 
spicuous along  the  East  Pacific  Rise,  where 
it  is  most  relevant  to  the  Hawaiian  convec- 
tion relationships. 

But,  as  explained  by  Inoue  et  al.  ( 1 990), 
their  map  for  478-629  km  does  not  have  as 
high  resolution  as  that  for  1203-1425  km, 
and  they  specifically  stated  that  the  East 
Pacific  Rise  anomaly  in  the  former  is  not 
reliable.  This  does  not  mean  that  the 
anomaly  is  invalid,  however.  Inoue  et  al. 
noted  also  that  an  increase  of  seismic  ve- 
locity at  400  km  beneath  the  East  Pacific 
Rise  had  been  observed  by  Suetsugu  and 
Nakanishi  ( 1 987)  in  a  Rayleigh  wave  study; 
and  maps  by  Dziewonski  and  Woodhouse 
(1987)  of  S-wave  velocity  variations  at  the 
670-km  discontinuity  similarly  show  high 
velocities  beneath  the  Rise  south  of  the 
Galapagos.  Thus,  the  possibility  of  sub- 
ridge  downwelling  at  400-670  km,  although 
still  wanting  of  strong  support,  is  still  con- 
sidered viable. 


GEOPHYSICAL  LABORATORY 


7 


Hawaiian  Relationships  and  Supercell 
Structure 

It  is  well  established  that  the  islands  and 
seamounts  of  the  Hawaiian  and  Emperor 
volcanic  chains  become  progressively  older 
to  the  west  and  north  from  Hawaii.  The 
widely  accepted  explanation  is  that  the 
volcanoes  formed  in  succession  from  a 
relatively  fixed  mantle  hotspot,  currently 
located  beneath  the  Big  Island,  as  the  Pa- 
cific plate  drifted  first  northward,  then 
westward  across  it  (cf.  Clague  and 
Dalrymple,  1987).  The  recent  volcanism 
on  Hawaii  has  been  dominated  by  the  tho- 
leiitic  eruptions  of  Kilauea  andMauna  Loa, 
but  the  newest  activity  features  eruptions 
of  alkalic  basalt  from  the  submarine  vol- 
cano Loihi  on  the  south  edge  of  the  island. 
Along  the  older  parts  of  the  volcanic  sys- 
tem, the  seamounts  at  the  Hawaiian-Em- 
peror bend  are  42-43  Ma  in  age;  those  at  the 
north  end  of  the  Emperor  chain  are  73-75 
Ma. 

In  1972,  Jackson  et  al.  pointed  out  that 
the  Hawaiian-Emperor  volcanoes  tend  to 
be  paired,  and  portrayed  them  as  being 
distributed  along  an  en  echelon  (discon- 
tinuous) series  of  sigmoidal  double  lines. 
Recently,  Garcia  et  al.  (1990)  identified  a 
newly  discovered  submarine  volcano 
(named  Makuhona)  just  west  of  Hawaii  to 
be  the  previously  missing  partner  of  the 
volcano  Kohala  on  the  northern  peninsula 
of  the  island.  With  these  observations  as 
background,  I  have  attempted  to  pair  the 
volcanic  structures  of  the  entire  system  into 
a  set  of  more  continuous  double  lines,  as 
shown  in  Fig.  2.  En  echelon  overlap  was 
required  through  the  Midway  Islands,  but 


otherwise  only  two  lines  were  necessary. 
Inasmuch  as  the  only  control  was  small- 
scale  topography,  the  pairing  is  frequently 
conjectural,  but  in  an  overall  count,  per- 
haps 60  of  some  70  possible  pairs  appear 
credible.  Thus,  given  that  there  are  prob- 
ably still  other  unrecognized  eruption  cen- 
ters of  the  Makuhona  type,  the  more  con- 
tinuous double  lines  seem  realistic.  The 
contention  is  that  they  reflect  the  continu- 
ous operation  of  the  mantle  convection 
supercell  outlined  in  parts  B  and  C  of  Fig. 
2  and  explained  below. 

Figure  2  also  shows  the  ocean-floor 
troughs  and  arches  that  are  associated  with 
the  volcanic  chains.  The  main  Hawaiian 
trough  (here  termed  the  "inner  trough") 
encircles  the  south  side  of  the  Big  Island 
and  extends  discontinuously  westward  on 
both  sides  of  the  island  chain-,  with  widths 
locally  exceeding  100  km.  This  trough  has 
long  been  ascribed  to  loading  of  the  ocean 
floor  by  the  volcanic  ridge  (e.g.,  Moore, 
1987).  Outboard  from  it  are  broad  arches, 
or  swells,  300-400  km  wide.  They  are 
topographically  prominent  only  along  the 
younger  half  of  the  Hawaiian  chain,  where 
their  relief  exceeds  1 000  m  and  brings  them 
to  depths  less  than  5000  m  (Fig.  2);  but  they 
can  be  readily  identified  throughout,  even 
along  the  Emperor  chain,  by  their  distinc- 
tive positive  gravity  signatures  (see  Haxby, 
1987).  Finally,  fringing  the  swells  around 
the  younger  part  of  the  Hawaiian  chain  is  a 
subtle  "outer  trough."  This  trough  has 
special  importance  in  the  present  context 
(see  below). 

Figure  3  depicts  an  Hawaiian  upper 
mantle  convection  supercell  designed  to 
account  for  the  paired  volcano  chains  and 


8 


CARNEGIE  INSTITUTION 


FIG.  2.  Maps  of  the  Hawaiian  and  Emperor  island  and  seamount  chains  in  which  the  volcanic  structures 
are  paired  into  two  lines  (four  lines  through  the  Midway  Islands).  Based  on  the  topography  map  of 
Mammerickx  (1989),  the  ocean  gravity  map  of  Haxby  (1987),  and  maps  of  earthquake  epicenters  and 
young  volcanoes  from  Moore  (1987)  and  Garcia  etal.  (1990).  See  text  for  discussion  of  the  upper  mantle 
convection  supercell. 


GEOPHYSICAL  LABORATORY 


WNW 


EAST  PACIFIC  RISE 


Mechanical  coupling 


Tholeiitic  shield  volcano 
Early  alkalic  lavas 
nner  Trough 
Outer  trough 

0  KM 


PLAN  VIEWS 
Vortex  core  flow 

©  up 

0  down 


11111*11  in  i  400 


FIG.  3.  Schematic  three-dimensional  representation  of  the  Hawaiian  upper  mantle  vortex  supercell.  See 
text  for  description. 


related  features.2  A  principal  feature  of  the 
interpretation  is  the  differential  lateral  flow 
of  the  upper  mantle  layers.  As  with  the 
supercell  proposed  last  year  for  Iceland, 
the  vorticity  in  the  supercell  is  ascribed  to 
thermal  tilting  of  this  flow  by  heat  rising 
from  the  underlying  lower  mantle  axis  of 
convective  up  welling  (cf.,  Irvine,  1990, 

2  Garcia  et  al.  (1990)  noted,  on  the  basis  of  an 
experimental  study  of  mantle  plume  dynamics  by 
Richards  and  Griffiths  (1989),  that  a  convection 
structure  of  this  general  type  might  account  for  the 
pairing  of  the  Hawaiian  volcanoes;  and  Sleep 
(1990)  used  a  "stagnation  streamline"  model  akin 
to  part  of  the  structure  in  Fig.  3  to  estimate  buoy- 
ancy flux  values  for  the  Hawaiian  and  other  mantle 
plumes. 


Fig.  1).  The  differential  flow  also  leads  to 
the  downwelling  at  400-670  km  under  the 
East  Pacific  Rise,  as  discussed  in  relation  to 
Fig.  2B. 

In  the  envisaged  action  of  the  supercell, 
a  swath  of  the  lower  part  of  the  lithosphere 
is  stripped  away  (delaminated)  and  replaced 
by  asthenospheric  material  upwelling  from 
near  the  400-km  interface.  The  stripping 
process  depresses  the  outer  ocean-floor 
trough,  and  the  buoyancy  of  the  replace- 
ment material  elevates  the  topographic 
swells.  The  swells  eventually  subside  as 
the  replacement  material  gradually  cools 
and  contracts,  but  through  its  increased 


10 


CARNEGIE  INSTITUTION 


density,  they  continue  to  have  their  distinc- 
tive gravity  signatures.  Further  postulates 
are  (1)  that  the  double -vortex  circulation 
holds  the  supercell  in  place  as  a  standing 
vortex  against  the  laterally  flowing  mantle 
layers  (see  Irvine,  1990,  Fig.  1),  (2)  that 
magma  melting  occurs  mainly  through  the 
adiabatic  rise  of  fertile  peridotite  along  the 
two  vortex  axes  of  upwelling,  and  (3)  that 
these  two  axes  also  deliver  the  magma  to 
paired  release  points  at  the  base  of  the 
lithosphere,  from  where  it  rises  to  the  paired 
volcanoes  at  the  surface. 

The  supercell  structure  can  also  be  used 
to  rationalize  much  of  the  general  history 
of  individual  Hawaiian  volcanoes  (for  sum- 
mary, see  Clague  and  Dalrymple,  1987). 
By  concept,  the  magma  from  the  main 
release  points,  having  melted  at  relatively 
high  pressures,  is  picritic  tholeiite  in  com- 
position, and  its  eruption  produces  large 
shield  volcanoes  like  Mauna  Loa  and 
Kilauea.  But  it  can  be  postulated  too  that 
some  alkalic  magma  should  form  in  the 
surrounding,  spreading  parts  of  the  vortex 
cells  by  differentiation  of  trapped  intersti- 
tial melt  at  the  more  moderate  pressures  of 
this  environment.  This  magma  could  be 
released  both  in  advance  of  the  tholeiitic 
shields,  as  at  Loihi,  and  in  arrears  of  them, 
as  in  the  "alkalic  post-shield  stage"  ob- 
served in  volcanoes  such  as  Mauna  Kea 
and  Hualalai.  I  suggest  too  that,  through  its 
buoyancy,  this  alkalic  magma  may  tend  to 
accumulate  in  substantial  quantities  at  the 
tops  of  the  inferred  vortex  axes  of 
downwelling,  from  where  it  might  be 
erupted  at  a  considerably  later  stage  in  the 
history  of  a  volcano  if  the  volcano  hap- 
pened to  pass  across  such  an  axis  by  virtue 


of  the  plate  motion.  This  kind  of  late 
eruption  could  correspond  to  the  "alkalic 
rejuvenated  stage"  that  is  observed  in  most 
of  the  Hawaiian  volcanoes  between  Maui 
and  Necker  Island. 

A  fundamental  general  tenet  in  this 
analysis  is  that  magma  from  the  mantle  is 
primarily  released  to  the  lithosphere  from 
specific,  critical  flow  points  in  the 
asthenospheric  convection  system. 


References 

Chase,  C.  G.,  Subduction,  the  geoid,  and  lower 
mantle  convection:  Nature,  282,  462-468, 
1979. 

Clague,  D.  A.,  and  G.  B.  Dalrymple,  The  Hawai- 
ian-Emperor volcanic  chain:  Part  I.  Geologic 
evolution,  U.  S.  Geol.  Surv.  Prof.  Paper  1350, 
5-54,  1987. 

Crough,  S.  T ,  and  D.  M.  Jurdy,  Subducted  litho- 
sphere, hotspots,  and  the  geoid,  Earth  Planet. 
Sci.  Letters,  48,  15-22.  1980. 

Dziewonski,  A.  M.,  andWoodhouse,  J.  H.,  Global 
images  of  the  Earth's  interior,  Science,  236, 
37-48,  1987. 

Garcia,  M.  O.,  M.  D.  Kurz,  and  D.  W.  Muenow, 
Mahukona:  The  missing  Hawaiian  volcano, 
Geology,  18,  1111-1114,  1990. 

Giardini,  D.,  L.  Xiang-Dong,  and  D.  H. 
Woodhouse,  Three-dimensional  structure  of 
the  Earth  from  splitting  in  free-oscillation  spec- 
tra, Nature,  325,  405-41 1,  1987. 

Haxby,  W.  F.,  Map  of  the  gravity  field  of  the 
world's  oceans,  Natl.  Oceanic  Atmos.  Adm. 
Rpt.  MGG-3,  1987. 

Inoue,  H.,  Y.  Fukao,  K.  Tanabe,  and  Y.  Ogata, 
Whole  mantle  P- wave  travel  time  tomography, 
Phys.  Earth  Planet.  Interiors,  59,  294-328, 
1990. 

Irvine,  T.  N.,  A  global  convection  framework: 
concepts  of  symmetry,  stratification  and  sys- 
tem in  the  Earth's  dynamic  structure,  Econ. 
Geol,  84,  2059-21 14,  1989. 

Jackson,  E.  D.,  E.  I.  Silver,  and  G.  B.  Dalrymple, 
Hawaiian-Emperor  chain  and  its  relation  to 
Cenozoic  Circum-Pacific  tectonics,  Geol.  Soc. 
Amer.  Bull.,  83,  601-618,  1972. 

Mammerickx,  J.,  Bathymetry  of  the  North  Pacific 
Ocean,  Geol.  Soc.  Amer.,  The  Geology  of 
North  America,  N,  1989. 


GEOPHYSICAL  LABORATORY 


11 


Moore,  J.  G.,  Subsidence  of  the  Hawaiian  Ridge, 
U.  S.  Geol.  Surv.  Prof.  Paper  1350,  85-100, 
1987. 

Nataf,  H.-C,  I.  Nakanishi,  and  D.  L.  Anderson, 
Measurements  of  mantle  wave  velocities  and 
inversion  for  lateral  heterogeneity  and  anisot- 
ropy:  III,  Inversion,  /.  Geophys.  Res.,  91, 
7261-7308,  1986. 

Olson,  P.,  Hot  spots,  swells  and  mantle  plumes,  in 
M.  P.  Ryan,  Magma  Transport  and  Storage, 
New  York,  J.  Wiley  &  Sons,  33-51,  1990. 

Romano wicz,  B.,  Seismic  tomography  of  the 
Earth's  mantle,  Ann.  Rev.  Earth  Planet.  Sci., 
79,77-99,  1991. 

Sleep,  N.,  Hotspots  and  mantle  plumes:  some 
phenomenology,  /.  Geophys.  Res.,  95,  6715- 
6736,  1990. 

Suetsugu,D.,andI.Nakashini,  Three-dimensional, 
velocity  map  of  the  upper  mantle  beneath  the 
Pacific  Ocean  as  determined  from  Rayleigh 
wave  dispersion,  Phys.  Earth  Planet.  Interi- 
ors, 47,  205-229,  1987. 

Woodhouse,  J.  H.,  and  A.  M.  Dziewonski,  Seis- 
mic modelling  of  the  Earth's  large  scale  three- 
dimensional  structure,  Phil.  Trans.  R.  Soc. 
Lond.,  A  328,  291-308,  1989. 

Richards,  M.  A.,  and  R.  W.  Griffiths,  Thermal 
entrainment  by  depleted  mantle  plumes,  Na- 
ture, 342,  900-902,  1989. 


Megacrystalline  Dunites  and 

Peridotites: 
Hosts  for  Siberian  Diamonds 

N.  P.  Pokhilenko* ,  D.  G.  Pearson  ** , 
F.  R.  Boyd,  andN.  V.  Sobolev* 

Several  investigations  have  identified 
xenoliths,  consisting  primarily  of 
ultracoarse  crystals  of  olivine,  which  ap- 
pear to  be  fragments  of  the  principal  host 
rocks  of  Siberian  diamonds  (Sobolev,  1974; 
Pokhilenko  et  ai,  1977;  Soboley  et  al., 
1984).    Twenty -three  of  these  xenoliths 

Inst,  of  Mineralogy  &  Petrology,  Siberian  Branch 
of  the  USSR  Academy  of  Sciences,  Novosibirsk 
Dept.  of  Terrestrial  Magnetism,  Carnegie  Instn. 
of  Washington,  Washington,  D.  C.  20015. 


contain  diamonds,  and  many  more  contain 
garnets  with  a  compositional  range  very 
similar  to  the  range  for  garnets  included  in 
diamonds.  These  megacrystalline  rocks 
have  been  found  as  xenoliths  in  the 
kimberlites  of  the  Daldyn-Alakit  region 
and  in  some  pipes  of  the  Upper  Muna 
group.  They  are  especially  abundant  in  the 
xenolith-rich  Udachnaya  kimberlite,  site 
of  one  of  the  world's  richest  diamond  mines. 

The  Siberian  megacrystalline  rocks  dif- 
fer from  diamondiferous  peridotites  from 
southern  Africa,  the  only  other  region  in 
which  a  number  of  such  xenoliths  have 
been  found.  African  occurrences  are  pri- 
marily lherzolites  and  harzburgites  having 
garnets  and  associated  minerals  with  com- 
positional ranges  that  depart  significantly 
from  the  majority  of  diamond  inclusions 
(Viljoen  etal.,  1991;  Boyd  etal.,  in  prepa- 
ration). Thus  the  principal  host  rocks  for 
African  diamonds  have  not  yet  been  dis- 
covered as  articulated  xenoliths. 

The  relative  abundances  and  composi- 
tional ranges  of  the  olivine  and  associated 
minerals  indicate  that  most  of  the  Siberian 
megacrystalline  rocks  are  extremely  re- 
fractor)'. They  may  be  residues  or  cumu- 
lates of  melting  events  in  which  a  large 
proportion  of  melt  was  removed.  Olivine 
forms  more  than  90%  of  most  specimens. 
The  primary  modal  proportions  are  diffi- 
cult to  estimate  because  of  disaggregation 
during  eruption,  but  it  is  likely  that  these 
ultracoarse  rocks  were  olivine-rich.  The 
molar  Mg/(Mg+Fe)  for  the  olivine  is  pre- 
dominantly 0.92-0.95,  and  most  of  the  gar- 
nets are  strongly  subcalcic  and  Cr-rich  (Fig. 
4).    Mineral  assemblages  of  the  33  new 


12 


CARNEGIE  INSTITUTION 


Table  1.  Mineral  assemblages  in  megacrystalline 
olivine-rich  xenoliths  from  the  Udachnaya 
kimberlite,  U.S.S.R. 


Assemblage 


No.  specimens 


olv  +  gar 

olv  +  gar  +  chr 

olv  +  gar  +  diam 

olv  +  gar  +  cpx 

olv  +  gar  +  opx 

olv  +  chr 

olv  +  opx  +  chr 

olv  +  gar  +  chr  +  diam 

olv  +  gar  +  opx  +  cpx  +  chr 


13 

8 

4 

2 

2 

1 

1 

1 

1 


specimens  chosen  for  this  investigation 
have  a  predominance  of  olivine  +  garnet 
(Table  1).  Five  are  diamond-bearing. 


Isotope  Results 

Preliminary  neodymium  isotope  analy- 
sis of  two  hand-picked,  acid-washed  garnet 
separates  (Uv  70/76  and  Uv  49/76)  reveal 
that  the  megacrystalline  rocks,  having  ex- 
perienced an  ancient  trace  element  enrich- 
ment event,  show  similar  Nd  isotopic  sig- 
natures to  peridotite  suite  garnet  inclusions 
in  diamond  from  southern  Africa  analyzed 
by  Richardson  et  al.  (1984).  The  two 
samples  record  £Nd  values  of  -28  to  -33  at 
350  Ma  (the  time  of  pipe  emplacement), 
and  yield  model  ages  relative  to  the  Bulk 
Earth  reservoir  (CHUR)  of  2.0  to  2.7  Ga. 
(Fig.  5).  Although  the  latter  age  is  late 
Archaean,  it  is  substantially  younger  than 
the  3.3  Ga.  model  ages  obtained  for  the 
diamond  inclusions  by  Richardson  et  al. 
(1984).  The  younger  model  ages  and  the 
age-range  obtained  from  the  two  samples 
from  Udachnaya  may  be  attributed  to  the 
more  "open  system"  behavior  of  the  small, 
coarse-grained  megacrystalline  samples 


o 

CO 

O 


12 

10 

5      8 

6 

4 
2 


— ■ r 

.      o1 

— r— 

■-> 1 ■ — 

- 1 ■ 1 *• 

1 1 

o     °    . 

a2 

•  3 

.        B4 

o 

•^^^s* 

0^ 

- . 

• 

• 

□           _ 
°  o* 

°  %° 

o       •  % 

o 

■ 
• 

•      111 

• 

3       o° 

•  ■    ■ 
c 

1 

_1        .        1        . 

I 

2  4  6  8  10 

Cr203,wt% 


12 


14 


Fig.  4.  A  plot  of  CaO  against  O2O3  for  garnets 
from  megaciystalline  xenoliths  of  different  asso- 
ciations: 1  -  without  chromite;  2  -  with  chromite; 
3  -  diamond-bearing  without  chromite;  4  -  dia- 
mond-bearing with  chromite.  Data  from  this 
study  and  Sobolev  et  al.  (1984). 

(compared  with  the  armored  diamond  in- 
clusions), which  could  have  allowed  the 
xenolith  garnets  to  be  subjected  to  later 
interaction  with  a  lighter  rare-earth-enriched 
component.  Further  analyses  of  a  suite  of 
larger  garnets  from  the  megacrystalline 
rocks  are  being  undertaken  to  resolve  this 
problem.  However,  the  results  obtained 
indicate  that  the  base  of  both  the  Kaapvaal 


0.512 


0       1000    2000     3000  4000 

Time,  Ma 

Fig.  5.  Neodymium  isotope  evolution  diagram  for 
two-garnet  separates  from  the  Udachnaya 
megacrystalline.  Vertical  arrows  indicate  model 
ages  relative  to  the  Bulk  Earth  or  CHUR  evolution 
curve.  Diamond  symbols  represent  the  isotopic 
composition  of  the  samples  at  the  time  of  pipe 
emplacement  350  Ma  ago. 


GEOPHYSICAL  LABORATORY 


13 


and  Siberian  cratons  experienced  similar 
incompatible  element  enrichment  events 
very  early  in  their  evolution. 


Petrography 

The  megacrysts  are  friable  and  are  not 
easily  separated  intact  from  hard  kimberlite. 
The  largest  has  a  long  dimension  of  19  cm, 
but  most  specimens  are  ovoidal  with  di- 
mensions of  the  order  of  5  cm.  Olivine 
crystals  range  up  to  10  cm,  and  in  a  few 
specimens  smaller  grains  of  olivine  (6-9 
mm)  with  differing  optic  orientation  are 
interspersed.  Prominent  parting,  resem- 
bling cleavage,  is  characteristic.  Rounded 
garnets  are  much  finer-grained,  0.2-6  mm, 
and  form  up  to  5  modal  percent.  The 
predominant  low-Ca  garnets  are  lilac -purple 
but  those  that  are  rich  in  Ca  as  well  as  Cr  are 
dark  purple  or  gray  or  even  green.  Alter- 
ation of  garnet  to  kelyphite  is  variable; 
some  grains  with  minor  kelyphite  are 
euhedral  with  a  rhombic  dodecahedral 
morphology  complicated  by  uneven  devel- 
opment of  faces.  The  distribution  of  gar- 
nets is  relatively  regular  in  the  larger  xeno- 
liths. 

Enstatite  is  present  in  about  20%  of  the 
megacrystalline  rocks,  but  clinopyroxene 
is  much  less  common.  Octahedra  of  chro- 
mite,  0.1-2  mm,  form  less  than  1  modal 
percent,  and  some  of  them  appear  to  show 
signs  of  resorption. 

Crystals  of  diamond  in  these  rocks  are 
characteristically  clear,  shaip-edged  octa- 
hedra, some  with  spinel  twins,  and  range  up 
to  4  mm.  Colored  diamonds  and  those  that 
are  corroded  or  cracked  are  unusual.  Small 


plates  of  graphite  (0.2-1.5  mm)  have  also 
been  observed  both  with  and  without  dia- 
mond. The  abundances  of  diamond  and 
graphite  are  usually  insignificant  but  one 
olivine-pyrope  specimen  having  a  volume 
of  only  2.5  cm3  contains  four  diamonds 
together  with  cavities  from  which  two  ad- 
ditional diamonds  have  broken  away. 


Mineral  Chemistry 

Most  olivines  in  the  megacrystalline 
rocks  have  Mg/(Mg+Fe)  in  the  range  0.92 
-0.95,  but  two  olivine-rich  wehrlites  in 
which  the  garnets  have  both  high  Ca  and 
high  Cr  (Fig.  4)  contain  substantially  more 
Fe-rich  olivines,  0.87  -  0.89  (Fig.  6).  Sur- 
prisingly, a  variation  in  NiO  from  0.32  to 
0.40  does  not  correlate  with  mg  number. 
Values  for  CaO  and  Cr203  in  the  olivines 
do  not  exceed  0.06  wt  %. 

Garnets  in  the  megacrystalline  xeno- 
liths  can  be  assigned  to  three  parageneses 
on  the  basis  of  their  CaO  and  Cr203  con- 
tents: proportions  for  more  than  200  speci- 
mens thus  far  studied  (this  Report  and 
Sobolev  et  al.,  1984)  are  harzburgite-dun- 
ite  (80%),  lherzolite  (15%),  and  wehrlite 
(5%).  Concentrations  of  FeO  andTi02  are 
least  in  the  harzburgite-dunite  group  (Tables 
2  and  3).  There  is  a  positive  correlation 
between  Ti02  and  CaO  in  these  garnets, 
apparently  reflecting  degree  of  depletion 
(Fig.  6B).  The  chromites  are  predomi- 
nantly rich  in  Cr203  (58-65  wt  %)  but 
spinels  with  Cr203  as  low  as  21.9  wt  % 
have  been  found.  A  good  correlation  in  Cr/ 
(Cr  +  Al)  between  the  garnets  and  chro- 
mites (Fig.  6C)  is  evidence  of  equilibra- 


14 


CARNEGIE  INSTITUTION 


a 
•  i— < 

*—> 

■a 

>* 

CX 
•  i— i 

CX 


t-l 
X> 


OS 

>> 

a 

c 
xi 
o 
a 

X3 


O 

OO 

.C 

4— > 

•  r-< 

c 
t> 

X 

<— > 

J— > 

o 

■c 

ex 

D 

c 

00 

p 

S3 

a 

> 

•  I— I 

a 

•*— > 
c 
<u 

oo 
<U 

ex 
u 

t-c 

oo 
t> 
oo 
j>. 

13 
c 
ed 

<u 

c 
•  i—i 


CN 

a 

£0 
< 


no 

oo 

o" 
no 

I 

> 
!D 


NO 
oo 

CN 

NO 

> 
ID 


NO 
OO 

in 

CN 

i 
> 

!D 


NO 
oo 

NO~ 

CN 

I 

> 

ED 


bfl 


J3 

o 


X 

a, 
o 


53 


CX 

o 


J3 

bJQ 


X 

Oh 

O 


> 
O 


o 


ton 


X 

ex 
o 


incococoTtoouoioco© 

mO'— i'— i  on  ©  *— t^^OO 


CN  Tf  CN  no  r-  no 


o  o  o 


oo 


en 


TfOOO-HO^tOOO 


r-»  OO  OO  OO  On  Os  c- 


©  CO  O 


oo  oo 


-a  -a 


©  ©  On  »0  *— i  ©  ^h  ©    c    c 


^••oiovp-'^ooNior^ 


m  d  ^'  ri  -h  o 


mo    .r 


©  t—  no  no  co  on  <— no    -g 
-4  -1  vd  ^  °i  "1  od  "I  ^.  H 


lOOOO^OcoOO   c 


^©©CNCN^hNOCN-CN 

^•OOOvO-HqO^CO 

Tj-ooor^o'mo  c  © 


^-*h^tOOO>^ 


oo 


CN  Tf  ^t 


<N 


CN 


NO  r-  O  CN  <N  o  o 


8rHlOOmHOOON 
<    CO    T-H    Tt    T— <    ^    r- I    ©    <— J 

loddd^dc^ddd 


en 

od 


On 


O  CO  On  CO  O  NO 


O  Tt 


O  O  CO 


Ttooor^-ou-iooo 


NO  NO  NO  NO  CN  On 
'-1  P  <*"*  -J  00*  "* 

o  o  r-  ^o  ^  © 


d  c  d 


coeoio^iocN©^ 


^t  O 


oo 


oo  Tt  >o 


CN 


NO 


NO  r^  O  CN  CN  o  o 


— '  On  ^t  NO  r-  <0  CN 


oo 


O^MCS 


m 


cn  o  ^ 


mOOOrJ-OcOOO    CS 


hOOOhh  ^_; 

^t  ©  ©  ©  r^  ©  uo 


Pd^ 
o  o  o 


KP<0£SSozz 


oo09,0 


C  BO 


d   cs  ^5 


oo 

r^ 

od 

ON 
ON 

d 
o 


in 

od 

On 


CN 


On 
CN 


On 
O 


On 

O* 

On 


On 
NO 


CO 

On 
On 


On 
CO 
On 


NO 
CN 

On 
On 


OO 


On 
On 


CO 


4—1 

o 
H 


CN  ^f  On 
ih  d  u-i 

r-  co  io 


NO 


OO 

OO 


CO 


CO 


oo 


CN  NO  OO 

"3*  no  »n 

On  tJ-  tj- 


CO  NO  iO 
O  od  iri 
OO  '-h  CN 


r.NO 


CO 

On 


CN 

CN 

On 


On 
CN 


">  NO  "^ 


5  lO 

OO  V) 


oo 


rn  -:  Tt 

On  °  CN 


lO 

CN 

On 


r-    On 

»o   OO 


^  OO 


oo 
On 


CO 

On 


CN 

On 


3  CN 


£3^ 

+  2  <- 

Sou 


GEOPHYSICAL  LABORATORY 


15 


•i— i 

*-> 

M 

Oh 
•  i— < 


X) 


a 

o 

a 

.13 

■4— > 

£ 
o 

<J3 

+- » 

•H 

c 

<D 

X 

<U 
■*-> 

•4— > 

O 

•c 

<u 
ex 

00 

C 


X> 

I 

C 

o 

B 


c 


o 
aj 

oo 


c 
< 


< 


ON 

oo 
O 

i 
> 


On 

oo 

r- 


> 

S3 


vo 

oo 

io 

VO 

I 

> 


vo 
oo 

^? 

On 
co 

i 

> 

S3 


vo 
oo 

vS1 

I 

> 

S3 


00 


53 


53 

00 


00 


I— * 

o 


J3 


J3 
oo 


OX) 


a^tMCOIOMOrH 


-JP 
^t  o 


vo 


oo 


On  «t 


<N 


r-~  vo  o  <n  co  o 


""t  m  tJ- tJ- Tf  \©  ^  ro  (N 


CN 


lOrHCO 


co 


OO 


Tj-ocN-^tr-ocNcNO 


On  O 


>-Hi>(sa\M      vo 
Ohh  jo       co 


^ooor-oino 


covor-»ocNcNO'— i 


eN 


ON 


CO  On  "tf  ^  CO  O 


TtO-HiONOONmO 


»— lQO'— iCIr-Hf—CO 

H'popcjHjq 
Tj-ooor^o>oo 


ON 
CO 


Tj-Q<— lOOVOOOVOCN 
Tt©— ivOUOOcNOO 


(N  OO  O  O  CO 


O  vo 


CO 


vo 

CO 


^ooo«oo»oo 


CN  CO  «o  OO  CO  CO 


^j  O  o  o  r-  o 


CN 


^tooovoomo 


O  <N  m  cn  oo  t-«-  oo 


CN  *-<  VO 


o  »o 


CN 


OOMnOhQhO 


cN^tcNooT-Hr^.r«-»o^HQ 

(vj  O  oo  CN  CO  CO  ^  <N  ©  © 

rtO'-nr-VOOCN'-HOO 


t--<Nr'»r--vo,r+<— *on 


cn 

tJ-  o 


oo 


CO  CN 

\6  vd 


^  (sqq 

<N  ">«  o  O 


VO 


i— i  CN  CO 

OON 


VO 


lO 

CO 


^tooovoo»oo 


m  en 

OOnnO _ 

co  H  <  <J  pu  ,§  S  <J  £  £ 


oo0§o 


a  oo 


o 


CN 


<—>  w  p  ©  CO  i— <  J  p  CO  © 

^t  o  o  o  r^  o  m  o      ©      2 


CN 


OO 

© 
o 


oo 

ON 

On 
On 


VO 


CN 

ON 

ON 
ON 


CO 


O        « 


>o 


vO 
CO 


o      Jr; 


CO 

oo 

ON 


VO 
CO 

On" 
On 


VO 

OO 

ON 
ON 


On 
On 

ON 


c3 

o 


oo 


VO 
CN 

ON 


CO 

oo 


oo 

CN 

ON 


VO 
OO 


VO 

CN 

On 


oo 
oo 


ON 


CO 

CO 

On 


CO 
ON 

iO 


oo 


CO 

oo 


vq 

CO 

ON 


+ 
00 


§     ^fc 


16 


CARNEGIE  INSTITUTION 


0.42 

0.40 

0.38h 

0.36 

0.34 


0.32 


' r      •        i 

■  a.  olivine 

— i 1 1 1 r 

■ 

- 

oo  • 

- 

■ 

a  •         a 

• 

. 

o    ■ 

■ 

- 

OOXD      ■  • 

- 

o 

0  •  <»«■ 

■ 

- 

*m   m 

" 

o 

m    m 

_J U . 1 L 

■ 

86       88       90       92       94       96 

mg# 


id 

■  b.  garnet 

— i ■ — 

8 

10 

; 

8 

- 

• 

o 

6 

o 
o        o 

o 

■        o-        0          o 

4 

•    °- 

- 

.    °o    *      o 

. 

•     o 

2 

"a        • 

i.i. 

1 

0.0  0.1  0.2 

Ti02,  wt  % 


0.3         0.4 


5U 

■    C. 

T ' 

1 •" 

1 ' 

40 

• 
■ 

30 

■  "■    . 

20 

• 

10 

"  • 

• 
1             ■ 

• 
1             i 

20  40  60  80 

Cr/(Cr+AI),  %,  chr 


100 


Fig.  6.  Compositional  variations  in  minerals  of  the 
megacrystalline  rocks.  Symbols  as  in  Fig.  4.  a)  a 
plot  of  NiO  against  Mg  number  for  olivines  of 
megacrystalline  rocks,  b)  a  plot  of  CaO  against 
Ti02  for  garnets  from  megacrystalline  rocks,  c)  a 
plot  of  Cr/(Cr  +  Al)  ratio  for  garnets  against  Cr/(Cr 
+  Al)  ratio  for  chromites  from  megacrystalline 
rocks. 


700 


30  40  50  60 

Pressure,  kbar 


70 


Fig.  7.  Temperature-Pressure  estimates  for  four 
xenoliths  of  megacrystalline  rocks  from 
Udachnaya  pipe  having  coexisting  pyrope  and 
enstatite.  Temperature  was  calculated  with  the 
olivine-garnet  thermometer  (O'Neill  and  Wood, 
1979)  and  pressure  from  the  isopleths  of 
MacGregor(1974). 

tion.  Enstatites  have  low  AI2O3  (Table  2), 
reflecting  relatively  low  temperatures  and 
high  pressures  of  crystallization. 


Thermobawmetry 

The  temperature  of  equilibration  of  a 
xenolith  can  be  estimated  from  the  parti- 
tion of  Fe  and  Mg  between  olivine  and 
garnet  (O'Neill  and  Wood,  1979),  but 
enstatite  must  also  be  present  to  obtain  an 
estimate  of  the  pressure  or  depth  of  equili- 
bration. Four  of  the  megacrystalline  rocks 
contain  enstatite  as  well  as  olivine  and 
garnet  and<F-r  estimates  for  them  (Fig.  7) 
plot  close  to  or  within  the  diamond  stabil- 
ity field. 

Puzzling  features  of  the  P-Tplot  in  Fig. 
7,  however,  are  that  the  estimated  points  for 
megacrystalline  rocks  show  a  wide  disper- 
sion in  pressure  and  that  they  plot  in  a 
temperature  range  below  the  shield 
geotherm  (40  mWm^)  and  below  estimates 


GEOPHYSICAL  LABORATORY 


17 


made  for  garnet  lherzolite  xenoliths  from 
Udachnaya  (Boyd,  1984).  One  of  the 
megacrystalline  rocks  contains  diopside  as 
well  as  olivine  and  garnet  (Uv-624/86,  Table 
2),  and  its  temperature  can  be  estimated 
with  the  diopside  solvus,  giving  800°C,  as 
well  as  with  the  Gar/Olv  thermometer, 
which  gives  840°C.  The  approximate  agree- 
ment suggests  that  the  discrepancy  with  the 
lherzolite  data  may  not  be  a  failure  of 
thermobarometry.  The  dispersion  in  pres- 
sure for  the  megacrystalline  rocks  could  be 
evidence  of  a  wide  range  in  depth  of  equili- 
bration. More  data  are  clearly  needed, 
however,  before  these  questions  can  be 
properly  addressed. 

Experimental  studies  of  coexisting  solid 
solutions  of  spinel  and  garnet  show  that  the 
solubility  of  the  knorringite  component 
(Mg3Cr2Si30i2)  in  pyrope  increases  with 
increasing  temperature  and  pressure 
(Malinovsky  andDoroshev,  1975).  Pyrope 
garnet  that  is  rich  in  Cr  can  coexist  with 


o 

03 
O 


4  6  8  10 

O2O3  ,Wt.% 


12         14 


Fig.  8.  A  plot  of  CaO  against  Q2O3  with  isopleths 
of  knorringite  component  in  garnet  estimated  from 
experimental  results  (Malinovsky  and  Doroshev, 
1975).  Fields  for  graphite  and  diamond  are  calcu- 
lated from  these  experimental  data  and  the  esti- 
mated temperature  and  pressure  of  the  diamond- 
graphite  transition  in  the  Siberian  mantle. 


chromite  only  at  pressures  within  the  dia- 
mond stability  field  (Kesson  and  Ringwood, 
1989).  The  experimental  data  can  be  used 
to  estimate  the  compositions  of  garnets  in 
equilibrium  with  Cr-spinel  at  P-T  condi- 
tions corresponding  to  the  diamond-graph- 
ite transition  in  the  Siberian  lithosphere 
(Fig.  8).  The  compositions  of  garnets  in 
megacrystalline  rocks  that  contain  diamond 
and  chromite  are  consistent  with  the  field 
for  diamond  shown  in  Fig.  8  and  with  the 
knorringite  isopleths  based  on  experiment. 
Analyses  for  garnets  in  graphite  +  garnet  + 
chromite  assemblages  (not  plotted  in  Fig. 
8)  all  have  less  than  15%  knorringite  and 
are  thus  also  consistent  (Pokhilenko  et  al., 
1988).  Thus,  it  may  be  possible  to  develop 
a  useful  barometer  based  on  the  Ca  and  Cr 
contents  of  the  garnets  for  assemblages  that 
include  both  pyrope  and  chromite. 


References 

Kesson,  S.  E.,  and  A.  E.  Ringwood,  Slab-mantle 
interactions  2.  The  formation  of  diamonds, 
Chem.  Geol.,  78,  97-118,  1989. 

Malinovsky,  I.  Yu.,  and  A.  M.  Doroshev,  Stability 
of  garnet  of  pyrope-knorringite  row  in  the  field 
7=1000-1500°C  and  P=20-50  kb,  Nauka. 
Novosibirsk,  p.  23-31,  1975. 

O'Neill,  H.  St.  C,  and  B.  J.  Wood,  An  experimen- 
tal study  of  Fe-Mg  partitioning  between  garnet 
and  olivine  and  its  calibration  as  a 
geothermometer,  Contrib.  Mineral.  Petrol.  70, 
59-70,  1979. 

Pokhilenko,  N.  P.,  A.  S.  Rodionov,  T.  M.  Blinchik, 
and  E.  V.  Malygina,  Graphite-diamond  phase 
transition  and  its  significance  for  estimations  of 
P-T  conditions  of  equilibrium  of  ultrabasic 
xenoliths,  Ext.  Abstr.  vol.  Intern.  Symposium 
on  Composition  and  Processes  of  Deep-seated 
Zones  of  Continental  Lithosphere,  pp.  64-65, 
Novosibirsk,  1988. 

Pokhilenko,  N.  P.,  N.  V.  Sobolev,  and  Yu.  G. 
Lavrent'ev,  Xenoliths  of  diamondiferous  ultra- 
mafic  rocks  from  Yakutian  kimberlites,  2nd 
Int.  Kimb.  Conf.  Ext.  Abstr.  Vol.  Santa  Fe, 
1977. 


18 


CARNEGIE  INSTITUTION 


Richardson,  S.  H.,  J.  J.  Gurney,  A.  J.  Erlank,  and 
J.  W.  Harris,  Origin  of  diamonds  in  old  en- 
riched mantle,  Nature,  310,  198-202,  1984. 

Sobolev,  N.  V.,  Deep  Seated  Inclusions  in 
Kimberlites  and  Problem  of  Upper  Mantle 
Composition,  p.  1-264  Nauka,  Novosibirsk, 
1974  (English  Translation),  AGU,  1977. 

Sobolev,  N.  V.,  N.  P.  Pokhilenko,  and  E.  S. 
Yefimova,  Diamond-bearing  peridotite  xeno- 
liths  in  kimberlites  and  the  problem  of  the 
origin  of  diamonds,  Sov.  Geol.  Geophys.,  25, 
62-76,  1984. 

Viljoen,  K.  S.,  D.  H.  Robinson,  and  P.  M.  Swash, 
Diamond  and  graphite  peridotite  xenoliths  from 
the  Roberts  Victor  Mine,  5th  Intern.  Kimb. 
Conf.  Araxa,  Brazil,  Ext.  Abstr.  Vol.,  1991. 


Mantle  Metasomatism:  Evidence  from  a 

MARIE)  -  Harzburgite  Compound 

Xenolith 

F.  R.  Boyd 

Interpretation  of  metasomatized  rocks 
in  the  Earth's  crust  depends  critically  on  the 
study  of  outcrops.  We  are  not  fortunate, 
however,  in  having  exposures  of  the  deep 
portions  of  continental  cratons  which  have 
been  the  sites  of  metasomatic  events  over 
the  course  of  3-4  billion  years.  In  the 
absence  of  outcrops,  the  rare  discoveries  of 
xenoliths  exhibiting  contacts  between  rock 
types  have  particular  importance.  This  is 
especially  true  if  the  rocks  are  fragments  of 
an  igneous  intrusion  and  a  metasomatized 
conduit  wall. 

A  xenolith  containing  a  contact  between 
a  mica-rich  igneous  cumulate  and  a 
metasomatized  peridotite  was  discovered 
and  made  available  for  study  by  staff  of  the 
Anglo  Axnerican  Research  Laboratories. 
The  xenolith  was  recovered  from  the  com- 
bined coarse  concentrate  of  the  De  Beers 
mines  in  Kimberley. 

Mica-rich  xenoliths  from  the  Kimberley 


pipes  include  varieties  of  both  igneous  and 
metasomatic  origin.  Those  believed  to  be 
igneous  have  been  designated  by  the  acro- 
nym MARID,  based  on  the  names  of  the 
constituent  primary  minerals:  mica,  am- 
phibole,  rutile,  ilmenite,  and  diopside 
(Dawson  and  Smith,  1977).  Metasomatized 
peridotites  contain  a  variety  of  introduced 
phases  that  include  phlogopite,  potassic 
richterite  and  pargasite  amphiboles,  il- 
menite, rutile,  and  a  number  of  exotic  po- 
tassium and  barium  titanates  (Erlank  et  al., 
1987).  It  has  been  a  problem  to  understand 
the  nature  of  the  metasomatizing  fluids  and 
whether  or  not  they  are  related  to  kimberlite. 
Is  kimberlite  the  magma  from  which 
MARID  rocks  are  derived  and  are  MARID 
rocks  the  remnants  of  metasomatic  sources? 
Only  one  xenolith  containing  a  contact 
between  a  MARID  rock  and  a  peridotite 
has  previously  been  described  (Waters  et 
al.,  1989).  This  xenolith  was  interpreted  as 
a  fragment  of  an  intrusive  with  attached 
wall  rock  and  study  of  it  has  helped  to 
establish  the  hypothesis  that  MARID  rocks 
are  remnants  of  sources  of  metasomatic 
fluids.  Extensive  post-metasomatic  alter- 
ation of  the  peridotite  to  carbonate  in  this 
xenolith,  however,  has  obscured  the  pri- 
mary mineralogy  and  the  metasomatic  im- 
print. 


Petrography 

The  sawed  fragment  of  the  compound 
xenolith  analyzed  in  the  present  study  is  6 
cm  in  maximum  dimension  and  2  cm  thick; 
the  fragment  is  numbered  FRB  1455.  The 
MARID  portion  of  the  xenolith  consists  of 


GEOPHYSICAL  LABORATORY 


19 


over  95  %  phlogopite,  predominantly  in  the 
form  of  tablets,  1-2  mm  in  length,  aligned 
parallel  to  the  contact  with  the  peridotite. 
The  coarse  mica  is  strained,  having  undu- 
late extinction  and  bent  cleavage.  Finer- 
grained  mica  neoblasts  are  interspersed 
between  the  coarse  tablets.  The  coarse 
mica  is  pale,  but  the  neoblasts  and  mantles 
on  the  mica  tablets  have  a  darker,  yellow- 
ish-brown color. 

Granules  of  diopside  a  few  tenths  of  a 
mm  in  diameter  are  dispersed  between  the 
mica  tablets  and  clustered  with  serpentine 
in  lenticles.  Elongate  grains  of  rutile  rang- 
ing up  to  a  mm  are  less  abundant.  Small 
patches  of  calcite  that  may  be  primary  are 
widely  dispersed  and  enclose  euhedral  crys- 
tals of  sphene,  0.1-0.2  mm.  Sphene  is  an 
unusual  mineral  in  mantle  rocks,  but 
Dawson  and  Smith  (1977)  have  noted  its 
occurrence  as  an  accessory  phase  in  a 
MARID  xenolith  from  the  Wesselton  mine. 

The  harzburgite  consists  primarily  of 
coarse  olivine,  ranging  up  to  1  cm,  with  a 
minor  proportion  of  smaller  grains  of 
enstatite.  Olivine  neoblasts  locally  form 
interstitial  zones  but  are  insufficiently  abun- 
dant to  envelope  the  primary  grains.  Coarse 
crystals  of  phlogopite,  mantled  and  seamed 
with  finer-grained  mica,  range  up  to  8  mm 
and  have  a  color  and  degree  of  strain  simi- 
lar to  the  mica  in  the  adjacent  MARID  rock. 
The  section  analyzed  contains  several  1- 
mm  grains  of  garnet  enveloped  in  fine- 
grained mica,  forming  a  replacement  tex- 
ture like  that  cited  by  Erlank  et  al.  (1987) 
and  observed  in  Jagersfontein  peridotites 
(Boyd  and  Mertzman,  1987).  The 
harzburgite  also  contains  irregularly- 
shaped,  coarse  blebs  of  ilmenite  mantled 


by  rutile  which  are  up  to  a  centimeter  in 
maximum  dimension.  The  contact  between 
the  peridotite  and  MARID  is  sharp;  minor 
serpentine  and  ilmenite  lenticles  are  local- 
ized in  the  peridotite  adjacent  to  the  con- 
tact. 


Mineral  Chemistry 

The  mica  and  accessory  phases  in  the 
MARID  portion  of  xenolith  1455  have 
compositions  that  are  comparable  to  those 
previously  reported  for  MARID  rocks.  The 
phlogopite  and  diopside  have  low  Cr203 
(Table  4);  the  diopside  is  also  relatively  low 
in  AI2O3.  The  mg  number  of  the  coarse 
phlogopite  is  0.87,  near  the  high  end  of  the 
range  reported  by  Dawson  and  Smith 
(1977).  The  diopside  with  an  mg  number 
of  0.86  is  markedly  more  Fe-rich  than  the 
Cr-diopsides  in  common  garnet  lherzolites. 
The  darker  mica  that  forms  neoblasts  and 
mantles  on  the  coarse  tablets  is  enriched  in 
Ti  but  not  in  total  Fe. 

The  olivine  and  enstatite  in  the  associ- 
ated peridotite  have  extremely  variable  Mg/ 
(Mg+Fe).  Cores  of  large  olivine  crystals 
are  Mg-rich  (0.94)  but  neoblasts  and  mar- 
gins on  the  coarse  grains  are  as  Fe-rich  as 
0.88  (Tables  4,  5).  These  variations  are 
irregular,  differing  widely  from  grain  to 
grain;  they  are  not  systematic  in  reference 
to  the  MARID-peridotite  contact.  Enstatite 
grains  have  comparable  enrichment  in  Fe 
and  the  secondary  enstatite  has  higher  Ca 
and  Ti.  Enstatite  close  to  the  MARID 
contact,  however,  is  markedly  depleted  in 
Al  (Table  5).  The  pale  mica  and  darker 
mica  in  the  harzburgite  have  compositions 


20 


CARNEGIE  INSTITUTION 


Table  4.  Compositions  of  primary  minerals  in  MARID 
1455,  Kimberley,  RSA. 


harzburgite  compound  xenolith,  FRB 


Harzburgite 

MARK) 

Olivine 

Enstatite 

Garnet 

Mica 

Diopside 

Rutile 

Sphene 

Si02 

41.5 

57.3 

41.4 

41.9 

54.3 

<0.03 

30.4 

Ti02 

<0.03 

0.10 

0.70 

1.77 

0.40 

97.2 

41.9 

AI2O3 

<0.03 

0.68 

20.5 

10.1 

0.50 

0.94 

0.55 

Cr203 

0.04 

0.48 

3.26 

0.18 

0.48 

1.89 

0.03 

FeO 

5.74 

4.63 

8.20 

6.22 

5.11 

0.09 

0.79 

MnO 

0.07 

0.11 

0.39 

0.04 

0.13 

<0.03 

0.04 

MgO 

53.1 

36.4 

20.6 

23.5 

17.6 

0.09 

0.10 

CaO 

<0.03 

0.50 

5.51 

<0.03 

20.6 

0.09 

24.8 

Na20 

n.d. 

0.21 

0.11 

0.10 

0.94 

n.d. 

2.50 

K20 

n.d. 

n.d. 

n.d. 

10.4 

<0.03 

n.d. 

0.07 

NiO 

0.36 

0.08 

<0.03 

0.15 

0.03 

0.03 

0.08 

total 

100.8 

100.5 

100.7 

94.4 

100.09 

100.3 

101.3 

Mg/(Mg+Fe) 

0.943 

0.934 

0.818 

0.871 

0.861 

- 

- 

Table  5.  Compositions  of  secondary  minerals  occurring  as  discrete  grains  or  mantles  on 
primary  grains  in  MARID-harzburgite  compound  xenolith,  FRB  1455,  Kimberley,  RSA. 


Harzburgite 


MARID 


Olivine 

Enstatite 

Ilmenite 

Rutile 

Mica 

Mica 

Si02 

40.5 

56.4 

<0.03 

<0.03 

40.6 

40.0 

Ti02 

0.03 

0.30 

53.6 

94.0 

3.73 

3.64 

A1203 

<0.03 

0.22 

0.81 

0.93 

11.1 

11.1 

Cr203 

0.04 

0.10 

2.59 

2.97 

0.19 

0.23 

FeO 

11.5 

8.16 

30.4 

0.29 

5.29 

5.91 

MnO 

0.13 

0.13 

0.27 

<0.03 

0.08 

0.07 

MgO 

48.9 

33.9 

13.6 

0.10 

23.2 

22.7 

CaO 

0.06 

0.91 

0.04 

0.04 

<0.03 

<0.03 

Na20 

n.d. 

0.14 

n.d. 

n.d. 

0.20 

0.17 

K20 

n.d. 

n.d. 

n.d. 

n.d. 

10.2 

10.3 

NiO 

0.38 

0.07 

0.19 

<0.03 

0.12 

0.12 

total 

101.5 

100.3 

101.5 

98.3 

94.7 

94.2 

Mg/(Mg+Fe) 

0.884 

0.881 

0.443 

- 

0.887 

0.873 

very  similar  to  counterparts  in  the  MARID 
portion  of  the  xenolith.  Rutile  in  both  the 
MARID  and  peridotite  contains  several 
percent  of  Cr203,  a  distinguishing  feature 
noted  for  MARID  rutiles  by  Dawson  and 
Smith  (1977). 


Thermobarometry 

The  occurrence  of  olivine  and  garnet 
together  with  enstatite  makes  it  possible  to 
estimate  the  temperature  and  depth  of  equili- 
bration of  the  harzburgite- MARID  assem- 


GEOPHYSICAL  LABORATORY 


21 


Table  6.  Estimates  of  the  temperature  and  depth  of  equilibration  of  the  harzburgite- 
MARID  xenolith,  FRB   1455. 


Thermometer 

Barometer         Temperature, 

°C 

Depth,  km 

olv-gar-opx 

O'Neill-Wood  (1979) 
O'Neill-Wood  (1979) 

MacGregor(1974) 
Nickel-Green  (1985) 

opx-gar-opx 

530 
560 

50 
70 

Harley  (1984) 

MacGregor(1974) 

850 

115 

Discussion 

blage.  Such  estimates  have  not  previously 
been  possible  for  MARK)  xenoliths  be- 
cause they  do  not  in  themselves  contain  a 
necessary  combination  of  phases.  Enrich- 
ment of  some  of  the  olivine  and  enstatite  in 
Fe  creates  an  obvious  uncertainty,  but  the 
most  magnesian  compositions  are  taken  to 
be  primary  (Table  4).  This  assumption  may 
not  be  correct  for  the  enstatite,  however, 
because  it  has  a  slightly  lower  mg  number 
than  the  olivine,  the  reverse  of  the  usual 
relationship.  Combining  the  olivine-gar- 
net  and  the  orthopyroxene-garnet  thermom- 
eters with  two  Al-enstatite  barometers  pro- 
vides results  that  suggest  a  shallow  mantle 
origin  for  this  xenolith  (Table  6),  perhaps 
close  to  the  upper  limit  of  garnet  peridotite 
stability.  There  is  a  discrepancy  of  about 
300°C,  however,  between  temperatures  cal- 
culated with  the  olivine-garnet  and  enstatite- 
gamet  thermometers,  and  because  of  the 
temperature  dependance  of  the  barometer 
there  is  a  corresponding  discrepancy  in 
depth.  This  is  probably  due  to  slight  en- 
richment of  the  primary  enstatite  in  Fe,  and 
the  olivine-gamet  estimates  are  believed  to 
be  more  reliable. 


The  pronounced  gradients  in  Fe/Mg  in 
the  olivine  of  xenolith  1455  may  be  evi- 
dence that  the  metasomatism  occurred  in 
association  with  the  eruptions  that  pro- 
duced the  Kimberley  diatremes.  Such  gra- 
dients are  unlikely  to  persist  in  the  upper 
mantle  for  periods  of  many  million  years. 
The  eruption  of  the  six  large  diatremes  at 
Kimberley,  each  with  multiple  units  of 
kimberl'te,  was  a  complex  volcanic  event, 
however,  and  the  origin  of  the  xenolith  may 
have  involved  more  than  one  body  of 
magma. 

A  second  factor  of  importance  is  the 
relatively  low  ambient  temperature  and 
depth  that  are  reflected  by  the  composi- 
tions of  the  primary  minerals  in  the 
harzburgite  (Table  6).  It  appears  that  the 
metasomatic  process  occurred  at  shallow 
levels  in  the  craton,  far  removed  from  the 
top  of  the  asthenosphere  and  from  the  zone 
of  kimberlite  generation.  In  such  circum- 
stances  it  is  probable  that  the 
metasomatizing  agent  was  either  the 
MARID  magma  or  the  kimberlite  itself, 
and  possibly  they  are  the  same.  A  relation- 


22 


CARNEGIE  INSTITUTION 


ship  between  MARID  rocks  and  kimberlite 
has  previously  been  proposed  (Dawson 
and  Smith,  1977).  The  only  alternative 
source  for  the  MARID  rock  that  is  known 
to  exist  are  intrusions  of  mica-rich  (Group 
II)  kimberlite  at  nearby  Loxtondal  that  have 
an  age  40  My  older  than  the  Kimberley 
diatremes  (E.M.W.  Skinner,  personal  com- 
munication). 

Many  of  the  peridotite  and  dunite  xeno- 
liths  from  Kimberley  have  been  subject  to 
Fe-Ti  metasomatism  with  introduction  of 
ilmenite  and  development  of  Fe-enriched 
marginal  zones  on  primary  olivine  and  py- 
roxene. These  effects  are  believed  to  have 
been  caused  by  kimberlite  magmatism 
(Boyd  etal.,  1983),  and  they  have  appeared 
to  be  distinct  from  the  introduction  of  coarse, 
Mg-rich  mica  and  less  commonly  pargasite 
that  may  have  originated  in  events  that  long 
pre- dated  kimberlite  eruption.  Crystalliza- 
tion of  metasomatic  K-richterite  in  perido- 
tite, however,  has  involved  introduction  of 
mica  together  with  titaniferous  phases 
(Erlankeftf/.,  1987).  The  latter metasomites 
appear  to  be  linked  to  MARID  intrusions 
by  similarities  in  mineralogy  and  by  the 
evidence  from  xenolith  1455  and  from  the 
similar  compound  xenolith  studied  by 
Waters  etal.  (1989).  Some  of  the  variations 
in  metasomatic  imprint  may  be  conse- 
quences of  the  varying  state  of  kimberlite- 
related  metasomatizing  fluid  and  varying 
depth  at  which  the  reactions  occurred. 

Compositional  changes  induced  in  the 
peridotite  of  xenolith  1455  are  extremely 
irregular,  and  their  magnitude  does  not 
vary  systematically  with  distance  from  the 
MARID  contact  on  a  scale  of  several  cen- 


timeters. Moreover,  the  irregular  spatial 
variation  does  not  seem  to  reflect  the  origi- 
nal presence  of  some  other  contact.  Most 
likely,  the  agent  of  metasomatism  was  a 
hydrothermal  fluid  derived  from  nearby 
magma  that  penetrated  the  harzburgite  ir- 
regularly along  fractures  and  grain  bound- 
aries. 

It  is  suggested  that  the  MARID  rock  of 
xenolith  1455  is  an  aggregate  of  mica  with 
minor  diopside  and  rutile  that  was  plated  on 
the  peridotite  wall  of  a  dike  or  other  conduit 
through  which  kimberlite  (?)  magma  was 
erupting.  The  magma  from  which  MARID 
rocks  have  crystallized  has  been  proposed 
to  be  lamproite  (Waters,  1987).  There  are 
no  lamproite s  in  the  Kimberley  area,  how- 
ever, and  kimberlite  itself  is  an  alternative 
possibility.  The  metasomatism  could  have 
been  produced  by  the  same  volume  of  melt, 
but  an  earlier  or  later  origin  during  the 
restricted  period  of  kimberlite  magmatism 
seems  possible.  The  more  darkly  colored 
Ti-rich  mantles  on  the  mica  in  both  the 
MARID  and  peridotite  portions  of  the  xe- 
nolith may  be  evidence  of  a  change  in 
composition  of  the  metasomatic  fluid  or 
even  a  two-stage  process. 


References 

Boyd,  F.  R.,  and  S.  A.  Mertzman,  Composition 
and  structure  of  the  Kaapvaal  lithosphere, 
southern  Africa,  in  Magmatic  Processes: 
Physicochemical  Principles,  B.  O.  My  sen, 
ed.,  Geochemical  Society  Special  Publication, 
No.  l,p.  13-24,1987. 

Boyd,  F.  R.,  R.  A.  Jones,  and  P.  H.  Nixon,  Mantle 
metasomatism:  the  Kimberley  dunites, 
Carnegie  Instn.  Washington  Year  Book,  82, 
330-336,  1983. 


GEOPHYSICAL  LABORATORY 


23 


Dawson,  J.  B.,  and  J.  V.  Smith,  The  MARID 
(mica-amphibole-rutile-ilmenite-diopside) 
suite  of  xenoliths  in  kimberlite,  Geochim. 
Cosmochim.  Acta,  41,  309-323,  1977. 

Erlank,  A.  J.,  F.  G.  Waters,  C.  J.  Hawkes  worth,  S. 
E.  Haggerty,  H.  L.  Allsopp,  R.  S.  Rickard,  and 
M.  Menzies,  Evidence  for  mantle  metasoma- 
tism in  peridotite  nodules  from  the  Kimberley 
Pipes,  South  Africa,  in  Mantle  Metasomatism, 
M.  Menzies  and  C.  J.  Hawkesworth,  eds., 
Academic  Press,  1987. 

Harley,  S.  L.,  An  experimental  study  of  the  parti- 
tioning of  Fe  and  Mg  between  garnet  and 
orthopyroxene,  Contrib.  Mineral.  Petrol,  86, 
359-373,  1984. 

MacGregor,  J.  D.,  The  system  MgO-Al203-SiC>2: 
Solubility  of  AI2O3  in  enstatite  for  spinel  and 
garnet  peridotite  compositions,  Amer.  Min- 
eral., 59,  110-119,  1974. 

Nickel,  K.  G.,  and  D.  H.  Green,  Empirical 
geothermobarometry  for  garnet  peridotites  and 
implications  for  the  nature  of  the  lithosphere, 
kimberlite s  and  diamonds,  Earth  Planet.  Sci. 
Lett.,  73,  158-170,  1985. 

O'Neill,  H.  St.  C.,  andB.  J.  Wood,  An  experimen- 
tal study  of  Fe-Mg  partitioning  between  garnet 
and  olivine  and  its  calibration  as  a 
geothermometer,  Contrib.  Mineral.  Petrol., 
70,  59-70,  1979. 

Waters,  F.  G.,  A  suggested  origin  of  MARID 
xenoliths  in  kimberlites  by  high  pressure  crys- 
tallization of  an  ultrapotassic  rock  such  as 
lamproite,  Contrib.  Mineral.  Petrol.,  95, 523- 
533, 1987. 

Waters,  F.  G.,  A.  J.  Erlank,  and  L.  R.  M.  Daniels, 
Contact  relationships  between  MARID  rock 
and  metasomatized  peridotite  in  a  kimberlite 
xenolith,  G eo chemical  J .,  23,  11-17,  1989. 


Boron  and  Beryllium  Concentrations  in 

Subduction-Related  Metamorphic  Rocks 

of  the  catalina  schist!  implications  for 

subduction-zone  recycling 

Gray  E.  Bebout,  Jeffrey  G.  Ryan*  and 
William  P.  Leeman** 


Enrichments  in  B  and  10Be  in  arc  volca- 
nic rocks  have  been  interpreted  to  reflect 
slab  additions  to  arc  source  regions  via 
hydrous  fluids  or  silicate  liquids  (Morris  et 
ai,  1990;  Ryan  and  Langmuir,  1991).  The 
two-component  mixing  relations  among  B, 
10Be,  and  9Be  demonstrated  by  Morris  et  al. 
(1990)  are  believed  to  result  from  the  addi- 
tion of  a  homogenized  slab-derived  compo- 
nent, probably  a  hydrous  fluid  which 
strongly  fractionates  B  from  Be.  The  con- 
centrations of  B  and  Be  in  sediments  and 
crustal  materials  believed  to  be  subducted 
are  reasonably  well  characterized.  Exten- 
sive B  and  Be  data  sets  exist  for  arc  volcanic 
rocks  (Teraef  a/.,  1986;  Morris  etal,  1990; 
Ryan  and  Langmuir,  1988, 1 991;  Leeman  et 
al.,  1990);  thus,  the  volcanic  output  of  these 
elements  can  be  estimated  (see  Ryan  and 
Langmuir,  1991,  for  discussion  of  B).  How- 
ever, relatively  little  is  known  about  the 
effects  of  slab  metamorphism  on  the  concen- 
trations of  3,  Be,  and  other  trace  elements  in 
subducted  rocks.  These  metamorphic  pro- 
cesses may  dictate  the  efficiency  with  which 

1 

these  elements  are  transferred  from  surface 
reservoirs  to  arc  source  regions;  Moran  et 

*  Department  of  Terrestrial  Magnetism,  Carnegie 
Institution  of  Washington 
**  Keith -Wiess  Geological  Laboratories,  Rice 
University,  Houston,  Texas  7725 1 


24 


CARNEGIE  INSTITUTION 


al.  (1991)  discuss  the  consequences  of  B 
loss  due  to  slab  metamorphism  for  the 
mass -balance  of  B  inputs  to  arc  source 
regions.  Also  of  particular  interest  are  the 
relative  capabilities  of  metamorphic 
de volatilization  and  migmatization  pro- 
cesses to  fractionate  B  and  Be  sufficiently 
to  produce  the  wide  ranges  of  B/Be  ob- 
served in  arc  volcanic  rocks  (5  to  200; 
Morris  etal,  1990). 

In  this  report,  B  and  Be  concentration 
data  are  presented  for  metamorphosed  sedi- 
mentary and  mafic  rocks  and  veins  and 
pegmatites  from  the  Catalina  Schist,  ex- 
posed on  Santa  Catalina  Island  in  southern 
California.  Mineral  reservoirs  for  these 
elements  during  devolatilization  of 
metasedimentary  and  metamafic  rocks  are 
considered,  and  the  extent  to  which  B  and 
Be  were  mobilized  during  progressive  vola- 
tile loss  and  migmatization  is  discussed. 


Finally,  the  implications  of  these  data  for 
trace  element  fractionation  among  miner- 
als, H20-rich  fluids,  and  silicate  liquids 
deep  in  subduction  zones  are  discussed,  as 
well  as  the  relevance  of  these  observations 
for  interpretation  of  B-Be  concentrations 
in  arc  volcanic  rocks. 

The  Catalina  Schist  consists  of  three 
major  metamorphic-tectonic  units  juxta- 
posed along  low-angle  faults  (Piatt,  1975; 
Sorensen  and  Barton,  1987;  Bebout  and 
Barton,  1989).  These  units  contain  similar 
lithologies  (metamorphosed  sandstones, 
shales,  and  cherts;  metabasaltic  and 
metagabbroic  rocks)  and  range  in  grade 
from  lawsonite-albite  to  amphibolite.  The 
Catalina  rocks  are  well  suited  for  examina- 
tion of  the  effects  of  metamorphism  on 
trace  element  and  stable  isotope  composi- 
tion over  a  wide  range  of  metamorphic 
conditions  (350°-750°C,  5-11  kbar).  Evi- 


ocmm     m 


Lawsonite-Albite 
Blueschist 


•  o 


Greenschist/Epidote 
Amphibolite 


(KCD   Amphibolite 


Seafloor  Sediments 
50  -  200  ppm  B 


0 


40 


80  120 

Boron  (ppm) 


160 


200 


Fig.  9.  Boron  content  in  whole-rock  metasedimentary  samples  from  the  Catalina  Schist.  The  large 
range  in  B  concentration  of  the  lawsonite-albite  and  blueschist  grade  rocks  probably  reflects 
variability  in  sedimentary  protoliths  (see  B  data  for  seafloor  sediments  in  Moran  et  al.,  1991). 
With  increasing  metamorphism,  B  content  is  decreased  and  becomes  more  uniform. 


GEOPHYSICAL  LABORATORY 


25 


dence  for  fluid  transport  and  associated 
mass  transfer  during  metamorphism  includes 
the  occurrence  of  veins,  reaction  zones  be- 
tween disparate  lithologies,  changes  in  bulk 
chemical  composition,  and  changes  in  iso- 
topic  composition  (Bebout  and  Barton, 
1989).  Stable  isotope  and  petrologic  stud- 
ies of  the  Catalina  Schist  have  yielded  evi- 
dence for  progressive  devolatilization  and 
km-scale  transport  of  H20-rich  C-O-H-S- 
N  fluid  during  metamorphism  (Bebout  and 
Barton,  1989;  Bebout,  1991).  Intheamphi- 
bolite  unit,  pegmatites  represent high-P mass 
transfer  via  silicate  liquids  derived  through 
vapor-saturated  partial  melting  of  sedimen- 
tary and  mafic  rocks  (Sorensen  and  Barton, 
1987;  Bebout  and  Barton,  1989). 


Boron  and  Beryllium  Concentration  Data 

Boron  and  Be  concentration  data  for  96 
metasedimentary,  metamafic,  and  meta-ul- 
tramafic  rocks,  mineral  separates,  veins  and 
pegmatites  were  obtained  by  inductively 
coupled  plasma  (ICP)  analytical  techniques; 
samples  were  fused  with  Na2C03  flux.  All 
chemistry  was  done  in  the  laboratory  of  J. 
D.  Morris  and  F.  Tera  at  the  Department  of 
Terrestrial  Magnetism. 

Boron  concentrations  and  B/Be  of 
metasedimentary  rocks  decrease  progres- 
sively with  increasing  metamorphic  grade 
(Figs.  9,  10);  the  decrease  in  B  content  is 
consistent  with  results  for  other  metamor- 
phic suites  (e.g.,  Shaw  etal.,  1988;  Nabelek 
et  al.,  1990;  Moran  et  ai,  1991).  Lowest- 
grade  lawsonite-albite  rocks  (inferred  meta- 
morphic conditions  of  350°-450°C,  5-8  kbar) 
range  in  B  content  from  12  to  181  ppm  B 


with  a  mean  of  73  ppm,  whereas  high-grade 
amphibolite  equivalents  (inferred  metamor- 
phic conditions  of  650°-750°C,  8-11  kbar) 
range  from  5.4  to  19  ppm  B  with  a  mean  of 
12.2  ppm.  Beryllium  contents  in  metasedi- 
mentary rocks  of  all  grades  range  from  0.3 
to  1 .2  ppm.  Thus,  the  decrease  in  B/Be  from 
a  mean  of -72  for  the  lawsonite-albite  grade 
metasedimentary  rocks  to  a  mean  of -27  for 
amphibolite  grade  equivalents  is  attributable 
to  loss  of  B  (Fig.  10).  Metamafic  rocks 
contain  from  3  to  20  ppm  B  (12  samples) 
and  from  0.21  to  1.31  ppm  Be,  with  the 
lower  values  for  both  elements  occurring  in 
the  highest  grade  rocks.  Metamafic  B/Be 
shows  no  significant  variation  with  increas- 
ing metamorphic  grade  and  ranges  from  3  to 
20  (Fig.  10). 

As  measures  of  B  and  Be  mobility  in 
hydrous  fluids  and  felsic  silicate  melts,  B 
and  Be  concentrations  were  obtained  for 
veins,  which  precipitated  from  the  hydrous 
fluids,  and  for  pegmatites,  which  reflect 
migmatization  in  the  amphibolite  unit.  In 
blueschist  metasedimentary  exposures, 


1000F 


100 


CD 

CD 

m 


Lawsonite-Albite 
Blueschist    —-*■""         *»  ^* 


Greenschist 
&  Epidote 
Atiphibolite 


10  100 

B  (ppm) 


1000 


Fig.  10.  B/Be  vs.  B  content  of  metasedimentary 
rocks  and  metamafic  rocks  (shaded  field).  B  and 
B/Be  of  metasedimentary  rocks  decrease  with 
increasing  grade.  With  the  exception  of  one 
glaucophanic  greenschist  sample  (60  ppm  B,  1.98 
ppm  Be,  B/Be  of  30),  all  of  the  measured 
metamafic  rocks  have  B/Be  less  than  20. 


26 


CARNEGIE  INSTITUTION 


i 1 — 

n  =  10;  32.4 
±10.3 

Na-amphibole 

(n  =  10;  33.4  ± 

11.5) 


Blueschist  Metasediments 


Blueschist  Veins 


OO  Host-Rocks 


O  Pegmatites 


O  Metasedimentary 
Metamafic 


0 


20 


40 


60 


B/Be 


80 


100 


Fig.  11.  Comparisons  of  B/Be  of  veins  and  pegmatites  with  those  of  host-rocks  from  the 
blueschistand  amphibolite  units. 


nearly  monomineralic  sodic  amphibole 
veins  have  B  and  Be  contents  and  B/Be 
ranges  similar  to  their  host  rocks  (Fig.  11). 
One  albite  +  graphite  vein  contains  4.6  ppm 
B  and  0.13  ppm  Be  and  also  has  B/Be 
(35.4)  similar  to  that  of  the  surrounding 
blueschist  grade  metasedimentary  rocks 
(32.4  ±10).  Pegmatites  in  the  amphibolite 
unit  show  wide  ranges  in  B  content  (6  to  26 
ppm)  and  Be  content  (0.6  to  4. 1  ppm)  but 
have  B/Be  similar  to  or  slightly  lower  than 
their  mafic  and  sedimentary  hosts  (Fig.  11). 
Consideration  of  B  and  Be  behavior 
during  devolatilization  and  migmatization 
requires  knowledge  of  B  and  Be  mineral 
residency.  Table  7  contains  data  for  whole- 
rock  and  mineral  separate  samples  from 
several  high-grade  metasedimentary  rocks 
and  pegmatites.  These  data  demonstrate 
that  B  is  concentrated  in  white  micas, 
whereas  Be  appears  to  be  somewhat  more 
evenly  distributed.  For  all  but  sample  7-3- 
23,  which  contains  tourmaline,  muscovite 


contains  more  B  than  the  whole-rock 
sample;  Be  content  of  the  muscovite  is 
comparable  to  that  of  the  whole-rock 
samples.  Preferential  enrichment  of  B  in 
micas  is  consistent  with  the  ion  microprobe 
results  of  Domanik  et  al.  ( 1 99 1 )  for  samples 
of  the  Catalina  Schist.  With  the  exception 
of  several  occurrences  in  felsic  pegmatites 
(amphibolite  unit)  and  in  greenschist-grade 
metacherts.  tourmaline  is  not  a  significant 
host  for  B  in  the  Catalina  samples. 


Discussion 

The  data  presented  in  this  report  are 
consistent  with  removal  of  B  and  Be  from 
metamorphosed  sedimentary  and  mafic 
rocks  by  both  H20-rich  fluids  and  felsic 
silicate  liquids.  The  B-Be  signatures  of 
these  two  "fluids"  were  apparently  dra- 
matically different.   The  H20-rich  fluids 


GEOPHYSICAL  LABORATORY 

Table  7.  Boron  and  beryllium  concentrations  of  mineral  separates  (in  ppm). 


27 


Whole-Rock 


Muscovite 


Feldspar 


Sample 


B 


Be 


B 


Be 


*  not  determined 

**  tourmaline-bearing  pegmatite  from  metasedimentary  exposure 


B 


Be 


Metasedimentary  Rocks 

6-3-41' 

(epidote  amphibolite) 

32.0 

4.1                   n.d.*            n.d. 

4.2 

1.1 

8-1-3 
(amphibolite) 

19.0 

0.41                 39.1            0.69 
Pegmatites  -  Amphibolite  Unit 

n.d. 

n.d. 

6-3-24 
6-3-75 
7-3-23** 

37.0 
25.7 
970 

1.3                  48.0           0.92 
4.1                  60.6            4.0 
2.1                   113             3.0 

n.d. 
11.1 
n.d. 

n.d. 
5.8 
n.d. 

liberated  through  progressive 
devolatilization  are  inferred  to  have  had 
high  B/Be;  their  removal  resulted  in  de- 
crease of  the  B/Be  of  the  residual  rocks.  To 
explain  the  similarity  of  the  Na-amphibole 
vein  and  blueschist  metasedimentary  rock 
B  and  Be  concentrations  and  B/Be,  a  model 
is  suggested  where  veins  precipitated  from 
fluids  which  were  previously  equilibrated 
with  respect  to  B  and  Be  with  host  rocks  (or 
with  similar  rocks  upstream  of  current  hosts). 
Stable-isotope  data  are  consistent  with  this 
model;  O,  H,  and  C  isotope  compositions  of 
vein  minerals  are  in  many  cases  similar  to 
those  of  the  same  minerals  in  host  rocks 
(Bebout  and  Barton,  1989).  If  this  model  is 
correct,  the  veins  would  have  had  mineral/ 
fluid  partition  coefficients  for  B  and  Be 
similar  to  the  bulk-rock/fluid  partition  coef- 
ficients of  the  host  rocks.  If  the  pegmatites 
are  directly  representative  of  silicate  liquid 
compositions  produced  during  melting,  the 
silicate  liquids  had  B/Be  similar  to  those  of 


host  rocks.  The  relatively  low  B/Be  of  the 
pegmatites  may  thus  reflect  earlier  removal 
of  B  from  the  source  rocks  during 
devolatilization  (Figs.  10, 1 1 ;  cf.  Leeman  et 
al,  1991). 

Release  of  B  and  other  fluid-mobile  trace 
elements  during  devolatilization  may  be 
imagined  as  a  result  of  discontinuous  reac- 
tions involving  breakdown  of  mineral  hosts 
(e.g.,  see  Nabelek  et  at.,  1990),  or  by  a 
process  of  continual  partitioning  from  min- 
eral hosts  into  fluids,  or  by  a  combination  of 
these  processes.  ThegoodfitoftheCatalina 
N  concentration  and  isotope  data  with  a 
Rayleigh  distillation  model  (Bebout  and 
Fogel,  1 991 ;  Bebout,  this  Report)  may  have 
implications  for  the  mechanisms  of  B  loss 
during  progressive  devolatilization.  Boron 
and  N  show  correlated  decreases  in  concen- 
tration with  progressive  volatile  loss  in  the 
Catalina  metasedimentary  rocks  (see 
Bebout,  this  Report),  and  both  appear  to  be 
concentrated  in  white  micas.  During  pro- 


28 


CARNEGIE  INSTITUTION 


gressive  de volatilization  of  the  Catalina 
Schist,  N  isotopic  composition  evolved 
through  incremental  loss  of  N2  equilibrated 
with  the  rock  N  reservoir,  which  was  largely 
dominated  by  the  micas.  Decreases  in  B 
concentration  may  also  reflect  progressive 
boron  partitioning  from  B-rich  minerals 
(primarily  the  micas)  into  H^O-rich  fluids 
derived  largely  by  chlorite-breakdown  re- 
actions (see  Bebout,  1991).  Thus,  the  B 
loss  may  not  simply  reflect  the  breakdown 
of  micas,  which  show  no  variation  in  com- 
bined modal  abundance  with  grade. 

These  results  virtually  eliminate  a  bulk 
sediment/slab  mixing  process  (presumably 
mixing  of  a  melt  derived  from  heteroge- 
neous slab  sources)  as  an  explanation  of  B- 
Be  systematics  in  high-B/Be  volcanic  arcs 
and  strengthen  the  arguments  of  Morris  et 
al.  ( 1 990)  for  addition  to  arc  magma  sources 
of  a  fractionated  (high-B/Be),  slab-derived, 
hydrous-fluid  component.  The  data  pre- 
sented here  predict  that,  before  subducted 
mafic  and  sedimentary  rocks  reach  depths 
of  the  Wadati-Benioff  zone  below  arcs  (80 
to  150  km),  the  B/Be  of  these  rocks  is  likely 
to  be  decreased  to  <40.  As  is  also  suggested 
by  Moran  et  al.  (1991),  the  simple  mixture 
of  sediment  and  slab  with  B/Be  reduced  by 
devolatilization  will  not  produce  a  mixing 
component  with  sufficiently  high  B/Be  to 
explain  the  linear  trends  in  the  arc  data  (see 
Fig.  2  in  Morris  et  al,  1990). 

Because  fractionation  of  B  and  Be  in  the 
mantle  wedge  during  melting  processes  is 
unlikely  (Morris  et  al,  1990;  Ryan  and 
Langmuir,  1991),  the  varying  impact  of 
hydrous  fluid  or  melt  additions  to  arc  source 
regions  may  explain  some  of  the  B/Be 
variability  observed  in  arc  volcanic  rocks 


(see  B-Be  data  for  arcs  in  Morris  et  al., 
1990;  Leeman  et  al,  1990;  Ryan  and 
Langmuir,  1991).  Subduction  zones  with 
cooler  inferred  thermal  structures  on  aver- 
age show  higher  B/Be  (e.g.,  Aleutian  and 
New  Britain  arcs  with  B/Be  of  5-40  and  40- 
200,  respectively),  consistent  with  addi- 
tion of  high  B/Be  hydrous  fluids.  Subduc- 
tion zones  with  hotter  inferred  thermal  struc- 
tures produce  volcanic  rocks  with  rela- 
tively low  B/Be  (e.g.,  the  Cascades  and 
Woodlark  Basin  with  B/Be  of  2-10  and  4- 
15,  respectively),  perhaps  consistent  with 
addition  of  a  melted  component  from  a  slab 
previously  stripped  of  B  by  lower-tempera- 
ture devolatilization.  Morris  et  al.  (1990) 
and  Ryan  and  Langmuir  (1991)  have  docu- 
mented decreases  in  B/Be  across  individual 
arcs,  in  all  cases  varying  from  high  front- 
arc  values  to  low  B/Be  indistinguishable 
from  MORB  in  back-arc  regions.  These 
decreases  could  presumably  represent  the 
diminishing  effects  of  hydrous  fluid  addi- 
tion (contributing  high-B/Be  signatures) 
and/or  the  onset  of  melt  dominated  B-Be 
transfer  (contributing  low-B/Be  signatures). 


Conclusions 

These  results  demonstrate  that  B  and  Be 
are  significantly  fractionated  during  meta- 
morphic  devolatilization;  this  process  pro- 
duces high-B/Be  hydrous  fluids  and  results 
in  the  dramatic  reduction  of  the  B/Be  of  the 
subducted  rocks.  In  contrast,  partial  melt- 
ing, which  does  not  strongly  fractionate  B 
and  Be,  may  produce  silicate  melts  with 
low  B/Be  inherited  from  previously 


GEOPHYSICAL  LABORATORY 


29 


devolatilized  source  rocks.  During 
devolatilization,  loss  of  B  and  N  occurred 
as  the  elements  partitioned  from  B-  and  N- 
rich  phases  (e.g.,  micas)  into  H20-rich 
fluids  produced  primarily  by  chlorite- 
breakdown  reactions.  The  B  and  Be  con- 
centrations of  metasedimentary  and 
metamafic  rocks,  veins,  and  pegmatites  of 
the  Catalina  Schist  place  constraints  on 
models  that  invoke  addition  of  these  ele- 
ments to  arc  source  regions  by  "fluids" 
(H20-rich  solutions  or  silicate  liquids). 
The  varying  impact  of  hydrous  fluid  or 
melt  additions  to  arc  source  regions  could 
in  part  explain  the  variability  in  B/Be  of 
arc  volcanic  rocks,  including  cross-arc 
variability  in  some  individual  arcs.  Boron 
and  Be  may  serve  as  analogues  for  other 
incompatible  elements  (e.g.,  Cs  and  Zr, 
respectively)  which  appear  to  behave  simi- 
larly in  hydrous  fluid-solid  and  melt-solid 
systems. 


References 


Bebout,  G.  E.,  Field-based  evidence  for 
devolatilization  in  subduction  zones:  impli- 
cations for  arc  magmatism,  Science,  251,413- 
416,  1991. 

Bebout,  G.  E.,  and  M.  D.  Barton,  Fluid  flow  and 
metasomatism  in  a  subduction  zone  hydro  ther- 
mal system:  Catalina  Schist  terrane,  Califor- 
nia, Geology,  17,  876-980,  1989. 

Bebout,  G.  E.,  and  M.  L.  Fogel,  Nitrogen-isotope 
compositions  of  metasedimentary  rocks  in 
the  Catalina  Schist:  implications  for  meta- 
morphic  devolatilization  history,  Geochim. 
Cosmochim.  Acta,  in  press,  1991. 

Domanik,  K.,  R.  L.  Hervig,  and  S.  M.  Peacock, 
Beryllium  and  boron  contents  of  subduction 
zone  minerals :  an  ion  microprobe  study,  EOS, 
72,293-294,  1991. 

Leeman,  W.  P.,  D.  R.  Smith,  W.  Hildreth,  Z. 
Palacz,  and  N.  Rogers,  Compositional  diver- 


sity of  Late  Cenozoic  basalts  in  a  transect 
across  the  southern  Washington  Cascades: 
Implications  for  subduction  zone  magmatism, 
Jour.  Geophys.Res.,  95,  19561-19582,  1990. 

Leeman,  W.  P.,  V.  B.  Sisson,  and  M.  R.  Reid, 
Boron  geochemistry  of  the  lower  crust:  Evi- 
dence from  granulite  terranes  and  deep  crustal 
xenoliths,  Geochim.  Cosmochim.  Acta,  in  press, 
1991. 

Moran,  A.  E.,  V.  B.  Sisson,  and  W.  P.  Leeman, 
Boron  in  progressively  metamorphosed  oce- 
anic crust  and  sediments:  implications  for  com- 
positional variations  in  subducted  oceanic  slabs, 
Earth  Planet.  Sci.  Lett.,  in  press,  1991. 

Morris,  J.  D.,  W.  P.  Leeman,  and  F.  Tera,  The 
subducted  component  in  island  arc  lavas:  Con- 
straints from  Be  isotopes  and  B-Be  systemat- 
ics,  Nature,  344,  31-36,  1990. 

Nabelek,  P.  I.,  J.  R.  Denison,  and  M.  D.  Glascock, 
Behavior  of  boron  during  contact  metamor- 
phism  of  calc-silicate  rocks  at  Notch  Peak, 
Utah,  Amer.  Mineral,  75,  874-880,  1990. 

Piatt,  J.  P.,  Metamorphic  and  deformational  pro- 
cesses in  the  Franciscan  Complex,  California: 
Some  insights  from  the  Catalina  Schist  terrain, 
Geol.  Soc.  Amer.  Bull,  86,  1337-1347,  1975. 

Ryan,  J.  G.,  and  C.  H.  Langmuir,  Beryllium  sys- 
tematics  in  young  volcanic  rocks:  Implications 
for  10Be,  Geochim.  Cosmochim.  Acta,  52, 237- 
244,  1988. 

Ryan,  J.  G.,  and  C.  H.  Langmuir,  The  systematics 
of  boron  abundances  in  young  volcanic  rocks, 
Geochim.  Cosmochim.  Acta,  in  press,  1991. 

Shaw,  D.  M.,  M.  G.  Truscott,  E.  A.  Gray,  and  T.  A. 
Middleton,  Boron  and  lithium  in  high-grade 
rocks  and  minerals  from  the  Wawa- 
Kapuskasing  region,  Ontario,  Can.  Jour.  Ear. 
Sci.,25,  1485-1502,  1988. 

Sorensen,  S.  S.,  and  M.  D.  Barton,  Metasomatism 
and  partial  melting  in  a  subduction  complex: 
Catalina  Schist,  southern  California,  Geology, 
15,  115-118,  1987. 

Tera,  R,  L.  Brown,  J.  Morris,  I.  S.  Sacks,  J.  Klein, 
and  R.  Middleton,  Sediment  incorporation  in 
island  arc  magmas:  Inferences  from  10Be, 
Geochim.  Cosmochim.  Acta,  50, 535-550, 1986. 


30 


CARNEGIE  INSTITUTION 


Laser  Fluorination  of  Sulfide  Minerals 
with  F2  Gas 

D.  Rumble,  J.  M.  Palin,  and  T.  C. 
Hoering 

Excitement  pervades  the  discipline  of 
stable-isotope  geochemistry  at  the  pros- 
pects for  scientific  advances  raised  by  the 
development  of  new,  microanalytical  and 
in  situ  sampling  techniques.  Plans  can  now 
be  made,  with  reasonable  chances  of  suc- 
cess, to  measure  fundamental  indicators  of 
mass  transfer  mechanisms  in  Earth's  crust, 
i.e.  submillimeter-scale  gradients  in  isoto- 
pic  compositions.  Theoretical  studies  dem- 
onstrate it  is  the  curvatures,  slopes,  and 
magnitudes  of  composition  vs.  distance 
profiles  that  provide  definitive  evidence  on 
flux  magnitude  and  direction  and  the  rela- 
tive importance  of  diffusive  and  infiltrative 
mass  transfer  (Bickle  and  McKenzie,  1987; 
Blattner  and  Lassey,  1989;  Baumgartner 
and  Rumble,  1988;  Bowman  and  Willett, 
1991).  Experience  shows  that  advancing 
the  understanding  of  mass  transfer  in  na- 
ture depends  on  the  ability  to  resolve  differ- 
ences in  isotopic  compositions  over  small 
distances  (Bickle  and  Baker,  1990a,b). 
Isotopic  microanalysis  by  laser-heating 


minerals  immersed  in  a  reactive  atmosphere 
has  been  successfully  demonstrated  for  sili- 
cate, oxide,  and  sulfide  minerals.  Sharp 
(1990)  showed  that  laser  heating  of  small 
amounts  of  powdered  quartz,  olivine,  po- 
tassium feldspar,  garnet,  biotite,  musco- 
vite,  diopside,  and  magnetite  in  a  BrFs 
atmosphere  released  O2  with  8^0  values 
that  are  in  good  agreement  with  the  results 
of  conventional  BrF5  analyses.  The  in  situ 
analysis  of  spots  on  the  surfaces  of 
uncrushed  grains  of  quartz,  plagioclase, 
olivine,  and  magnetite  has  also  been  car- 
ried out  (Sharp,  1990;  Schiffries  and 
Rumble,  Annual  Report  1989-1990,  p.  37- 
40;  Elsenheimer  et  ai,  1990;  Conrad  and 
Chamberlain,  1991).  In  situ  analysis  for 
<534S  in  the  minerals  pyrite,  pyrrhotite, 
sphalerite,  galena,  and  chalcopyrite  has 
been  accomplished  by  laser  heating  and 
combustion  in  an  O2  atmosphere  and 
(Kelley  and  Fallick,  1990;  Crowe  et  al., 
1990;  Crowe  and  Shanks,  1991).  The  di- 
ameter of  analysis  spots  achieved  in  these 
studies  is  100-300  Jim,  with  a  precision  of 
±  0.2%o. 

We  are  currently  testing  the  efficacy  of 
F2  gas  in  laser  fluorination  to  analyze  sul- 
fide, silicate,  and  oxide  minerals  for  ^4S 
and  ^l  80.  It  is  desirable  to  develop  alterna- 


Table  8.  Precision  of  laser  fluorination  using  F2  gas  in  analyzing  mineral  powders 


Sample 


#4S 


533S 


#6S 


a 


CDT  +0.16  (±0.13)  17 

NBS-123  +17.62  (±0.15)  18 

E40  -26.72  (±0.3)  14 

M-pyrite  +3.17  (±0.3)  33 


+0.03 

(±0.06) 

8 

-4.1 

(±6.1) 

9 

+8.98 

(±0.13) 

9 

+34 

(±11) 

9 

-13.82 

(±0.16) 

7 

-33.7 

(±4.5) 

4 

+1.58 

(±0.17) 

22 

+24 

(±74) 

11 

o~=  Standard  Deviation 

CDT:  Canyon  Diablo  Troilite.  NBS-123:  Sphalerite.  E40:  Synthetic  Ag2S  prepared  from 
pyrrhotite;  three  wildly  errant  values  for  <5^6S  have  been  dropped.  M-pyrite:  Pyritehedrons,  1 
cm  in  diameter. 


GEOPHYSICAL  LABORATORY 


31 


tive  fluorination  reagents  in  order  to  opti- 
mize analysis  conditions.  Conventional 
analyses  of  silicates  for  <5180  show  that 
choosing  BrFs  or  F2  may  lead  to  incom- 
plete fluorination  and  inaccurate  analytical 
results  for  certain  minerals  (Taylor  and 
Epstein,  1962;  Clayton  and  Mayeda,  1963). 
In  the  laser  analysis  of  sulfides,  it  may 
prove  that  laser  fluorination  gives  less  frac- 
tionated analytical  results  than  laser  com- 
bustion with  O2  (cf.  Kelley  and  Fallick, 
1990;  Crowe  et  ai,  1990).  We  report, 
below,  verification  of  the  feasibility  of  F2 
gas  for  micro  analysis  of  5^4S,  cf^S,  and 
<536S  in  sulfide  minerals. 


Feasibility  of  Sulfide  Laser  Fluorination 
with  F2  Gas 

The  feasibility  of  using  F2  gas  as  fluori- 
nating  agent  to  produce  SF6  for  the  analysis 
of  533s,  #4S,  and  ^36s  during  laser  heat- 
ing of  sulfide  minerals  has  been  established 
by  repeated  analysis  of  aliquots  of  pow- 
dered troilite,  pyrite,  sphalerite,  and  syn- 


? 

v> 
« 

e 

CO 

00 


dSJ 

_  I 

1  1  1  1  1  1 

I     I    I    |    I     I    I    I    |    I    I     I 

* 
*                — 

- 

+   NBS-123 

10 



X    CDT 

• 

_ 

— 

O    E40-SO2 

* 

— 

- 

O  E40-BrF5 

♦ 

— 

0 

— 

.«* 

- 

-10 

** 

— 

• 

- 

* 
* 

- 

-20 

— 

™ 

—                 * 

-*r*\  1  1  1  i  1  1  1  1  I  1  1  1  1  I  1  1  1 

Illl1 

-30 


-20 


-10  0 

534S,  accepted 


10 


20 


Fig.  12.  Plot  of  accepted  vs.  measured  values  of 
#4SforNBS-123  sphalerite,  CDT-Canyon  Diablo 
Troilite,  and  E40  -  synthetic  Ag2S  prepared  from 
pyrrhotite  and  analysed  by  conventional  combus- 
tion (SO2)  or  conventional  fluorination(BrF5). 

thetic  Ag2S.  The  precision  of  analysis  for 
troilite  and  sphalerite  is  ±  0.06-0. \5%c  for 
S&S  and  #4S  and±  6-1  Wcfor  &£>S  (Table 
8).  The  precision  for  pyrite  and  synthetic 
Ag2S  is  ±  0. 1 6-0  3%o  for  S^S  and  #4S  and 
±  27-74%o  for  ^S  (Table  8).  The  preci- 
sion for  534S  is  comparable  to  that  obtained 
by  laser  oxidation/combustion  of  sulfide 
minerals  in  an  O2  atmosphere  (Crowe  et 
al.,  1990).  The  scatter  in  the  results  for 
pyrite  may  be  explained,  at  least  in  part,  by 
the  fact  that  the  mineral  was  used  for  prac- 


Table  9.  Accuracy  of  laser  fluorination  using  F2  gas.  Mineral  powders. 


&*S 

Sample 

(measured) 

(accepted) 

(measured- accepted) 

CDT 

NBS-123 

E40 

E40* 

+0.16 
+17.62 
-26.72 
-26.72 

0.0 
+17.09 
-27.12 
-26.88 

+0.16 
+0.53 
+0.40 
+0.16 

CDT:  Canyon  Diablo  Troilite;  standard  #4S  value  defined  asO.O.  NBS-123:  Sphalerite: 
Value  of  +17.09  reported  as  average  of  results  of  intercomparsion  of  1 1  laboratories. 
(International  Atomic  Energy  Agency,  1986).  E40:  Synthetic  Ag2S  prepared  from 
pyrrhotite.  Accepted  value  analyzed  by  conventional  combustion  (Oliver  et  al.,  Annual 
Report  1989-1990,  p.  30-33).  E40':  Synthetic  Ag2S  prepared  from  pyrrhotite.  Accepted 
value  analyzed  by  conventional  BrFs  fluorination  (Oliver  et  al.,  Annual  Report  1989- 
1990,  p.  30-33). 


CARNEGIE  INSTITUTION 


2.0 


_  I  I  I  I  |  I  I  I  I  J...!...!...!...!...|...!...j...j....i...i...i....i....i.%  i  - 


1.5    - 


£  1.0  ^ 

o 


W  «     IT 

S     0.5 


-0.5 


+  NBS-123 

X  CDT 

O  E40-SO2 

O  E40-BrFs  (shift  +2  along  X-axis) 


o.o  E-  Q 


X 

N 

x 


- 1  i  i  i  i  i  i  i  i  i  i  i  i  i  i  i  i  i  i  i  i  i  i  i- 


-30 


-20 


-10  0 

634S,  accepted 


10 


20 


Fig.  13.  Plot  of  accepted  values  of  $4S  vs.  the 
difference  between  measured  and  accepted  values 
of^S. 

tice  in  learning  how  to  use  F2  gas.  The 
scatter  in  results  for  synthetic  Ag2S  has  not 
yet  been  fully  explained  but  may  be  related 
to  the  relatively  smaller  amount  of  avail- 
able sulfur  as  constrained  by  stoichiom- 
etry.  The  accuracy  of  the  results  was  evalu- 
ated by  comparing  5^4S  of  SF6  derived 
from  F2  laser  fluorination  of  CDT,  NBS- 
123,  and  E  40  to  "accepted"  #4S  values 
obtained  by  conventional  combustion  with 
O2  and  fluorination  with  BrF5.  The  results 
show  deviations  of  from  +0.16  to  +0.53%c 
in  <P4S  for  F2  laser  fluorination  samples 
relative  to  accepted  values  (Figs.  12,  13, 
Table  9). 

The  experimental  apparatus  used  for 
F2  laser  fluorination  resembles,  in  general 
outline,  that  used  by  Sharp  (1990)  for  BrF5 
fluorination  of  silicate  minerals.  There  has 
been  added,  however,  a  fluorine  gas  gen- 
eration and  delivery  system,  a  heated  KBr 
trap  to  dispose  of  excess  F2,  and  an  Inconel 
capacitance  manometer  to  measure  yields 
and  to  monitor  the  progress  of  reactions 
and  the  course  of  cryogenic  transfers  of 
condensable  gases.  Fluorine  is  generated 
by  heating  the  compound  K2NiF6  •  KF  in 
an  evacuated  nickel  reservoir  to  tempera- 


tures of  290°-320°C  (Asprey,  1976).*  El- 
emental fluorine  is  incorporated  at  low 
temperatures  (~250°C)  and  released  at 
higher  temperatures  according  to  the  re- 
versible reaction  (Asprey,  1976) 

2(K2NiF6  •  KF)  =  2K3N1F6  +  F2      (1) 
solid  solid       gas 

Use  of  the  compound  solves  two  problems 
encountered  in  fluorine  chemistry.  (1)  F2 
gas  can  be  obtained  free  from  contamina- 
tion by  N2  or  O2.  We  outgas  the  compound 
at  100°C  under  vacuum  prior  to  evolution 
of  F2.  (2)  The  hazard  of  handling  large 
quantities  of  F2  gas  is  eliminated.  We 
generate  only  the  small  amounts  of  F2 
needed  for  laser  fluorination.  Excess  F2 
can  be  resorbed  by  the  reversal  of  reaction 
(1)  at  lower  temperature. 

The  experimental  procedure  consists  of 
generating  F2  gas,  expanding  it  into  an 
evacuated  sample  chamber  loaded  with 
samples,  and  aiming  and  firing  the  laser. 
The  reaction  chamber  is  open  to  a  liquid 
nitrogen  cold  trap  to  remove  SF6  cryogeni- 
cally  from  the  reaction  site  as  soon  as  it  is 
formed.  After  laser  heating,  excess  F2  is 
removed  by  reaction  with  KBr.  The  prod- 
uct SF6  is  held  back  in  a  cold  trap  to  avoid 
mixing  it  with  the  bromine  formed  when  F2 
reacts  with  KBr.  The  apparatus  is  not  fully 
optimized  for  laser  fluorination  of  sulfides 
because  at  this  step  of  the  procedure  prod- 
uct SF6  must  be  removed  in  a  Ni-metal 
container  and  carried  to  another  vacuum 

1  We  are  indebted  to  J.  O'Neil  for  alerting  us  to  the 
existence  of  the  K2NiF6  •  KF  compound,  to  G.  P. 
Landis  for  furnishing  unpublished  information  on 
its  use,  and  to  Ozark- Mahoning,  Inc.  for  manufac- 
turing it. 


GEOPHYSICAL  LABORATORY 


33 


line  for  purification.  The  purification  con- 
sists of  reacting  SF6  with  moist  KOH  to 
eliminate  traces  of  F2  and  HF  followed  by 
gas  chromatography  to  remove  trace  con- 
taminants that  give  isobaric  interferences 
with  the  ion  beams  of  SF5+  in  the  mass 
spectrometer  (Hoering,  Annual  Report 
1989-1990,  p.  128-131). 

Experimental  parameters,  such  as  pres- 
sure of  F2  gas  and  laser  power,  were  varied 
systematically  to  establish  optimal  operat- 
ing conditions.  Pressure  of  F2  gas  in  the 
reaction  chamber  was  varied  from  55  to 
175  torr.  Fluorination  proceeds  readily  at 
lower  pressures  and  is  preferred  to  econo- 
mize on  consumption  of  the  reagent.  The 
20-watt,  CO2  laser  was  operated  in  both 
continuous  and  pulsed  mode.  In  the  analy- 
sis of  powders,  minimum  power  was  used 
in  order  to  protect  the  fragile  BaF2  win- 
dows from  damage.  For  most  samples, 
pulsed  operation  with  a  pulse  spacing  of  10 
milliseconds  and  pulse  width  of  10  milli- 
seconds at  a  laser  power  setting  of  20%  is 
adequate  to  achieve  complete  fluorination. 
An  additional  problem  is  encountered  if  the 
initial  laser  shot  is  set  at  high  power:  the 


10 
5 
0 
-5 


_i  i  i  i  |  i  i  i  i  |  i  i  i  i  |  i  i  i  i  |  i  i  i  i_ 

.4? 


j¥ 


<c*» 


<*6 


■  r^  •  *  * 


-10     - 

_  * 

.15  Eg-  i  i  i  1  i  1  1  i  1  1  1  1  i  i  i  1  i  i  1  1  1  1- 


►£•  Sulfide  Powders 


-30  -20  -10 


10  20 


534S 


Fig.  14.  Plot  of  $3S  vs.  #4S  showing  measured 
vs.  theoretical  mass  fractionation. 


powdered  sample  may  be  scattered  around 
the  sample  chamber.  The  problem  may  be 
mitigated  by  starting  the  laser  at  low  power 
and  low  pulse  width  and  increasing  incre- 
mentally until  reaction  is  seen.  It  is  diffi- 
cult to  avoid  some  scattering  of  powdered 
samples.  For  this  reason,  yields  of  SF6  in 
relation  to  weighed  amounts  of  samples  are 
less  than  stoichiometric. 

Analysis  for  four  sulfur  isotopes,  32s, 
33S,  34S,  and  36s,  is  made  simultaneously 
on  aFinnigan-M AT '251  mass  spectrometer 
with  custom  quadruple  collector.  The  mea- 
surement of  all  the  isotopes  provides  a 
useful  test  for  the  quality  of  the  analysis 
(Fig.  14). 

The  theory  of  isotope  fractionation  in 
chemical  processes  predicts  that  the  ratio 
<534SA533s  =  1.94  (±0.01)  in  samples  such 
as  terrestrial  ores,  where  anomalous  nucleo- 
synthetic  effects  are  absent  (Hulston  and 
Thode,  1965).  We  obtained  a  value  of 
1.950  (±  0.003)  for  the  ratio  (Fig.  14). 

The  results  reported  above  validate  the 
use  of  F2  gas  in  laser  fluorination  of  sul- 
fides. The  laser  fluorination  of  sulfides  is 
far  quicker  than  conventional  fluorination: 
reactions  are  completed  in  seconds  or  min- 
utes rather  than  overnight.  In  comparison 
to  laser  combustion  of  sulfide  minerals  in 
O2  (Crowe  et  al.  1 990),  laser  fluorination  is 
slower,  because  of  the  preparative  gas  chro- 
matography that  is  required  to  eliminate 
isobaric  interferences  in  the  mass  spec- 
trometer. The  advantages  of  the  method 
are  clear,  however.  One  obtains  precise 
data  on  three  of  the  four  isotopes  32S,  33S, 
and  34S.  The  additional  uncertainties  in- 
troduced by  oxygen  isotope  corrections  in 


34 


CARNEGIE  INSTITUTION 


mass  spectrometry  of  SO2  are  eliminated, 
for  fluorine  has  only  one  stable  isotope. 
Because  SF6  is  chemically  inert,  does  not 
sorb  onto  the  interior  surfaces  of  vacuum 
lines,  and  is  not  sensitive  to  moisture,  it  is 
potentially  a  better  working  gas  than  SO2 
for  mass  spectrometry  of  very  small  samples 
(Rees,  1978). 


References 

Asprey,  L.  B.,  The  preparation  of  very  pure  fluo- 
rine gas,  J.  Fluorine  Chem.,  7,  359-361,  1976. 

Baumgartner,  L.  P.,  and  D.  Rumble,  Transport  of 
stable  isotopes.  I.  Development  of  a  kinetic 
continuum  theory  for  stable  isotope  transport, 
Contrib.  Mineral.  Petrol,  98,  417-430,  1988. 

Bickel,  M.,  and  D.  McKenzie,  The  transport  of 
heat  and  matter  by  fluids  during  metamor- 
phism,  Contrib.  Mineral.  Petrol.,  95,  384-392, 
1987. 

Bickel,  J.,  and  J.  Baker,  Advective-diffusive  trans- 
port of  isotopic  fronts:  an  example  from  Naxos, 
Greece,  Earth  Planet.  Sci.  Lett.,  97,  78-93, 
1990a. 

Migration  of  reaction  and  isotopic 

fronts  in  infiltration  zones:  Assessments  of  fluid 
flux  in  metamorphic  terrains,  Earth  Planet.  Sci. 
Lett.,  98,  1-13,  1990b. 

Blattner,  P.,  and  K.  R.  Lassey,  Stable-isotope 
exchange  fronts,  Damkohler  numbers,  and  fluid 
torockratios,  Chem.  Geoi,  78,  381-392, 1989. 

Bowman,  J.  R,  and  S.  D.  Willett,  Spatial  patterns 
of  oxygen  isotope  exchange  during  one  dimen- 
sional fluid  infiltration,  Geophys.  Res.  Lett.,  18, 
971-974,  1991. 

Clayton,  R.N.,  and  T.  K.  Mayeda,  The  use  of  BrF5 
in  the  extraction  of  O2  from  oxides  and  silicates 
for  isotopic  analysis,  Geochim.  Cosmochim. 
Acta,  27,  43-52,  1963. 

Conrad,  M.  E.,  and  C.  P.  Chamberlain,  Laser- 
based  analyses  of  small-scale  variations  in  the 
oxygen  isotope  ratios  of  hydrothermal  quartz, 
£05,72,292,1991. 

Crowe,  D.  E.,  J.  W.  Valley,  and  K.  L.  Baker, 
Micro-analysis  of  sulfur-isotope  ratios  and  zo- 
nation  by  laser  microprobe,  Geochim. 
Cosmochim.  Acta,  54,  2075-2092,  1990. 

Crowe,  D.  E.,  and  W.  C.  Shanks,  Laser  micro- 
probe  $4S  study  of  coexisting  sulfide  pairs: 
seeing  through  metamorphism,  EOS,  72,  1991. 

Elsenheimer,  D.,  J.  W.  Valley,  and  K.  Baker,  In- 
situ  laser  microprobe  determinations  of  (5180, 


GSAAbstractsw. Programs, 22, 160-161, 1990. 

Hulston,  J.  R.,  andH.  G.,  Thode,  Variations  in  the 
33S,  34S,  and  36S  contents  of  meteorite  and  their 
relation  to  chemical  and  nuclear  effects,  /. 
Geophys.  Res.,  70,  3475-3484,  1965. 

Kelley,  S.  P..  and  A.  E.  Fallick,  High  precision 
spatially  resolved  analysis  of  £*4S  in  sulfides 
using  a  laser  extraction  technique,  Geochim. 
Cosmochim.  Acta,  54,  883-888,  1990. 

Rees,  C.  E.,  Sulfur  isotope  measurements  using 
SO2  and  SF6,  Geochim.  Cosmochim.  Acta,  42, 
383-389,  1978. 

Rumble,  D.,  T.  C.  Hoering,  and  J.  M.  Palin,  Mi- 
croanalysis for  $AS  in  sulfide  minerals  with 
laser  fluorination,  EOS,  72,  292,  1991. 

Sharp,  Z.  D.,  A  laser-based microanalytical  method 
for  the  in  situ  determination  of  oxygen  isotope 
ratios  of  silicates  and  oxides,  Geochim. 
Cosmochim.  Acta,  545,  1353-1357,  1990. 

Taylor,  H.P.,  and  S.  Epstein,  Relationships  be- 
tween 180/160  ratios  in  coexisting  minerals  of 
igneous  and  metamorphic  rocks,  Geol.  Soc. 
Amer.  Bull.,  73,  461-480,  1962. 


Stable  Isotope  and  Trace  Element  Indi- 
cators of  Devolatilization  History  in 
Metashales  and  Metasandstones 

Gray  E.  Bebout 

Devolatilization  of  carbonate-poor 
metashales  and  metasandstones  has  the  po- 
tential to  release  large  amounts  of  H2O- 
rich  C-O-H-S-N  fluid  (e.g.,  Walther  and 
Orville,  1982).  However,  few  studies  have 
directly  examined  this  process,  in  contrast 
with  the  many  studies  that  have  dealt  with 
devolatilization/inf iltration  in  meta-carbon- 
ate  systems  (e.g.,  Valley,  1986).  Because 
metashale  and  metasandstone  make  up  a 
significant  fraction  of  the  continental  crust, 
their  devolatilization  may  be  extremely 
important  for  crustal  chemical,  thermal, 
and  rheological  evolution.  The 
metasedimentary  rocks  of  the  Catalina 
Schist  (California)  are  well  suited  for  a 


GEOPHYSICAL  LABORATORY 


35 


study  of  devolatilization;  rocks  of  similar 
bulk  composition  have  been  metamor- 
phosed under  a  wide  range  of  metamor- 
phic  conditions  (350°-750°C,  5-11  kbar; 
Piatt,  1975;  Sorensen  and  Barton,  1987; 
Bebout  and  Barton,  1989).  Trace  element 
and  stable  isotope  compositions,  mineral 
modes,  and  mineral  compositions  show 
distinct  covariance  with  increasing  meta- 
morphic  grade.  In  this  study,  the  integra- 
tion of  petrologic  and  geochemical  data 
results  in  a  distillation  model  for  progres- 
sive devolatilization.  This  model  neces- 
sarily implies  specific  devolatilization 
mechanisms  which  have  consequences  for 
models  of  crustal  chemical  evolution,  fluid 
transport  dynamics,  and  metamorphic  re- 
action kinetics. 

The  three  major  metamorphic  units  of 
the  Catalina  Schist  (lawsonite-albite/ 
blueschist,  glaucophanic  greenschist/epi- 
dote-amphibolite,  and  amphibolite)  con- 
tain sedimentary,  mafic,  and  ultramafic 
rocks  underplated  and  metamorphosed 
during  early  Cretaceous  subduction  (Piatt, 
1975;  Bebout,  1991a).  Metasedimentary 
rocks  from  the  three  units  show  a  similar 
range  in  lithology  from  metapelites  to  meta- 
graywackes;  average  grain  size  increases 
with  increasing  grade  (several  \im  to  sev- 
eral mm).  Trends  in  H2O  content  with 
increasing  metamorphic  grade  and  pro- 
grade  reaction  histories  inferred  from  min- 
eral modes  indicate  that  devolatilization  of 
the  metasedimentary  rocks  of  the  Catalina 
Schist  principally  involved  chlorite  break- 
down reactions  over  the  approximate  tem- 
perature interval  400°-600°C  (Bebout, 
1991a).  H2O  content,  determined  by  H- 
isotope  extraction  techniques,  decreases 


from  2-5.5  wt  %  in  lowest-grade  rocks  to 
1.5-2.5  wt  %  in  highest-grade,  amphibolite- 
facies  rocks  (Bebout,  1991a).  Chlorite 
breakdown  reactions  resulted  in  the  produc- 
tion of  muscovite-,  biotite-,  garnet-,  and 
kyanite-bearing  mineral  assemblages 
through  reactions  of  the  following  general 
types: 

2Phengite  +  Chlorite  =  Muscovite 

+  Biotite  +  Quartz  +  4H2O  (1) 

2Chlorite  +  4Quartz 

=  3Garnet  +  8H2O  (2) 

3  Chlorite  +  7  Muscovite  +  Quartz 

=  Al2Si05(Kyanite)  +  7Biotite 

+  12H20  (3) 

Chlorite  +  Muscovite  =  Biotite  + 
Al2Si05(Kyanite)  +  Quartz  +  8H2O     (4) 

In  addition  to  the  decrease  in  H2O  con- 
tent, devolatilization  resulted  in  preferential 
loss  of  some  trace  elements  and  in  shifts  in 
the  C  and  N  isotope  compositions  of  the 
rocks.  The  metasedimentary  rocks  show 
trends  of  decreasing  N  concentration  and 
increasing  whole-rock  8  5N  with  increas- 
ing metamorphic  grade  (Bebout  and  Fogel, 
Annual  Report  1989-1990,  p.  19-26;  Fig. 
15).  The  £13C  of  carbonaceous  matter  in 
these  rocks  increases  from  values  of  from 
-26  to  -24  %o  in  the  lowest-grade  rocks  to 
values  of  from  -21  to  -19  %o  in  the  highest- 
grade,  amphibolite-facies  rocks  (Fig.  15). 
Concentrations  of  carbonaceous  matter  do 
not  vary  systematically  with  increasing  grade 
and  average  -0.6  wt  %  for  all  grades.  De- 
spite metamorphism  at  temperatures  ex- 


36 


CARNEGIE  INSTITUTION 


■18 


o 

CO 


-20   - 


§    -22 

Q. 


-24 


-26 


-28 


Lawsonite-     • 
Albite  & 
Blueschist 

JL 


Greenschist  & 

Epidote 

Amphibolite 


j i_ 


j i_ 


515N 


4 
Air 


Fig.  15.  Whole-rock  515N  vs.  <513C  of  carbon- 
aceous matter  for  metasedimentary  rocks  of  the 
Catalina  Schist.  Note  parallel  trends  toward  higher 
isotope  values. 

ceeding  350°C,  C/N  of  the  lowest-grade 
metasedimentary  rocks  (5-20;  mean  -13) 
is  similar  to  C/N  of  many  unmetamorphosed 


sedimentary  rocks,  including  those  in  trench 
and  off-trench  environments  (cf.  Patience 
et  al.,  1990).  Higher-grade  rocks  have 
higher  C/N  that  is  attributable  to  preferen- 
tial N  loss  (for  6  greenschist  and  epidote 
amphibolite  samples,  range  is  from  5  to 
1 25;  for  four  amphibolite  samples,  range  is 
from  28  to  237).  Boron  concentration  in 
the  same  rocks  decreases  with  increasing 
metamorphic  grade  from  an  average  of -73 
ppm  in  lawsonite-albite  grade  rocks  to  an 
average  of  -12  ppm  in  amphibolite -grade 
rocks  (Fig.  16;  Bebout  et  a/.,  this  Report). 
Bebout  and  Fogel  (Annual  Report  1989- 
1990,  p.  19-26)  interpreted  the  N  concen- 
tration and  isotope  data  to  be  results  of 
Rayleigh  distillation  behavior  during  pro- 
gressive devolatilization  (see  equation  [4] 
in  Valley,  1986).  A  remaining  difficulty  in 
the  Rayleigh  distillation  calculations  stems 
from  variability  in  composition  within  an 
individual  grade.  This  variability,  which  in 


1000 


800 


E 

Q. 


^    600 


CD 

D) 

O 


400  - 


200 


Lawsonite-Albite  & 
Blueschist  (• ) 


Greenschist  &  Epidote 
Amphibolite  (A ) 


Amphibolite  (o ) 


100 

Boron  (ppm) 


200 


Fig.  16.  N  and  B  concentrations  of  metasedimentary  rocks  of  the  Catalina  Schist. 
See  discussion  of  the  significance  of  B  loss  for  subduction  zone  recycling  in 
Bebout  et  al.  (this  Report). 


GEOPHYSICAL  LABORATORY 


37 


E 

CL 
Q. 

C 
CD 
D) 
O 


800 


600 


400 


200 


Mean  F  Calculated 

from  Sample  Pairs 

0.26  +  0.07 

(n  =  12) 


Lawsonite-Albite  & 
Blueschist 


F       0.4  - 


K>0  (weight  %) 

Fig.  17.  Relationship  of  N  and  K2O  concentra- 
tions in  metasedimentary  rocks  of  the  Catalina 
Schist.  Tie  lines  connect  samples  with  similar 
K2O  content  and  demonstrate  that  high-grade 
samples  contain  0.26  ±  0.07  of  the  N  in  low-grade 
samples  with  similar  K  content;  calculated  F  is 
shown  for  several  of  the  tie  lines.  The  mean  8  N 
shift  between  low-  and  high-grade  samples  in 
these  same  pairs  is  1.85%c  (standard  deviation  of 

1.0%©). 

part  represents  protolith  variability,  compli- 
cates comparisons  of  trace  element  compo- 
sition and  volatile  content  of  low-  and  high- 
grade  metamorphic  equivalents.  The 
protolith  variability  problem  can  be  allevi- 
ated by  normalization  of  the  N  data  to  K2O 
data  for  the  same  rocks.  Because  N  is 
preferentially  partitioned  into  micas  (Honma 
and  Itihara,  1981),  and  because  the  micas 
constitute  the  only  significant  K  reservoir  in 
the  rocks,  the  concentrations  of  N  correlate 
well  with  K2O  content  of  the  rocks  (Fig. 
17).  By  comparing  only  samples  with  simi- 
lar K2O  content  (i.e.,  with  similar  modal 
abundance  of  micas),  fluid-rock  isotope  frac- 
tionation factors  can  be  more  tightly  con- 
strained, as  shown  in  Fig.  18. 


Fluid  -  Rock 


Fig.  18.  Demonstration  of  the  interdependencies 
of  8  N  shift  due  to  Rayleigh  distillation  (labelled 
curves),  fluid-rock  N-isotope  fractionation 
(A15Nfiuid-rqck;  related  to  alpha,  the  fractionation 
factor,  by  8 ^Ifiuid  -  8  Nrock  =  103  In  alpha),  and 
inferred  N  loss  (F  indicates  the  fraction  of  the 
original  N  remaining  in  the  rock.)  Indicated  are 
estimates  of  N  loss  %  based  on  differences  in 
mean  N  content  in  the  lawsonite-albite  and  am- 
phibolite  facies  rocks  (0.22;  dashed  line)  and 
based  on  normalizations  by  use  of  the  K2O  data 
indicates  an  F  or  ~0.26±0.07;  see  solid  horizontal 
lines  that  indicate  mean  ±  one  standard  devia- 
tion). Curves  are  for  the  mean  of  the  8  N  shifts 
for  the  pairs  with  similar  K2O  content  (±  one 
standard  deviation  on  lower  and  higher  8  N 
sides;  1.85  ±  1 .0  %p-  solid  curves),  and  the  differ- 
ence in  the  mean  8  N  of  the  lawsonite-albite  and 
amphibolite  samples  (dashed  curve  labeled  2.4%o; 
see  Bebout  and  Fogel,  Annual  Report  1989-1990, 
p.  19- 26  for  data).  Shaded  region  indicates  range 
of  fractionations  (-1.511.0  %6)  compatible  with 
the  Catalina  data  based  on  these  data  for  N  loss 
and  8  N  shift.  Vertical  arrows  indicate  frac- 
tionations calculated  by  use  of  differences  in 
mean  N  and  8  N  of  the  low-  and  high-grade 
rocks  (—1.6  %o)  and  by  use  of  the  K^O-normal- 
ized  data  (—1.4  %o). 

Values  of  A15Nfiuid-rock  (<515Nfiuid  - 
<515Nrock)  of  about  -1 .5+1 .0%o  inferred  in 
Fig.  1 8  are  similar  to  values  of  from  -1 .0  to 
-4.0%o  suggested  by  Haendel  et  ai  (1986) 
based  on  isotope  and  concentration  shifts 
in  metamorphic  suites,  but  are  smaller 


38 


CARNEGIE  INSTITUTION 


than  ranges  of  from  -3.0  to  -4.0  %oand  -6.0 
to  -11.5  %o  obtained  by  Kreulen  et  al 
(1982)  and  Bottrell  et  al  (1988),  respec- 
tively, for  fluid  inclusion-mineral/rock  pairs 
in  low  to  medium  grade  metasedimentary 
rocks.  The  range  of  fractionation  factors 
indicated  in  this  study  is  similar  to  the 
range  calculated  by  Hanschmann  (1981) 
for  N2-NH4"1"  exchange  equilibrium  at  400°- 
600°C  (A15Nfiuid-rock  of  from  -2.25  to  -2.9 
%o  for  the  temperature  range  of  400°- 
600°C).  Speciation  of  N  as  N2  in  the 
Catalina  fluids  is  further  suggested  by  the 
presence  of  small  amounts  of  N2  in  fluid 
inclusions  from  metasomatized  eclogitic 
blocks  in  blueschist  melange  (<1  mol  %;  T. 
C.  Hoering,  personal  communication,  1991; 
determined  by  quadrupole  mass  spectrom- 
etry). Furthermore,  calculations  of  fluid  C- 
O-H-N  equilibria  indicate  that  N2  is  the 
dominant  N  species  under  most  crustal 
metamorphic  conditions  (Duit  etal,  1986; 
Ferry  and  Baumgartner,  1987;  Bottrell  et 
al.,  1988). 

Carbon  and  O  isotope  data  for  the  same 
rocks  are  compatible  with  the  model  of 
Rayleigh  distillation  derived  from  the  N 
data.  Progressive  fractionation  of  12C  from 
carbonaceous  matter  into  CH4  in  fluids  by 
a  Rayleigh  distillation  process  could  pre- 
sumably explain  the  observed  shift  in  <513C 
(Fig.  15;  seeBebout,  1989).  Because  of  the 
relatively  large  A13C  of  CH4-graphite  ex- 
change at  these  temperatures  (-3  to  -7  %o\ 
see  Bottinga,  1969),  such  a  shift  in  £13C 
could  be  achieved  without  a  large  decrease 
in  whole -rock  C  concentration.  Heating/ 
freezing  experiments  (Sorensen  and  Barton, 
1987)  and  quadrupole  mass  spectrometry 
indicate  that  CH4  is  the  dominant  C  species 


(<1-10  mol  %)  in  fluid  inclusions  from 
metasomatized  mafic  blocks  in  blueschist 
and  amphibolite  unit  melange.  The  unifor- 
mity of  <5180  of  the  metasedimentary  rocks 
at  all  grades  (quartz  <5180  of  approximately 
+16  to  +19  %o;  Bebout,  1991a)  indicates 
that  the  metasedimentary  rocks  did  not 
pervasively  reequilibrate  with  the  H2O- 
rich  fluids  that  produced  veins  and  infil- 
trated more  permeable  portions  of  the 
Catalina  Schist  (Bebout,  1991a,  b).  This 
inferred  relative  impermeability  of  the 
metasedimentary  rocks  supports  the  as- 
sumption of  closed- system  behavior  im- 
plicit in  the  Rayleigh  distillation  calcula- 
tions. 

The  Rayleigh  calculations  assume  that 
isotopic  equilibrium  is  maintained  between 
the  fluid  phase  and  the  remaining  mineral 
phases  and  among  mineral  phases  during 
incremental  loss  of  fluid.  In  the  Catalina 
metasedimentary  rocks,  the  mechanisms 
affording  this  continual  reequilibration  pre- 
sumably involved  diffusive  exchange  and 
dissolution/reprecipitation  during  grain 
coarsening.  The  increase  in  white  mica 
grain  size,  a  trend  in  white  mica  chemistry 
{decrease  in celadonite  substitution,  [(Mg, 
Fe2+)  +  Si  =  2A1];  Sorensen,  1986;  Bebout, 
1989},  and  increases  in  the  grain  size  and 
degree  of  crystallinity  of  carbonaceous  mat- 
ter (Bebout  1 989)  presumably  reflect  these 
processes.  For  other  rocks,  Duit  et  al 
(1986)  reported  coupled  decreases  in  mica 
and  whole-rock  N  concentration  with  in- 
creasing metamorphic  grade.  Such  de- 
creases are  consistent  with  progressive  par- 
titioning of  N  into  fluids  equilibrated  with 
the  remaining  micas,  as  opposed  to  selec- 
tive release  of  N  due  to  mica  breakdown 


GEOPHYSICAL  LABORATORY 


39 


(i.e.,  with  the  remaining  mica  retaining  its 
original  higher  N  content) .  Similarly,  in  an 
ion  microprobe  study  of  the  rocks  ana- 
lyzed in  this  study,  Domanik  et  al.  (1991) 
report  a  general  decrease  in  the  B  concen- 
tration of  white  micas  with  increasing  meta- 
morphic  grade. 

The  proposed  model  requires  that  flu- 
ids released  by  devolatilization  escape 
without  appreciable  infiltration  by  exter- 
nally-derived fluids  significantly  out  of 
0-,  C-,  or  N-isotopic  equilibrium  with  the 
rocks.  Such  a  scenario  might  arise  as 
locally-derived  fluids  migrate  within  lay- 
ers of  similar  composition  and  at  similar 
P-Tconditions  toward  fractures  along  pore 
fluid  pressure  gradients  (cf.  Etheridge  et 
al.,  1984).  Veins,  which  represent  the 
larger  fractures,  are  abundant  in  the  Catalina 
Schist,  particularly  at  lower  grades  (Bebout 
and  Barton,  1989).  The  fluids  in  this 
scenario  would  always  be  equilibrated  with 
the  rocks  along  their  flow  paths  and  would 
not  leave  a  significant  imprint  on  trace 
element  and  stable  isotope  compositions 
in  the  host  rock.  However,  flow  across 
layering  (e.g.,  of  mixed  sandstone  and 
shale  sequences)  would  possibly  homog- 
enize elemental  and  isotopic  variations 
that  resulted  from  local  devolatilization 
history  or  protolith  variability,  depending 
on  the  magnitude  of  the  variations,  fluid- 
to-rock  ratios,  and  fluid  composition.  The 
degree  of  homogenization  would  depend 
in  part  on  fracture  density,  which  would 
control  the  scale  of  the  intergranular  flow, 
and  the  relative  permeabilities  of  the 
interlayered  lithologies,  which  would  dic- 
tate whether  or  not  intergranular  flow  oc- 
curred across  layering.  To  address  further 


these  issues,  ongoing  research  involves 
analyses  of  trace  element  concentrations 
and  stable  isotope  composition  in  finely 
interlayered,  veined  sequences  of  lawsonite- 
albite  and  blueschist  grade  metasandstone 
and  metashale.  Preliminary  data  indicate 
the  preservation  of  -1.0  %o  gradients  in 
8  5N  within  one  finely  interlayered  (cm- 
scale)  lawsonite-albite  grade  meta-sedimen- 
tary  exposure. 

For  carbonate-poor  metamorphosed 
sandstones  and  pelitic  rocks,  N  and  C  (and 
perhaps  S)  isotope  systematics  appear  to  be 
more  useful  measures  of  the  extent  of 
devolatilization  than  either  the  H  or  O  sys- 
tems, owing  to  a  more  favorable  mass- 
balance  between  the  fluids  and  rocks.  Be- 
cause of  the  large  reservoir  of  H  in  high- 
X(H20)  fluids  and  the  low  H/O  of  the  rocks, 
trends  in  H  isotope  compositions  due  to 
fluid  loss  are  probably  modified  more  easily 
by  interaction  with  the  small  amounts  of 
fluid  derived  within  local  fluid  generation/ 
extraction  systems.  The  Catalina  meta- 
sedimentary  rocks  show  no  obvious  trend  in 
SD  with  increasing  metamorphic  grade  and 
range  from  -80  to  -50  %oat  all  grades  (Bebout, 
1989).  Because  of  the  large  O  reservoir  in 
the  rocks,  Rayleigh  distillation  results  in 
only  minor  change  in  whole-rock  5  °0  with 
progressive  devolatilization;  this  change  is 
probably  unresolvable  given  the  marked 
protolith  variability  in  <5180.  Similar  argu- 
ments were  made  by  Oliver  et  al.  (Annual 
Report  1989-1990,  p.  30-33)  for  S-isotope 
systematics  in  sulfidic  schists  from  the 
Waterville-Augusta  area,  Maine.  Because 
of  the  low  S  content  in  the  metamorphic 
fluids  (2-5  mol  %;  Ferry,  1981),  greenschist- 
grade  metamorphic  rocks  retained  S-iso- 


40 


CARNEGIE  INSTITUTION 


tope  signatures  inherited  from  diagenetic 
and  lower-grade  metamorphic  de- 
volatilization  histories.  These  arguments 
are  also  analogous  to  those  made  for  O  and 
C  isotope  systematics  during  decarbon- 
ation  of  impure  carbonates  infiltrated  by 
H20-rich,  C-poor  fluids.  Because  of  the 
low  C  content  of  the  infiltrating  fluids,  C- 
isotope  composition  of  the  rocks  is  re- 
garded as  a  better  measure  of  the  extent  of 
decarbonation  than  the  O-isotope  system- 
atics, which  are  commonly  controlled  by 
infiltration  (Rumble,  1982;  VaUey,  1986). 
A  distillation  mechanism  like  that  pro- 
posed for  the  Catalina  N  data  may  dictate 
the  mobilities  of  other  fluid-mobile  trace 
elements  (e.g.,  B,  Cs,  U;  see  Heier,  1973; 
Leeman  et  aL,  1990)  residing  in  a  range  of 
mineral  reservoirs  (i.e.,  not  only  in  micas, 
which  are  the  predominant  N  and  B  reser- 
voir). The  efficiency  with  which  equilib- 
rium is  maintained  between  fluids  and  min- 
erals during  incremental  loss  may  be  dic- 
tated by  the  relative  intracrystalline  diffu- 
sion rates  of  the  elements  in  their  respective 
mineral  reservoirs  or  by  dissolution/ 
reprecipitation  rates  of  the  host  minerals 
during  fluid  loss.  Boron  concentration, 
like  N  concentration  and  isotope  composi- 
tion, may  have  been  controlled  by  the  bulk 
fluid  loss  and  fluid-rock  exchange  history 
of  the  rocks.  If  so,  B  partitioned  progres- 
sively from  B-rich  minerals  (primarily  the 
white  micas)  into  H20-rich  fluids  derived 
largely  by  chlorite  breakdown.  Thus,  as 
with  N,  the  B  loss  need  not  be  attributed  to 
breakdown  of  host  minerals  such  as  white 
mica  and  biotite,  which  show  no  obvious 
decrease  in  their  combined  modal  abun- 


dance with  increasing  metamorphic  grade 
[see  equations  ( 1  -4)] .  Continuous  exchange 
reactions  like  those  which  stabilized  micas 
during  progressive  metamorphism  may  also 
allow  the  continual  reequilibration  of  fluid 
and  rock  stable  isotope  and  trace  element 
composition  during  progressive 
devolatilization  and/or  infiltration  by  ex- 
ternally-derived fluids. 


References 


Bebout,  G.  E.,  Geological  and  geochemical  inves- 
tigations of  fluid  flow  and  mass  transfer  during 
subduction-zone  metamorphism,  Ph.  D.  Dis- 
sertation, University  of  California,  Los  Ange- 
les, 1989. 

Bebout,  G.  E.,  Field-based  evidence  for 
devolatilization  in  subduction  zones:  Implica- 
tions for  arc  magmatism,  Science,  251,  413- 
416,  1991a. 

Bebout,  G.  E.,  Geometry  and  mechanisms  of  fluid 
flow  at  15  to  45  kilometer  depths  in  an  early 
Cretaceous  accretionary  complex,  Geophys. 
Res.  Lett.,  18,  923-926,  1991b. 

Bebout,  G.  E.,  and  M.  D.  Barton,  Fluid  flow  and 
metasomatism  in  a  subduction  zone  hydrother- 
mal  system:  Catalina  Schist  terrane,  Califor- 
nia, Geology,  17,  876-980,  1989. 

Bottinga,  Y.,  Calculated  fractionation  factors  for 
carbon  and  hydrogen  isotope  exchange  in  the 
system  calcite-C02-graphite-methane-hydro- 
gen  and  water  vapor.  Geochim.  Cosmochim. 
Acta  33,  49-64,  1969. 

Bottrell,  S.  H.,  L.  P.  Carr,  and  J.  Dubessy,  A 
nitrogen-rich  metamorphic  fluid  and  coexist- 
ing minerals  in  slates  from  North  Wales,  Min- 
eral. Mag.,  52,  451-457,  1988. 

Domanik,  K.,  R.  L.  Hervig,  and  S.  M.  Peacock, 
Beryllium  and  boron  contents  of  subduction 
zone  minerals:  an  ion  microprobe  study,  EOS, 
72,293-294,1991. 

Duit,  W.,  Jansen,  J.  B.  H.,  Van  Breemen,  A.,  and 
A.  Bos,  Ammonium  micas  in  metamorphic 
rocks  as  exemplified  by  Dome  de  L'Agout 
(France).  Amer.  J.  Sci.,  286,  702-732,  1986. 

Etheridge,  M.  A.,  Wall,  V.  J.,  Cox,  S.  R,  and  R.  H. 
Vernon,  High  fluid  pressures  during  regional 
metamorphism  and  deformation:  Implications 
for  mass  transport  and  deformation  mech- 
anisms,/ Geophys.Res.,89, 4344-4358, 1984. 


GEOPHYSICAL  LABORATORY 


41 


Ferry,  J.  M.,  Petrology  of  graphitic  sulfide-rich 
schists  from  south-central  Maine:  An  example 
of  desulfidation  during  prograde  regional  meta- 
morphism.  Amer.  Mineral.,  66, 908-930, 198 1 . 

Ferry,  J.  M.,  andL.  Baumgartner,  Thermodynamic 
models  of  molecular  fluids  at  the  elevated  pres- 
sures and  temperatures  of  crustal  metamor- 
phism,  Rev.  Mineral.,  17,  323-365,  1987. 

Haendel,  D.,  K.  Muhle,  H.-M.  Nitzsche,  G.  Stiehl, 
and  U.  Wand,  Isotopic  variations  of  the  fixed 
nitrogen  in  metamorphic  rocks,  Geochim. 
Cosmochim.  Acta,  50,  749-758,  1986. 

Hanschmann,  G.,  Berechnung  von  Isotop  Effekten 
auf  quantenchemischer  Grundlage  am  Beispiel 
stickstoffhaltiger  Molekule,  Zfl-Mitt.,  41,  19- 
39,  1981. 

Heier,  K.  S.,  Geochemistry  of  granulite  facies 
rocks  and  problems  of  their  origin,  Phil.  Trans. 
R.  Soc.  Lond.  A.,  273,  429-442,  1973. 

Honma,  H.,  and  Y.  Itihara,  Distribution  of  ammo- 
nium in  minerals  of  metamorphic  and  granitic 
rocks,  Geochim.  Cosmochim.  Acta,  45,  983- 
988,  1981. 

Kreulen,  R.,  A.  Van  Breeman,  and  W.  Duit,  Nitro- 
gen and  carbon  isotopes  in  metamorphic  fluids 
from  the  Dome  de  L'Agout,  France,  Proceed- 
ings of  the  Fifth  International  Conference  on 
Geochronology,  Cosmochronology,  and  Iso- 
tope Geology,  p.  191,  1982. 

Leeman,  W.  P.,  A.  E.  Moran,  and  V.  B.  Sisson, 
Compositional  variations  accompanying  meta^ 
morphism  of  subducted  oceanic  lithosphere: 
Implications  for  genesis  of  arc  magmas  and 
mantle  replenishment,  Abstr.,  Seventh  ICOG 
Mtg.,  58,  1990. 

Patience,  R.  L.,  C  J.  Clayton,  A.  T.  Kearsley,  S.  J. 
Rowland,  A.  N.  Bishop,  A.  W.  G.  Rees,  K.  B. 
Bibby,  and  A.  C.  Hopper,  An  integrated  bio- 
chemical, geochemical,  and  sedimentological 
study  of  organic  diagenesis  in  sediments  from 
Leg  1 12,  in  Proceedings  of  the  Ocean  Drilling 
Program,  Scientific  Results,  112,  E.  Suess  et 
al.,  eds.,  Ocean  Drilling  Program,  College  Sta- 
tion, Texas,  135-153,  1990. 

Piatt,  J.  P.,  Metamorphic  and  deformational  pro- 
cesses in  the  Franciscan  Complex,  California: 
some  insights  from  the  Catalina  Schist  terrain, 
Geol.  Soc.  Amer.  Bull,  86,  1337-1347,  1975. 

Rumble  D.,  Stable  isotope  fractionation  during 
metamorphic  volatilization.  Rev.  Mineral.  10, 
327-353,  1982. 

Sorensen,  S.  S.,  Petrologic  and  geochemical  com- 
parison of  the  blueschist  and  greenschist  units 
of  the  Catalina  Schist  terrane,  southern  Califor- 
nia, Geol.  Soc.  Amer.  Mem.,  164,  59-75,  1986. 

Sorensen,  S.  S.,  and  M.  D.  Barton,  Metasomatism 
and  partial  melting  in  a  subduction  complex: 
Catalina  Schist,  southern  California,  Geology, 
15,  115-118,  1987. 


Valley,  J.  W.,  Stable  isotope  geochemistry  of  meta- 
morphic rocks,  Rev.  Mineral.  16,  445-489, 
1986. 

Walther,  J.  V.,  and  P.  M.  Orville,  Volatile  produc- 
tion and  transport  in  regional  metamorphism, 
Contrib.  Mineral.  Petrol,  79,  252-257,  1982. 


The  fa  Content  of  Normative  ol 
Felix  Chayes 

Replacement  of  conventional  "wet- 
way"  analytical  procedures  by  instrumented 
physical  techniques  is  generating  a  large 
and  rapidly  expanding  reservoir  of  rock 
analyses  in  which  Fe  is  not  partitioned  by 
oxidation  state.  (Of  the  14,722  analyses  of 
igneous  rocks  included  in  the  current  ver- 
sion of  the  base  IGBADAT,  for  instance, 
almost  a  quarter  lack  Fe  partition.)  Com- 
parison of  analyses  subject  to  this  defi- 
ciency with  older  data,  indeed,  plotting 
them  in  many  common  petrographic  varia- 
tion diagrams,  requires  some  external 
nonanalytical  assignment  of  Fe  oxidation 
state. 

In  the  several  procedures  now  in  use 
(for  a  thorough  review  see  Middlemost, 
1989),  this  adjustment  consists  of  the  ap- 
plication of  a  simple  formula  to  convert  the 
lone  Fe  or  Fe-oxide  value  into  two.  With 
few  exceptions  the  conversion  is  made 
without  regard  to  the  amounts  of  other 
components  reported  in  the  bulk  analysis. 
And  however  important  petrologic  and 
mineralogical  factors  may  have  been  in  the 
development  of  these  relations,  they  play 
no  direct  role  in  their  application.  In  prac- 
tice, the  preferred  rule  is  simply  applied  by 
rote  to  any  analysis  that  is  to  be  used  in  a 
fashion  requiring  some  estimate  of  the  oxi- 


42 


CARNEGIE  INSTITUTION 


dation  state  of  Fe  in  the  specimen  in  ques- 
tion. 

In  olivine1  bearing  rocks  this  may  lead 
to  ol  unrealistically  rich  in  fa.  Fortunately, 
loss  of  information  about  the  oxidation 
state  of  Fe  in  bulk  analyses  of  igneous 
rocks  has  been  accompanied  by  a  manifold 
increase  in  the  amount  of  information  about 
the  chemical  compositions  of  their  con- 
stituent minerals.  It  has  been  pointed  out 
(Chayes,  Annual  Report  1989-1990,  p.  40- 
42)  that  whenever  such  information  is  in 
fact  available  for  an  analyzed  specimen,  it 
would  be  a  simple  matter  to  adjust  the  Fe 
oxidation  ratio  inversely,  by  using  the  fa 
content  of  modal  olivine,  or  some  multiple 
of  it,  as  an  upper  limit  on  the  normative 
ratio  fa/ ol.  What  does  one  do,  however,  if, 
as  is  still  true  of  many  rocks  and  will  always 
be  true  of  some,  no  pertinent  mineralogical 
information  is  available? 

Rather  than  fall  back  on  a  single  rule  for 
all  rocks,  one  might  then  prefer  an  inverse 
adjustment  based  on  the  average/a/0/  con- 
tent found  from  more  complete  analyses  of 
rocks  similar  to  that  in  question.  This  note 
presents  relevant  summaries  drawn  from 
IGBADAT.  Of  the  14,722  specimen  de- 
scriptions in  the  current  version  of  that 
base,  11,294  are  accompanied  by  "com- 
plete" bulk  analyses,  which  sum  to  be- 
tween 95  and  105%  and  include  determina- 
tions for  both  oxides  of  Fe.  The  cumulative 
frequency  distribution  offal  ol  in  the  4,589 
of  these  that  are  6>/-normative  is  shown  by 
the  upper  line  in  Fig.  19.  The  lower  line 
shows  the  same  information  for  the  3,476 

1  In  this  note,  "olivine"  denotes  the  minerall,  ol  and 
fa  the  normative  components  computed  from  the 

analysis 


0.3   0.4  0.5   0.6 

Lower  Class  Mark 


Fig.  19.  Cumulative  frequency  distributions  of fa/ 
ol  in  analyses  with  summations  in  the  range  95- 
105  and  containing  analytical  determinations  for 
both  oxides  of  Fe.  Data  from  IGB  ADAT4  (4,589 
analyses  in  all,  3,476  with  H2O  <2%,  class  width 
=  0.05). 


which  also  contain  less  than  2%  of  H2O. 
Values  of  fa/ ol  in  excess  of  0.5  comprise  no 
more  than  2.5%  of  the  entries  in  either  data 
set.  In  fact,  in  only  6%  of  all  values 
included  in  either  summary  is  fa/ol  greater 
than  0.4,  and  in  only  12%  is  it  greater  than 
0.35. 

The  mean  and  standard  error  of  this 
statistic  in  each  of  a  number  or  rock  types 
(really  name  groups)  are  shown  in  Table 
10.  Despite  the  limited  overall  range  of fa/ 
ol  and  its  rather  broad  within  group  disper- 
sion, there  appear  to  be  marked  differences 
between  group  means;  in  randomly  drawn 
samples  several  of  these  differences  would 
be  considered  statistically  significant.  In 
fact,  with  the  exception  of  the  tephrites  and 
gabbros  the  groups  seem  to  fall  into  three 
sets,  with  mean  values  of  fa/ol  in  the 
ranges,  respectively,  0.28-0.31,  0.20-0.21 
and  0.11-0.13.  In  Table  1 0  the  members  of 
each  set  are  listed  in  order  of  decreasing 
size  of  the  sample  available  for  calculation 
of  the  mean  and  its  error. 

As  of  this  writing,  calculations  have 
been  completed  only  for  complete  groups; 
from  prior  experience  it  is  anticipated  that 


GEOPHYSICAL  LABORATORY 

Table  10.  Average  and  standard  error  of  fa/ ol,  by  rock  type,  data  from  IGBADAT4 


43 


Rock  Type 

Number  of  Analyses 

falol 

All 

ol  >0.5% 

mean 

std.  error 

Tholeiite 

339 

202 

0.306 

0.005 

Dolerite 

359 

140 

0.285 

0.009 

Diorite 

358 

78 

0.310 

0.013 

Diabase 

200 

62 

0.298 

0.015 

Andesite 

794 

41 

0.281 

0.021 

Basalt 

2,532 

1,629 

0.206 

0.003 

Gabbro 

504 

342 

0.181 

0.007 

Basanite,  hawaiite,  mugearite 

384 

320 

0.197 

0.007 

Trachyandesite,  benmoreiite,  tristainite 

246 

158 

0.208 

0.012 

Phonolite 

319 

116 

0.208 

0.019 

Mafic  plutonics,  ultramafic  nodules 

463 

384 

0.114 

0.005 

Ultramafic  volcanic s  &  dikes 

313 

224 

0.128 

0.008 

Tephrite 

101 

94 

0.156 

0.011 

elimination  of  hydrated  materials  will  shift 
means  upward  slightly,  and  since  in  all 
these  groups  high  values  of  falol  are  either 
very  rare  or  lacking,  this  may  well  reduce 
observed  differences  between  means.  It 
should  also  materially  reduce  within  group 
dispersion,  however,  so  that  its  effect  on  the 
significance  of  intergroup  differences  re- 
mains to  be  seen. 

The  underlying  situation,  alas,  is  prob- 
ably not  as  simple  as  Table  10  may  make  it 
seem.  Even  ignoring  the  evident  unruli- 
ness  of  the  tephrites  and  gabbros,  elaborate 
and  sometimes  rather  elliptical  rationaliza- 


tions about  qualifications  for  membership 
in  some  of  the  groups  cannot  be  altogether 
avoided  in  any  organization  of  data  based 
on  the  actual  usage  of  rock  names  rather 
than  on  their  a  priori  definitions.  A  more 
thorough  discussion  of  the  problems  this 
raises  is  in  preparation. 


References 


Middlemost,  E.A.K.,  Iron  oxidation,  norms,  and 
the  classification  of  igneous  rocks,  Chem. 
GeoL,  77,  19-26,  1989. 


GEOPHYSICAL  LABORATORY 


45 


Igneous  and  Metamorphic  Petrology 
Experimental  studies 


Raman  Spectra  of  High-Temperature 

Silicate  Melts:  NA2O-S1O2,  K2O-S1O2, 

and  L12O-S1O2  Binary  Compositions 

John  D.  Frantz  and  Bjorn  O.  My  sen 

The  structure  of  silicate  liquids,  deter- 
mined at  high  temperature,  and  relation- 
ships between  structure  and  properties  are 
centrally  important  to  our  understanding  of 
natural  magmatic  processes.  Principally 
from  studies  of  quenched  melts,  for  the 
compositional  range  of  most  natural  mag- 
matic liquids,  a  simple  equilibrium  of  the 
form; 

T205(2Q3)  *=>  T03(Q2)  +  TOiiQ4),1    (1) 

describes  the  principal  elements  of  the  struc- 
ture (e.g.,  Virgo  et  ai,  1980;  Matson  et  al., 
1983;  Stebbins,  1988),  where  T represents 
tetrahedrally  coordinated  cations.  From 
spectroscopic  data  on  quenched  binary 
metal  oxide-silica  melts  (glasses),  it  ap- 
pears that  equation  (1)  shifts  to  the  right 
with  increasing  Z/r2  of  the  network-modi- 
fying metal  cation. 


1  In  this  paper,  stoichiometric  units  are  used  to 
describe  the  structural  units.  T  represents  tetrahe- 
drally coordinated  cation(s)  such  as,  for  example, 
Si4+  and  Al-K  In  view  of  the  frequent  usage  of  Q- 
notations  in  the  NMR  literature,  the  equivalent 
notations  are  shown  here  for  convenience.  The 
superscript  in  the  Q-notation  refers  to  the  number 
of  bridging  oxygens  on  the  unit. 


Whether  this  relationship  holds  true  in 
the  molten  state  is  not  known.  Studies 
using  NMR  (Stebbins,  1988;  Brandriss  and 
Stebbins,  1988)  and  Raman  spectroscopy 
(Seifertertf/.,  1981;  My  sen,  1990;  Cooney 
and  Sharma,  1990)  indicate  systematic 
structural  changes  with  temperature.  There- 
fore, even  though  the  spectra  of  glasses  and 
melts  qualitatively  resemble  one  another, 
at  least  for  the  limited  compositions  stud- 
ied, structural  data  from  glasses  may  not  be 
used  to  characterize  relationships  between 
melt  properties  and  melt  structure.  In-situ, 
high-temperature  spectra  of  melts  are  re- 
quired. Measurements  of  the  Raman  spec- 
tra of  melts  in  their  molten  state  to  tempera- 
tures of  above  1600°C  are  now  possible.  In 
the  present  study,  the  effect  of  temperature 
on  the  spectra  of  the  Na20-Si02,  K2O- 
Si02,  and  Li02-Si02  are  investigated. 

The  integration  of  a  micro-heating  stage 
with  the  focusing  capability  of  the  micro- 
Raman  system  is  fundamental  to  the  suc- 
cess of  the  proposed  research.  The  heater/ 
thermocouple  is  fabricated  from  pieces  of 
0.8-mm  Pt  and  Pt9oRhiO  wire  which  are 
welded  and  flattened  to  500  jum  thickness 
at  the  join.  A  1-mm  diameter  hole  is  drilled 
through  the  Pt-PtooRh  10  junction  (Fig.  20). 
In  the  heating  stage,  the  Pt-PtooRhio  ther- 
mocouple serves  a  dual  function  as  both 
thermocouple  and  heater  (Ohashi  and 
Hadidiacos,  Year  Book  75,  828-834).  The 
heater  responds  to  applied  power  in  a  mat- 
ter of  seconds.  The  temperature  increments 


46 


CARNEGIE  INSTITUTION 


Fig.  20.  Microphotograph  of  heater  with  glass 
sample  in  place  as  indicated. 


reported  here  were  accomplished  within 
10-30  seconds.  After  a  sample  is  melted 
into  the  hole  it  is  held  in  place  by  surface 
tension  during  high  temperature  experi- 
ments. Sample  thickness  near  the  edge  is 
about  500  Jim,  and  thicker  in  the  center. 
Although  the  emf  from  the  thermocouple 
yields  an  approximate  temperature  (uncer- 
tainties are  introduced  by  electrical  con- 
nections between  dissimilar  metals  and  al- 
loys), accurate  temperature  calibration  is 
achieved  by  suspending  a  0.04  mm  Pt- 
Pt9oRhio  thermocouple  into  the  melt.  The 
temperatures  measured  with  this  design  are 
accurate  to  within  4 °C  anywhere  within  the 
sample.  The  microheater  was  then  placed 
on  the  microscope  stage  of  a  custom  de- 
signed microscope  port  of  a  Dilor  XY  con- 
focal  micro-Raman  system  equipped  with 
an  EG&G  Model  1433-C  cryogenic  CCD 
detector.  Small  chips  (~  1  mg)  of  the  glass 
were  placed  over  the  hole  in  the  microheater 
(Fig.  20)  and  melted.  The  samples  were 
excited  with  the  488-nm  line  of  a  Spectra 
Physics  model  2025  Ar+  ion  laser,  operat- 


ing near  850  mW  at  the  sample. 

The  compositions2  chosen  for  high-tem- 
perature spectroscopy  were  on  binary  M2O- 
Si02  joins,  with  M  =  Li,  Na,  andK.  These 
three  different  metal  cations  were  chosen 
so  as  to  evaluate  the  influence  of  electronic 
properties  of  the  alkali  metal  on  the  tem- 
perature-dependence of  their  structure. 
Relationships  between  bulk  melt  polymer- 
ization (NBO/Si)  and  the  temperature  de- 
pendence of  the  structure  were  addressed 
by  changing  the  Si/O  in  each  binary  sys- 
tem. 

The  widest  range  of  metal  oxide/silicon 
compositions  were  studied  in  the  system 
Na20-Si02  (Fig.  21)  as  the  problem  of 
glass  devitrification  appeared  to  be  mini- 
mal in  this  system.  The  high-frequency 
region  (800-1300  cnr*)  of  the  NS7,  NS5, 
NS3,  and  NS2  glass  spectra  (marked  25  in 
Fig.  21a,  b,  c,  d)  are  similar  to  those  re- 
ported by  Mysen  et  al.  (1982).  There  is  an 
approximately  1 00  cm-  1  wide  (full  width  at 
half  height)  band  centered  near  1 100  cm-1. 
A  high-frequency  shoulder  near  1150 
cm-  *  can  be  discerned  in  the  spectra  of  NS7 
and  NS5  glasses.  All  spectra  exhibit  a  dis- 
tinct band  near  950  cm-1,  the  intensity  of 
which  visually  increases  systematically 
with  increasing  Na/Si  (increasing  NBO/Si 
of  the  melt).  The  glass  spectra  have  been 
deconvoluted  previously  (Mysen  et  al., 


2  These  compositions  are  designated  by  the  molar 
ratio  of  M2O  to  Si02,  so  that  NS7,  NS5,  NS3,  and 
NS2  are  compositions  Na207Si02,  Na20»5Si02, 
Na20*3Si02,  and  Na20«2Si02,  and  KS5,  KS4, 
LS3,  and  LS2  are  K20»5Si02,  K20»4Si02, 
Li2O3Si02,  Li20»2Si02,  respectively. 


GEOPHYSICAL  LABORATORY 


47 


6^ 
CO 

c 

CD 


200 


150 


100  V 


.'4 

r*    •:  #  -  v.-  .;% 

fitist  < 


NS7 

T|iq:1367°C 
Tn  :820°C 


200 


900   1000   1100   1200 

Wavenumber,  cm" 


CO 

c 

CD 


B 


150 


100 


50 


*t 


.***  V 
WAS 


NS5 


T 


liq 


1255°C 
945°C 


% 


^■N.    ^  ^y  v*v .*>.>.% vs    «*.      ■-<  y     ■.*     t 

•.  •x::-;..::;;::v     ..;**■  V-.      .-  *>.    • 


800    900   1000   1100   1200 

Wavenumber,  cm"1 


1300 


CO 

c 

CD 


200  - 


150  - 


NS3 

T„q:  805°C 

Tg:  445°C 


100 


800        900       1000      1100      1200 

Wavenumber,  cm" 


1300 


co 

c 

CD 


200  - 


150 


100 


D 


VX        ff£f*jJST 


NS2 

T,iq:  870'C 

Tg:  490 °C 


l\%\...  ......  w,..%v1 1 55 


900       1000      1100 

Wavenumber,  cm" 


Fig.  21.  In-situ,  high-temperature  Raman  spectra  of  glasses,  supercooled  melts,  and  melts  on  the  join 
Na20-Si02  as  a  function  of  temperature  (numbers  on  the  right  side  of  each  spectrum  represent 
temperature  in  °C).  The  liquidus,  Tug,  and  glass  transition,  To,  temperatures  for  each  composition  are 
indicated  on  the  figures.  The  intensities  are  calculated  relative  to  the  greatest  intensity  within  each 
spectrum.  A  -  composition  NS7  (Na20»7Si02,  bulk  melt  NBO/Si=0.28);  B  -  NS5  (Na20»5Si02,  bulk 
melt  NBO/Si=0.4);  C  -  NS3  (Na203Si02,  bulk  melt  NBO/Si=0.67);  D  -  NS2  (Na202Si02,  bulk  melt 
NBO/Si=l).  In  each  panel,  the  spectra  from  successively  higher  temperature  are  offset  by  10%  for 
clarity. 


48 


CARNEGIE  INSTITUTION 


100 


80 


60 

'55 

c 

£ 

S  40 


20 


800 


NS2 


& 


Na20-Si02  glass 
25°C 


[usa\  Jz 

;  >NS5  Nst. 
*  .?.,*«*{«'.  * 
*   ,<NS7* 

"ft L_ 


c 
c 


900 


1000 


1100 


1200 


1300 


Wavenumber,  cm" 


100 


80 


A  Na20-Si02  melts 
*\|  1144-1165°C 


•  w 

NS2            •  ; 

• 

60 

-                      #%                     .  * 

.•;       :* 

•  .i 

*.               »* 

•  • 

•                         if 

■  • 

•  ■  • 

.  #• 

•  •• 

40 

#        :        *  -• 

9  \ 
•  .i 

r      :     V7 

•  .  t 

.-.    • 

•% 

«                 .„-. 

•   > 

*        NS3         i . 

•    •:•• 

20 

-      •    >S    /: 

*.  ** 

;          ;      rHE&*$  •' 

.;       f     :J*v^-  • 

t  * 

..      '     f  JT   •    •  .' 

\  'i 

0 

•fr'     ?*\    NS7   / 

vs. 

800 


900 


1000 


1100 


1200 


1300 


Wavenumber,  cm" 


Fig.  22.  Comparison  of  room  temperature  Raman  spectra  of  glasses  on  the  join  Na20-Si02  (A)  with 
spectra  of  the  same  materials  at  temperatures  between  1144°C  and  1165°C  (B)  for  the  sample 
compositions  as  shown  in  Fig.  21.  The  temperatures  for  all  but  the  NS7  samples  are  above  the  liquidus 
temperature.  For  the  NS7  sample,  the  temperature  of  spectra  acquisition  is  above  the  glass  transition 
temperature  (see  Fig.  21  for  glass  and  liquidus  temperatures  of  these  samples).  The  intensities  are 
calculated  relative  to  the  greatest  intensity  within  each  spectrum. 


1982)  and  the  950,  1100  and  1150  cm-1 
bands  have  been  assigned  to  Si-0  stretch 
vibrations  in  specific  structural  units  (see 
Virgo  et  al,  1980;  Furukawa  et  ai,  1981; 
McMillan,  1984,  for  discussion  of  band 
assignments).  The  sharp  band  near  950 
cm-1  is  assigned  to  Si-O"  stretching  in 
structural  units  with  2  nonbridging  oxygens 
per  silicon  (Q2).  The  main  band  centered 
near  1 100  cm-1  is  due  to  Si-O"  stretching  in 
units  with  NBO/Si  =  1  (1100  cm1  band; 

3  For  convenience,  in  this  paper,  we  will  refer  to 
the  950, 1 100,  and  1 150  cm"1  bands  as  the  Q2,  iP 
and  Q^  bands 


Q3).  The  high  frequency  shoulder  on  the 
1 100  cm-  *  band  at  approximately  1 1 50  cnr 
1  results  from  the  presence  of  fully  poly- 
merized units  (Q4).  Thus,  all  the  glass  spec- 
tra from  the  Na20-Si02  system  are  consis- 
tent with  the  existence  of  structural  units 
with  their  individual  NBO/Si  values  =  2 
(Q2  or  Si032"  units),  1  (Q3  or  Si20s2" 
units)  and  0  (Q4  or  Si02  units)3.  There  is  no 
evidence  for  structural  units  less  polymer- 
ized than  NBO/Si  =  2  in  these  spectra,  in 
accord  with  previous  Raman  and  NMR 
data  (e.g.,  Virgo  et  al.  1980;  My  sen  et  al, 
1982;  Stebbins,  1987).  Equation  (1)  can  be 


49 


100  i_ 


80 


60 


CO 

c: 


40 


20 


:t» 


0  # 


NS5 


NS5  &  KS5 
25°C 


%• 


/ 


'55 

c 

"E 


1100 


1200 


1300 


Wavenumber,  cm" 


100 


80 


60  - 


B 


« #  „  • 

NS5  &  KS5 
1 370-1 380°C 

•        • 

• 

T^5:1255°C 
Tf 5:  745°C 

1; 

T^5:  940°C 

. 

TgS5:  540°C 

• 

"»• 

• 

»  • 

6      * 

^t 

40 

v      # 

'"  • 

S 

,  • 

^       * 

"• 

* 

"  » 

.3=         * 

\\ 

r.             • 

o  • 

20 

NS5     »w      ; 

»J 

*4*.Vv      .# 

*_v 

"e        KS5    * 

o  tft&ffW' 

i     ■.     •*  i                   i 

\ 

v&f20»K 

800     900     1000    1100    1200 

Wavenumber,  cm' 


1300 


Fig.  23.  Comparison  of  room  temperature  (25°C  -  A)  and  high  temperature  (1370-1380°C  -  B)  spectra 
of  KS5  (K205SiC>2)  and  NS5  (Na20»5Si02)  glasses  (A)  and  melts  (B).  The  bulk  melt  NBO/Si  for  both 
samples  is  0.4.  The  intensities  are  calculated  relative  to  the  greatest  intensity  within  each  spectrum. 


used,  therefore,  to  describe  the  anionic 
equilibria  for  these  compositions. 

Increasing  temperature  has  three  effects 
on  the  Raman  spectra.  (1)  The  band  near 
950  cm-1  (Q2  band)  relative  to  the  band 
near  1100  cm-l  increases  in  intensity  as  a 
function  of  increasing  temperature  (Fig. 
21 A  through  2 ID).  An  increase  of  about 
60%  from  the  25° C  spectrum  to  the  high- 
est-temperature spectrum  was  determined 
for  all  four  compositions  (Fig.  22).  A  simi- 
lar increase  in  intensity  with  temperature 
was  also  noted  by  Seifert  tt  al.  ( 1 98 1 )  in  the 
high-temperature  spectra  of  (Na2Si205)85- 
(Na2(NaAl)205)i5  melt.  (2)  The  intense 
envelope  centered  near  1100  cm-1  broad- 


ens with  temperature  (from  slightly  less 
than  100  cm-1  in  spectra  of  room  tempera- 
ture glass  to  more  than  1 20- 1 30  cm- 1  in  the 
spectra  of  melts  and  supercooled  liquids. 
The  -1150  cm-1  shoulder  evident  in  the 
NS5  and  NS7  room  temperature  spectra  is 
not  so  clear.  (3)  The  frequency  of  the  -950 
cm-1  band  as  well  as  the  envelope  centered 
near  1100  cm-1,  shifts  to  slightly  lower 
temperature. 

The  relationships  between  temperature 
and  electronic  properties  of  the  network- 
modifying  metal  cation  (K,  Na,  and  Li)  are 
illustrated  in  Figs.  23  and  25.  At  room 
temperature  (Fig.  23A),  the  spectrum  of 
NS5  glass  shows  a  more  distinctive  Q2 


50 


CARNEGIE  INSTITUTION 


1001- 


C 


NS3  &  LS3 
25°C 


100 


80 


60 


1      40 


1300 


B 


/"A  Nf 

.  *  •  1 1 

»  :  1  • 


NS3  &  LS3 
55-1165°C 


T^S3:  805°C 


LS3  / 


NS3 

g  • 

LS3 

iiq: 


»:T9 

1: 

\\ 

ft 


LS3 


445  °C 
1224°C 
725°C 


/  ns3  /  ;\ 

QEteJL  ■ i i ^^i 


Wavenumber,  cm 


800         900        1000       1100 

Wavenumber,  cm 


1200 

-1 


1300 


100  1- 


80  - 


■*— • 
CO 

c 

CD 


60  - 


40  - 


20; 


800 


pc 

i 

rV.    NS2  &  LS2 

9 
• 
r 
• 

:.:   25'C 

•  • 

- 

/ 

7 
• 
• 

:  »• 
•• 

- 

/  .' 

•• 

•• 
• 

,Als2,'    : 

•• 
•• 
•• 

i 

9 

%  *     1 

V 

—                                                 • 

v 

• 

« 

ft       < 

> 

• 

:* 

\      j 

\ 

•7 

*NS2/ 

• 
9 

1 

I    t     1 

1 

I                I                l 

100 


80 


60 


CO 

I  40 


20 


900   1000   1100 

Wavenumber,  cm 


1200 
-1 


1300 


D 


NS2  &  LS2 
1205-1 21 5°C 


LS2,"     J 

X  /  / 

*    1/    / 

?     'J 

:        Vns2 


* 


ST 


7 

\ 


NS2 
Iiq    • 

NS2. 

g    : 

LS2. 
Iiq    ■ 

LS2 


870°C 
490*C 
1033°C 
600°C 


y 


_  ^*^^j 


800        900       1000      1100 


Wavenumber,  cm 


1200 
-1 


1300 


Fig.  24.  Comparison  of  Raman  spectra  of  glasses  and  melts  in  the  systems  Li20-Si02  and 
Na20  for  trisilicate  (bulk  melt  NBO/Si  =  0.67)  (A  -  room  temperature  and  B  - 1 1 55- 1 1 65°C) 
and  and  disilicate  (bulk  melt  NBO/Si  =  1 .0)  compositions  (C  -  room  temperature,  D  - 1 205- 
1215°C).  Symbols:  NS3  -  Na20«3Si02,  LS3  -  Li20»3Si02,  NS2  -  Na20«2Si02,  LS2  - 
Li20»2Si02.  Relevant  liquidus  and  glass  transition  temperatures  are  indicated  on  the 
figures.  The  intensities  are  calculated  relative  to  the  greatest  intensity  within  each  spectrum. 


band  than  that  of  KS5  (in  both  glasses,  the 
bulk  melt  NBO/Si  =  0.4).  When  comparing 
the  spectra  of  LS3  and  NS3  (NBO/Si  = 
0.667)  and  LS2  and  NS2  (NBO/Si  =  0.14) 
(Figs.  24A  and  24C),  the  intensity  of  the  Q2 
band  is  again  more  pronounced  in  the  spec- 


tra of  glasses  with  the  smallest  metal  cation 
(Li).  This  increased  intensity  of  the  950 
cm-1  band  would  indicate  enhanced  abun- 
dance of  Si032_  (or  Q2)  structural  units. 
High -temperature  spectra  of  KS5  and 
KS4  melts  reveal  the  same  broadening  of 


GEOPHYSICAL  LABORATORY 


51 


250  i- 


200  - 


150 


m 


a) 

I     100 


50 


KS5 

T,iq:940°C 
Tn  :  536° C 


900         1000        1100 

Waven umber,  cm 


1200 

-1 


1300 


250  - 


200  - 


KS4 

Tliq:  772°C 
: 423°C 


o 


&   150  P 


05 


—     100 


800         900        1000       1100 

Wavenumber,  cm 


1200 

1 


1300 


Fig.  25.  Raman  spectra  of  KS5  (K20»5SiC>2)  glasses  and  melts  (A)  and  KS4  (K20«4SiC>2)  glasses  and 
melts  (B)  as  a  function  of  temperature  (°C)  as  indicated  on  the  right  side  of  each  spectrum.  The  intensities 
are  calculated  relative  to  the  greatest  intensity  within  each  spectrum,  and  spectra  at  successively  higher 
temperatures  are  offset  by  10  %  for  clarity. 


the  1100  cm-1  envelope  in  the  KS5  and 
NS5  spectra  (Fig.  25).  The  KS5  spectrum 
does  not  show  evidence  for  a  950  cm-1 
band  at  any  temperature  studied,  whereas 
that  of  NS5  does  with  its  intensity 
increasing  with  temperature.  In  spectra  of 
the  KS4  composition,  the  950  cm-1  band  is 
present  in  the  room  temperature  spectra  as 
a  very  weak  band  or  shoulder  in  the  glass 
spectrum,  and  shows  a  distinctive  intensity 
increase  with  increasing  temperature  (Fig. 
25B).  Thus,  it  would  appear  that  Si032_ 
structural  units  are  no  longer  discernible 
(within  the  sensitivity  of  the  spectroscopic 
technique)  in  glasses  and  melts  on  the  K2O- 
Si02  join  for  compositions  with  bulk  melt 
NBO/Si  of  0.4.  Increased  temperature  (at 
least  to  1 380°C)  does  not  change  this  obser- 
vation. In  contrast,  in  the  Na20-Si02  sys- 
tem, all  structural  units  appear  to  be  present 
at  all  temperatures  with  melts  at  least  as 


polymerized  as  NS7  (NBO/Si  =  0.28). 

The  comparison  of  the  glass  and  melt 
spectra  of  the  tri-  and  di-silicates  of  Na  and 
Li  (Fig.  24)  reveal  (1)  that  the  intensity  of 
the  Q2  band  in  LS3  glass  is  greater  by  a 
factor  of  about  3  compared  with  that  of 
NS3  (Fig.  24 A),  whereas  at  high  tempera- 
ture in  the  molten  range,  the  difference  has 
decreased  to  about  a  factor  of  2  (Fig.  24B). 
These  relative  intensity  changes  can  also 
been  seen  in  the  glass  and  melt  spectra  of 
LS2  and  NS2  composition  (Fig.  24C,D). 
(2)  In  the  1100  cm-1  envelope,  the  maxi- 
mum is  at  lower  frequency  in  the  spectra  of 
the  lithium  samples  than  in  the  sodium 
samples  and  there  might  be  a  slight  in- 
crease in  frequency  difference  as  the  the 
glasses  are  transformed  to  melts. 

Previous  deconvolutions  of  this  high- 
frequency  envelope  (e.g.,  My  sen  et  al., 
1982;  Mysen,  1990)  demonstrated  that  the 


52 


CARNEGIE  INSTITUTION 


1150  cm-1  band  (Si02  or  Q4)  on  the  high- 
frequency  limb  of  the  1100  cm-1  band 
always  increased  when  the  intensity  of  the 
Q2  band  increased.  These  increases  were 
accompanied  by  a  concomitant  decrease  in 
the  1100cm-1  (Si2052"  or  Q3)  band  inten- 
sity. Observations  such  as  these,  also  con- 
sistent with  interpretation  of  29Si  NMR 
spectra  (e.g.,  Stebbins,  1987),  lead  to  the 
conclusion  that  in  metal  oxide  silicate 
glasses  whose  anionic  equilibrium  can  be 
described  with  equation  ( 1 ),  increased  abun- 
dance of  Q2  (or  Si032_)  structural  units  is 
always  accompanied  with  an  increase  in 
Q4  (Si02)  and  a  decrease  in  Ql  (Si2052) 
units.  The  spectra  of  glass,  supercooled 
liquid,  and  liquid  for  Na20-Si02  and  K2O- 
Si02  show  visual  evidence  for  a  shift  of 
equation  (1)  to  the  right  with  increasing 
temperature.  In  the  absence  of  detailed 
statistical  deconvolution,  the  evidence  for 
the  glasses  and  melts  in  the  Li20-Si02 
glasses  and  melts  is  less  obvious.  It  would 
appear,  from  comparison  of  NS3  with  LS3 
and  of  NS2  with  LS2  glass  and  melt  spectra 
that  increasing  temperature  has  less  effect 
on  the  spectra  of  Li-silicates  than  on  those 
of  the  Na-silicates.  Visually,  the  intensity 
near  950  cm-1  in  the  Li-silicate  spectra 
does  not  change  significantly  with  tem- 
perature, whereas  those  of  NS3  and  NS2 
do.  The  intensity  difference  between  the 
two  sets  of  spectra  decreases,  therefore,  as 
the  glasses  are  heated  and  eventually  melted. 
In  summary,  high-quality  Raman 
spectra  of  silicate  melts  can  be  recorded  in- 
situ  at  magmatic  temperatures  and  above 
with  sample  acquisition  times  on  the  order 
of  one  minute  or  less.  From  such  spectra,  it 
has  been  found  that  in  binary  alkali  metal- 


silica  systems  in  the  bulk  melt  polymeriza- 
tion range  between  0.28  and  1.0,  glasses, 
supercooled  melts,  and  melts  in  the  tem- 
perature range  25-1475°C  generally  con- 
sist of  coexisting  Si032"  (g2),  Si2052" 
(Q3),  and  Si02  (Q4)  structural  units.  In 
potassium-bearing  systems,  the  upper  NBO/ 
Si  limit  for  Si032~  units  probably  is  at 
NBO/Si  between  0.4  and  0.5.  No  tempera- 
ture-dependence of  this  limit  was  observed. 
No  additional  units  were  identified  within 
this  temperature  range.  For  compositions 
with  the  Z/r2  of  the  alkali  metal  ranging 
from  2.8  to  0.6  (Li,  Na,  and  K),  equation  (1) 
shifts  to  the  right  with  increasing  tempera- 
ture. The  spectra  probably  indicate  a  com- 
positional dependence  of  the  free  energy 
for  reaction  (1).  Qualitatively,  the  effect  of 
temperature  on  equilibrium  (1)  decreases 
as  the  Zjr2  of  the  metal  cation  increases 
(Li>Na>K) ,  but  quantitative  evaluations  of 
its  values  have  not  been  been  carried  out. 


References 


Brandriss,  M.  E.,  and  J.  F.  Stebbins,  Effects  of 
temperature  on  the  structures  of  silicate  liq- 
uids: 29Si  NMR  results,  Geochim.  Cosmochim. 
Acta,  52,  2659-2669,  1988. 

Cooney,  T.  F.,  and  S.  K.Sharma,  High  temperature 
Raman  spectral  study  of  Ge02  andRb4SigOi8 
crystals,  glasses  and  melts,  EOS,  71,  1672, 
1990. 

Furukawa,  T.,  K.  E.  Fox,  and  W.  B.  White,  Raman 
spectroscopic  investigation  of  the  structure  of 
silicate  glasses.  HI.  Raman  intensities  and 
structural  units  in  sodium  silicate  glasses,  J. 
Chem.Phys.,  153,    3226-3237,1981. 

Matson,  D.  W.,  S.  K.  Sharma,and  J.  A.  Philpotts, 
The  structure  of  high-silica  alkali-silicate 
glasses  — A  Raman  spectroscopic  investiga- 
tion, J.  Non-Cryst.  Solids,  58,  323-352, 1983. 

McMillan,  P.,  A  Raman  spectroscopic  study  of 
glasses  in  the  system  CaO-MgO-Si02,  Amer. 
Mineral.,  69,  645-659,  1984. 


GEOPHYSICAL  LABORATORY 


53 


Mysen,  B.  O.,  The  role  of  aluminum  in  depoly- 
merized,  peralkaline  aluminosilicate  melts  in 
the  systems  Li20  -  AI2O3  -  Si02,  Na20- 
Al203-Si02  and  K20-Al203-Si02,  Amer. 
Mineral.,  75,  120-134,  1990. 

Mysen,  B.  O.,  D.  Virgo,  and  F.  A.  Seifert,  The 
structure  of  silicate  melts:  Implications  for 
chemical  and  physical  properties  of  natural 
magma,  Rev.  Geophys.,  20,  353-383,  1882. 

Seifert,  F.  A.,  B.  O.  Mysen,  and  D.  Virgo,  Struc- 
tural similarity  between  glasses  and  melts  rel- 
evant to  penological  processes,  Geochim. 
Cosmochim.  Acta,  45,  1879-1884,  1981. 

Stebbins,  J.  F.,  Effects  of  temperature  and  compo- 
sition on  silicate  glass  structure  and  dynamics: 
Si-29  NMR  results,  /.  Non-Cryst.  Solids,  106, 
359-369,  1988. 

Stebbins,  J.  F.,  Identification  of  multiple  structural 
species  in  silicate  glasses  by  29Si  NMR,  Na- 
ture, 330,  465-467,  1987. 

Virgo,  D.,  B.  O.  Mysen,and  I.  Kushiro,  Anionic 
constitution  of  silicate  melts  quenched  at  1  atm 
from  Raman  spectroscopy:  Implications  for 
the  structure  of  igneous  melts,  Science,  208, 
1371-1373,  1980, 


Peralkalinity  and  H2O  Solubility 
Mechanisms  in  Silicate  melts 

Bjorn  Mysen 

Relationships  between  properties  of 
magmatic  liquids  and  their  water  content 
have  been  examined  since  the  pioneering 
work  of  Bowen  (1928).  Despite  an  exten- 
sive literature  on  the  subject,  the  detailed 
nature  of  the  interaction  between  H2O  and 
the  silicate  melt  structure  remains  unclear, 
and  characterization  of  the  relations  be- 
tween the  structure  of  hydrous  silicate  melts 
and  their  physical  and  chemical  properties 
remains  a  topic  of  intense  interest. 

The  principal  anionic  equilibrium  that 
describes  the  melt  structure  in  magmatic 
liquids  under  anhydrous  conditions  is  (Virgo 
etal,  1980;  Mysen  et  al.,  1982;  Matsonef 
al,  1983;  Stebbins,  1987) 


T2O52-  <^>  TO32-  +  IO2. 


(1) 


Dissolved  H2O  can  interact  not  only  with 
TO2.  but  with  all  the  structural  units  in  the 
melts.  A  study  has  been  conducted,  there- 
fore, under  conditions  of  constant  degree 
of  bulk  melt  polymerization  (NBO/T  of 
anhydrous  melts  is  0.5)  with  Al/(A1+Si) 
and  water  content  as  the  compositional 
variables.  The  NBO/T  value  corresponds  to 
magma  compositions  intermediate  between 
tholeiite  and  andesite  (Mysen,  1988).  The 
range  in  Al/(A1+Si)  (0-0.3)  covers  that  found 
in  most  natural  magmatic  liquids. 

Starting  materials  were  mixtures  of 
Na2C03+Al203  +  Si02  on  thetetrasilicate 
(Na2Si409)-tetra-aluminate 
[Na2(NaAl)409]  join  (Fig.  26).  Exchange 


Fig.  26.  Composition  of  anhydrous  starting  mate- 
rials superimposed  on  simplified  liquidus  phase 
relations  in  the  system  Na20-Al203-Si02  (from 
Osborn  and  Muan,  1960). 


54 


CARNEGIE  INSTITUTION 


Anhydrous 


850  1075  1300 

Wavenumber,  cm"1 


-  100r 


7.5wt%H20 


830  1065  1300 

Wavenumber,  cm-1 


Fig.  27.  Curve-fitted  Raman  spectra  of  composi- 
tions indicated.  The  V950  and  vnoo  bands  are 
shaded  for  clarity. 


of  Al3+  for  Si4+  does  not  affect  the  bulk 
melt  polymerization  (NBOIT  =  0.5)  under 
anhydrous  conditions  because  both  Al3+ 
and  Si4+  occupy  tetrahedral  coordination 
(Mysen,  1990). 

The  glass  starting  materials  were  from 
the  same  batch  of  glasses  used  by  Mysen 
(1990).  About  20  mg  of  finely  crushed 
glass  together  with  distilled,  deionized  H2O 
was  placed  in  sealed  Pt  containers  for  high- 
pressure  synthesis.  All  water  contents  (<7.5 
wt  %)  were  less  than  that  needed  to  saturate 
the  melts  atthe  12kbarand  1400°Cusedfor 
sample  preparation.  These  samples  were 
subjected  to  12  kbar  at  1400°C  in  the  solid- 
media,  high-pressure  apparatus  (Boyd  and 
England,  1960)  for  90  min  and  tempera- 
ture-quenched at  a  quenching  rate  near 
100°C/s  between  the  experimental  tem- 
perature and  ~500°C.  The  quenching  rates 
were  similar  for  all  materials  studied.  The 
pressure  uncertainty  (as  calibrated  against 
the  quartz  <=>  coesite  and  albite  <=>  jadeite  + 
quartz  transitions)  is  near  ±1  kbar.  The 
effect  of  pressure  on  the  Pt-Pt9oRhio  ther- 
mocouples is  about  ±10°C  (Mao  etal.,  Year 
Book   70,  p.  281-287).  With  no  pressure 
correction  on  the  emf  output  from  this 
thermocouple,  the  temperature  is  consid- 
ered accurate  to  ±10°C. 

Structural  information  was  obtained 
from  analysis  of  Raman  spectra  of  the 
quenched  melts.  The  spectra  were  recorded 
with  an  automated  single-channel  Raman 
spectrometer  system  with  the  frequency- 
doubled  532-nm  line  of  an  NdYAG  laser 
operating  at  1  W  for  sample  excitation. 

The  abundance  of  the  structural  units 
was  determined  from  the  Raman  spectra, 
as  described  by  Mysen  {Annual  Report 


GEOPHYSICAL  LABORATORY 


55 


1988-1989,  p.  47-54).  From  deconvoluted, 
high-frequency  Raman  spectral  envelopes 
such  as  illustrated  with  Figure  27,  the  rela- 
tive intensities  of  the  V950  [(Si,  Al)-0"  stretch 
band  from  TO32-  structural  units]  and  vi  100 
bands  (Si,  Al)-0  stretch  band  from  T2O52- 
structural  units]  were  used  for  this  purpose. 
The  distribution  of  Al  among  those  units  as 
well  as  the  fully  polymerized  structural 
unit  (7T)2.;  V1150  and  V1200  bands)  was 
evaluated  from  frequency  shifts  of  the  indi- 
vidual bands  as  a  function  of  H2O  content 
and  bulk  melt  Al/(A1+Si). 

There  are  systematic  shifts  in  frequency 
of  important  Raman  bands  with  increasing 
water  content  and  with  increasing  bulk  Al/ 
(Al+Si)  (Fig.  28).  The  general  descent  of 
the  vi  150  and  V1200  frequencies  with  in- 
creasing H2O,  also  noted  for  hydrous 
NaA102-Si02  melts  (Mysen  and  Virgo, 
1986),  indicates  that  Al/(A1+Si)  of  the 
fully  polymerized  structural  units  (TO2.)  in 
the  melts  is  positively  correlated  with  H2O 
concentration.  It  is  notable,  however,  that 
even  in  anhydrous  samples,  the  V1150  and 
vi 200  frequencies  at  given  Al/( Al+Si)  are 
lower  than  those  observed  in  the  spectra  of 
fully  polymerized  (NBO/T  =  0)  NaA102- 
Si02  quenched  melts  with  the  same  bulk 
Al/(A1+Si)  (Fig.  28).  If  the  frequency  of 
these  bands  were  used  as  a  quantitative 
measure  of  Al/( Al+Si)  in  the  7X)2.  units, 
[Al/Al+Si)]7T)2.  in  the  anhydrous  sodium 
aluminosilicate  melts  would  be  0.32, 0.34, 
and  0.37,  for  NS4-A10  [bulk  melt  Al/ 
(Al+Si)  =  0.1],  NS4-A20  [bulk  melt  Al/ 
(Al+Si)  =  0.2],  and  NS4-NA30  [bulk  melt 
Al/( Al+Si)  =  0.30],  respectively.  The  er- 
rors in  these  numbers,  however,  are  quite 
large  due  to  the  shallow  slope  of  the  fre- 


1300 


§    1200 


900 


Anhydrous 


_» 1 i- 


0.0    0.1     0.2    0.3    0.4    0.5    0.6 
AI/(AI+Si) 


1300 


B 

>  - » 


o  1200CT-- 


tf^- - -O 


o 

§  1100, 
ex 

CD 

1000 


3  wt%  H2O 


900 


0.1         0.2        .0.3        0.4 
AI/(AI+Si) 


1300 


1000 


900 


5wt%H20 


0.0        0.1         0.2         0.3        0.4 
AI/(AI+Si) 


Fig.  28.  Frequencies  of  selected  (indicated  on 
Figure)  Raman  bands  as  a  function  of  bulk  melt 
Al/(A1+Si)  and  H20.The  bands  shown  are  as- 
signed to  (Si,Al)-0°  stretch  vibrations  in  fully 
polymerized  structural  units.  The  dashed  line  is 
from  the  system  NaA102-Si02-H20  (data  from 
Mysen  and  Virgo,  1986). 


56 


CARNEGIE  INSTITUTION 


1.0 
0.8 


A 


NS4;AI/(AI+Si)=0;  B     r  NS4-A10;  Al/(AI+Si)=0. 


6  0  2 

wt%  H20 


8 


Fig.  29.  Abundance  of  structural  units  (as  shown  on  figure)  as  a  function  of  H2O  content 
and  Al/(A1+Si).  Open  symbols  are  results  from  samples  with  D2O. 


quency  vs.  Al/(A1+Si)  curves.  The  results 
are,  nevertheless,  consistent  with  a  distinct 
preference  of  Al3+  for  the  most  polymer- 
ized among  the  coexisting  structural  units 
in  the  anhydrous  melts,  a  conclusion  con- 
sistent with  other  Raman  and  NMR  data 
(e.g.,  Mysen  etal,  1981;  Engelhardt  etal., 
1985;  Kirkpatrick  et  al.,  1986;  Oestrike  et 
al.,  1987;  Domine  and  Piriou,  1986; 
Merzbacher  et  al,  1990;  Mysen,  1990). 
Thus,  the  Al/(A1+Si)  of  the  TO2  structural 
units  would  exceed  that  of  the  bulk  melt. 
The  frequency  reduction  of  the  V1150 
and  vi 200  bands  with  increasing  H2O  con- 
tent leads  to  the  suggestion  that  the  [Al/ 
Al+Si)]7U2  in  hydrous  melts  is  enhanced 


further.  The  effect  of  water  on  these  Raman 
frequencies  is  less  pronounced,  however, 
in  the  peralkaline  aluminosilicate  melts 
studied  here  than  in  hydrous  NaA102-Si02 
melts  (Mysen  and  Virgo,  1986).  If  the  fre- 
quency trajectories  in  spectra  from  anhy- 
drous Si02-NaA102  quenched  melt  were 
employed  to  calibrate  the  Al/(A1+Si)  of  the 
fully  polymerized  units  in  hydrous  melts, 
the  minimum  frequency  (near  1080  cm-1) 
corresponds  to  Al/(A1+Si)  >  0.4  for  for 
7T)2.  units  hydrous  aluminosilicate  melts. 
The  influence  of  water  on  the  abun- 
dance of  structural  units  in  the  melts  (Fig. 
29)  can  be  inferred  from  the  intensity  varia- 
tions of  the  bands.  At  low  water  contents, 


GEOPHYSICAL  LABORATORY 


57 


the  AT2O5  in  hydrous  NS4  increases  (the 
melts  become  depolymerized)  with  water 
content.  This  hydroxy lation  mechanism 
may  be  expressed  with  the  equation 


2Si02  +  H20  »  SiO(OH)2, 


(2) 


which  when  considered  with  equation  (1) 
probably  results  in  an  increased  abundance 
of  S12O52-  accompanied  by  a  similar  de- 
crease in  Si032_.  If  this  were  the  only 
solubility  mechanism,  the  rate  of  depoly- 
merization  would  be  0.0133  per  mol  % 
dissolved  H2O. 

With  high  H2O  contents  there  is  evi- 
dence, however,  that  these  melts  undergo  a 
slight  polymerization  (Fig.  29  A).  Hydroxy  - 
lation  of  network-modifying  Na+  can  be 
described  by  the  system  of  equations 


2Na+  +  H20  +  Si032- 

<=>  2Na(OH)  +  Si02, 

2Na+  +  H20  +  Si2052" 

«>  2Na(OH)  +  2Si02, 

and 

2Na+  +  H20  +  2Si032" 

<=>  2Na(OH)  +  Si2052-, 


(3) 


(4) 


(5) 


in  which  the  network-modifying  cation 
(Na+)  reacts  with  water  to  form  Na(OH). 
Thus,  even  in  this  compositionally 
simple  Na20-Si02-H20  system,  the  spec- 
tral data  are  consistent  with  water  solubil- 
ity mechanisms  that  include  three  different 
types  of  OH-bonding  [H-OH  (molecular 
H2O),  Na-OH  (to  form  NaOH  complexes), 
and  Si-OH. 


The  abundance  trends  of  structural  units 
in  the  aluminous  samples  have  four  fea- 
tures in  common.  (1)  Whether  anhydrous 
or  water-bearing,  the  abundance  of  fully 
polymerized  structural  units  is  several  tens 
of  percent  higher  than  in  the  absence  of  Al 
(Figs.  29B-D).  (2)  The  abundance  of  TO2 
passes  through  a  maximum  with  water  con- 
tents between  1.5  wt  %,  and  3  wt  %.  (3) 
Both  depolymerized  ^Os2"  (NBO/T=  1) 
andTC>32  (NBO/T=2)  units  generally  coex- 
ist with  fully  polymerized  TO2  (NBO/T=0). 
(4)  There  is  a  minimum  in  abundance  of 
depolymerized  units  (T2O5  and  TO3)  at 
water  contents  corresponding  to  the  maxi- 
mum TO2  concentration. 

The  abundance  patterns  of  the  struc- 
tural units  in  Al-bearing  samples  indicate 
that  at  low  H2O  concentration  (<1.5  wt  % 
H2O)  solution  of  water  results  in  polymer- 
ization of  the  melts.  The  solubility  mecha- 
nism consistent  with  polymerization  is  in- 
teraction between  Na+  and  H2O  along  the 
lines  of  equation  (3). 

The  Al-bearing  melts  become  depoly- 
merized with  H2O  contents  >  1 .5-3  wt  %  as 
the  abundance  of  TO3  and  T2O5  units  is 
positively  correlated  with  H2O  concentra- 
tion, whereas  that  of  702.  is  negatively 
correlated  (Fig.  29B-D).  Depolymeriza- 
tion  via  interaction  of  H2O  with  expulsion 
of  tetrahedrally  coordinated  Al3+  (because 
of  reduction  in  Na+  charge -balance,  or 
Al(OH)3  formation,  or  both)  is  a  principal 
structural  mechanism  describing  this  be- 
havior. 

Although  Al3+  does  not  reside  exclu- 
sively in  fully  polymerized  anionic  units 
[the  frequencies  of  the  V95oand  vi  100  bands 
are  weakly  dependent  on  Al/(A1+Si),  in 


58 


CARNEGIE  INSTITUTION 


particular  in  the  low  water  concentration 
ranges],  available  information  (e.g, 
Engelhardt  et  al.,  1985;  Kirkpatrick  et  al., 
1986;  Oestrike  et  al,  1987;  Mysen,  1990) 
indicates  a  strong  preference  of  aluminum 
for  the  most  polymerized  of  the  coexisting 
units  at  least  in  anhydrous  melt  systems. 
We  will  discuss,  therefore,  the  solubility 
mechanism  on  the  basis  of  all  Al  in  such 
fully  polymerized  units  and  refer  to  them  as 
NaA102.  The  principles  derived  from  that 
discussion  will  not,  however,  be  affected 
by  some  Al3+  in  more  depolymerized  struc- 
tural units. 

The  depolymerization  reactions  of  wa- 
ter-bearing alkali  aluminosilicate  melts  rest 
on  the  premise  that  if  Al3+  in  4-fold  coordi- 
nation in  anhydrous  melts  forms  Al(OH)3 
complexes  after  hydrolysis  (with  Al  no 
longer  in  tetrahedral  coordination),  a  frac- 
tion of  the  charge -balancing  cation  (Na+  in 
the  present  system)  equal  to  the  proportion 
of  Al3+  in  these  complexes  becomes  net- 
work-modifying. This  mechanism  could 
be  described  with  the  following  set  of 
equations: 

2NaA102  +  2Si02  +  3H20 

<^>  2A1(0H)3  +  2Na+  +  Si2052",      (6) 

2NaA102  +  Si02  +  3H20 

^>  2A1(0H)3  +  2Na+  +  Si032",       (7) 

2NaA102  +  Si2052"  +  3H20 

<=>  2A1(0H)3  +  2Na+  +  2Si032".     (8) 

An  alternative  or  additional  mechanism 
would  operate  if  the  charge -balancing  cat- 
ion (Na+)  interacts  with  dissolved  H2O  to 
form  OH  complexes.  Then,  an  equivalent 


fraction  of  Al3+  originally  in  tetrahedral 
coordination  will  become  a  network-modi- 
fying cation. 

2NaA102  +  3Si02  +  H20 

<=>  2Na(OH)  +  2A13+  +  Si032',  (9) 

2NaA102  +  6Si02  +  H2O 

<=>  2Na(OH)  +  2A13+  +  3Si2052-,  (10) 

2NaA102  +  3Si2052"  +  H20 

<=>  2Na(OH)  +  2A13+ +  6Si032".    (11) 

Manning  et  al.  (1980)  and  Pichavant 
(1987)  inferred  from  liquidus  phase  equi- 
libria in  hydrous  quartz-feldspar  systems 
that  this  mechanism  would  explain  their 
observations.  Kohnera/.  (1989)  interpreted 
their  NMR  spectra  of  hydrous  NaAlSi30s 
glass  as  consistent  with  NaOH  complexing. 
In  the  case  of  Na-aluminosilicate  melts,  the 
latter  mechanism  (NaOH  complexing)  more 
efficiently  depolymerizes  the  melt,  be- 
cause for  each  charge-balancing  Na+  in 
anhydrous  melts  that  forms  the  NaOH  com- 
plex in  the  hydrous  environment,  three 
nonbridging  oxygens  can  be  stabilized  with 
Al3+.  For  each  tetrahedrally  coordinated 
Al3+  in  anhydrous  melts,  transformed  to 
Al(OH)3  in  hydrous  melts  only  one 
nonbridging  oxygen  can  be  stabilized  with 
the  Na+  released  in  this  process. 

These  expressions  are  consistent  with 
the  observations  in  Figs.  29B-D,  where  the 
abundance  of  both  T2O52-  and  TO32 
increasesand  that  of  TO2  decreases  as  the 
water  concentration  increases  above  1.5 
wt%.  Equation  (5)  (depolymerization)  to- 
gether with  equation  (3)  (polymerization) 
may  be  an  adequate  approximation  to  the 


GEOPHYSICAL  LABORATORY 


59 


principal  solution  mechanism  of  H2O  in 
highly  aluminous,  melts  such  as  NS4-A30 
[Al/(A1+Si)=0.3]. 

Liquidus  phase  relations  in  the  system 
Na20-Al203-Si02  in  the  compositional 
region  near  the  Na2Si409-Na2(NaAl)409 
join  can  be  employed  to  illuminate  the 
relations  between  H2O  activity  and  liquidus 
phase  relations.  Due  to  the  lack  of  experi- 
mental data  on  necessary  high-pressure 
liquidus  phase  relations,  the  1  -bar  informa- 
tion (Osborn  and  Muan,  1960)  will  be  used 
for  the  purpose.  It  is  recognized  that  by 
using  the  1-bar  data,  significant  errors  are 
introduced,  as  it  is  well  established  that 
pressure  by  itself  affects  both  liquidus  tem- 
peratures and  liquidus  volumes  in  this  (and 
other)  systems  (e.g.,  Boettcherera/.,  1984). 
Nevertheless,  it  is  informative  to  discuss 
the  consequences  of  dissolved  water  only. 
Isopleths  of  2.0  and  7.5  wt  %  were  recast 
in  terms  of  Na20,  AI2O3,  and  Si02.  The 


liquidus  temperatures  were  corrected  for 
water  content  by  assuming  ideal  mixing  of 
H2O  in  melts  on  the  basis  of  the  8 -oxygen 
model  of  Burnham  (1975)  and  by  using  the 
heat  of  fusion  data  for  liquidus  phases 
summarized  by  Richet  and  Bottinga  ( 1 986) 
with  the  expression 


7(K) 


1 


-1— fail 


v  melt\ 

AHoo/ 


R 


(12) 


a  o  fusion 


where  it  was  assumed  that  the  heat  of 
fusion  of  the  relevant  phases  was  tempera- 
ture independent. 

It  is  evident  that  with  small  amounts  of 
water  in  melt  solution  (2  wt  %  H2O),  the 
albite  liquidus  volume  is  greatly  expanded 
at  the  expense  of  both  quartz  on  its  low-Al 
side  and  nepheline  on  its  high- Al  side  (Fig. 
30).  The  small  liquidus  surface  of  crystal- 
line sodium  disilicate  (NS2;  Na2Si20s)  in 


1200 


1000 


O 

o 


£    800 

■*— » 
co 

g.   600 

E 
a> 

•"    400 


200 


■    $\ 

— 1 

^  Ab 

1 T  ■ 

""'Ab    y 
NS2 

— , 1 1              1 

1          _                     ^ 1 

^Ab 
1 

NS2 

• 

0.0 


0.1 


0.2 
AI/(AI+Si) 


0.3 


0.4 


Fig.  30.  Calculated  liquidus  surfaces  along  the  anhydrous  (NBO/T  =  0.5),  2  and 
7.5  wt  %  H20  isopleths,  as  a  function  of  Al/(A1+Si). 


60 


CARNEGIE  INSTITUTION 


the  anhydrous  system  has  disappeared.  Ex- 
pansion of  the  albite  liquidus  volume  at  low 
PH20  (-1.25  kbar)  along  the  join 
NaAlSi308-Na2Si205  has  also  been  docu- 
mented by  experimental  phase  equilibrium 
results  (Mustart,  1972).  The  liquidus  sur- 
faces of  fully  polymerized  albite  and  neph- 
eline  and  complete  lack  of  depolymerized 
crystalline  phases  reflect,  therefore,  the 
tendency  toward  polymerization  of  the 
peralkaline  aluminosilicate  melts  with  small 
amounts  of  water  in  solution. 

With  7.5  wt  %  H2O  in  solution  the 
liquidus  temperatures  are  greatly  depressed 
(Fig.  30).  There  is  no  quartz  liquidus  vol- 
ume, and  crystalline  NS2  (sodium  disilicate) 
is  an  important  phase  from  Al-free  compo- 
sitions to  Al/(A1+Si)  near  0.15,  where 
albite  appears.  The  albite  liquidus  volume 
ranges  from  Al/(A1+Si)  =  0.15  to  about 
0.22,  whereas  with  2  wt  %  H2O  the  volume 
ranges  from  near  0.05  to  about  0.26.  This 
very  significant  expansion  of  the  NS2 
liquidus  volume  at  the  expense  of  albite  is 
a  consequence  of  the  depolymerization  of 
the  melt  caused  by  the  dissolved  water.  The 
activity  of  the  nepheline  component  is  in- 
creased relative  to  that  of  albite  as  reflected 
in  the  encroachment  of  the  nepheline 
liquidus  volume  on  that  of  albite.  Even 
though  the  abundance  of  TO2  structural 
units  has  been  lowered,  most  likely  this 
expansion  of  nepheline  relative  to  albite 
results  from  an  enhancement  of  Al/(A1+Si) 
in  the  remaining  fully  polymerized  struc- 
tural units. 


References 

Boettcher,  A.L.,  Q,  Guo,  S.  Bohlen,  andB.  Hanson, 
Melting  of  feldspar-bearing  systems  to  high 
pressures  and  the  structures  of  aluminosilicate 
liquids,  Geolog,  12,  202-204,  1984. 
Bowen,  N.  L.  The  Evolution  of  the  Igneous  Rocks, 
332  pp.,  Princeton  Univ.  Press,  Princeton,  1 928 
Boyd,  F.  R.,  and  England  J.  L.,  Apparatus  for 
phase  equilibrium  measurements  at  pressures 
up  to  50  kilobars  and  temperatures  up  to 
1750°C,/. Geophys.Res.,  65, 741-748, 1960. 
Burnham,  C.  W\,  Thermodynamics  of  melting  in 
experimental   silicate-volatile  systems, 
Geochim.  Cosmochim.  Acta,  39,  1077-1084, 
1975. 
Domine,  R,  and  B.  Piriou,  Raman  spectroscopic 
study  of  the  Si02-Al203-K20  vitreous  sys- 
tem: Distribution  of  silicon  and  second  neigh- 
bors, Amer.  Mineral,  71,  38-50,  1986. 
Engelhardt,  G.,  M.  Nofz,  K.  Forkel,  F.  G. 
Wishmann,  M.  Magi,  A.  Samoson,  and  E. 
Lippmaa,  Structural  studies  of  calcium  alumi- 
nosilicate glasses  by  high  resolution  solid  state 
29Si  and  ^Al  magic  angle  spinning  nuclear 
magnetic  resonance,  Phys.  Chem.  Glasses,  26, 
157-165,  1985. 
Kirkpatrick,  R.  J.,  R.  Oestrike,  C.  A.  Weiss,  K.  A. 
Smith,  and  E.  Oldfield  High-resolution  27A1 
and  29Si  NMR  spectroscopy  of  glasses  and 
crystals  along  the  join  CaMgSi206-CaAl2Si06, 
Amer.  Mineral,  77,705-711,1986. 
Kohn,  S.  C.,  R.  Dupree,  and  M.  E.  Smith,  A 
Nuclear  magnetic  resonance  study  of  the  struc- 
ture of  hydrous  albite  glasses,  Geochim. 
Cosmochim.  Acta,  53,  2925-2935,  1989. 
Manning,  D.  A.  C,  C.  M.  B.  Hamilton  C.  M.  B. 
Henderson,  and  M.  J.  Dempsey,  The  probable 
occurrence  of  interstitial  Al  in  hydrous  F- 
bearing  and  F-free  aluminosilicate  melts, 
Contr.  Mineral.  Petrol,  75,  257-262,  1980. 
Matson,  D.  W.,  S.  K.  Sharma,  and  J.  A.  Philpotts, 
The  structure  of  high-silica  alkali-silicate 
glasses  — A  Raman  spectroscopic  investiga- 
tion, /.  Non-Cryst.  Solids,  58,  323-352, 1983. 
Merzbacher,  C,  B.  L.Sherriff,  S.  J.  Hartman,  and 
W.  B.  White,  A  high-resolution  29Si  and  27A1 
NMR  study  of  alkaline  earth  aluminosilicate 
glasses,  /.  Non-Cryst.  Solids,  124,  194-206, 
1990. 
Mustart,  D.  A.  Phase  relations  in  the  peralkaline 
portion  of  the  systemNa20-Al203-Si02-H20 . 
Ph.  D.  thesis,  Stanford  University  ,  1972. 
Mysen,  B.  O.,  Effect  of  pressure,  temperature,  and 
bulk  composition  on  the  structure  and  species 
distribution  in  depolymerized  alkali  alumino- 
silicate melts  and  quenched  melts,  J.  Geophys. 
Res.,  95,  15733-15744,  1990. 


GEOPHYSICAL  LABORATORY 


61 


My  sen,  B.  O.,  Structure  and  Properties  of  Silicate 
Melts,  354  pp.,  Elsevier,  Amsterdam,  1988. 
Mysen,  B.  O.,  D.  Virgo,  and  F.  A.  Seifert,  The 
structure  of  silicate  melts:  Implications  for 
chemical  and  physical  properties  of  natural 
magma,  Rev.  Geophys.,  20,  353-383,  1982. 
Mysen,  B.  O.,  and  D.  Virgo,  The  structure  of  melts 
in  the  system  Na20-CaO-Al2C>3-Si02-H20 
quenched  from  high  temperature  at  high  pres- 
sure. 2.  Water  in  melts  along  the  join  NaAlC>2- 
Si02  and  a  comparison  of  solubility  mecha- 
nisms of  water  and  fluorine,  Chem.  GeoL,  57, 
333-358,  1986. 
Mysen,  B.  O.,  D.  Virgo,  and    I.  Kushiro,  The 
structural  role  of  aluminum  in  silicate  melts  — 
A  Raman  spectroscopic  study  at  1  atmosphere, 
Amer.  Mineral,  66,678-701,1981. 
Oestrike,  R.,  W.-H.  Yang,  R.  J.  Kirkpatrick,  R. 
Hervig,  A.  Navrotsky,  and  B.  Montez,  High- 
resolution  23Na,  27A1  and  29Si  NMR  spectros- 
copy of  framework-aluminosilicate  glasses, 
Geochim.  Cosmochim.  Acta,  51,  2199-2210, 
1987. 
Osborn,  E.  F.,  and  A.  Muan,  Phase  equilibrium 
diagrams  for  ceramists.  Plate  4.  The  system 
Na20-Al203-Si02,  Am.  Ceram.  Soc.,  Colum- 
bus Ohio,  1960. 
Pichavant,  M.,  Effects  of  B  and  H2O  on  liquidus 
phase  relations  in  the  haplogranite  system, 
Amer.  Mineral,  72,1056-170,1987. 
Richet,  P. ,  and  Y.  Bottinga,  Thermochemical  prop- 
erties of  silicate  glasses  and  liquids:  A  review, 
Rev.  Geophys.,  24,  1-26,  1986. 
Stebbins,  J.  F.,.  Identification  of  multiple  struc- 
tural species  in  silicate  glasses  by  29Si  NMR, 
Nature,  330,  465-467,  1987. 
Virgo,  D.,  B.  O.  Mysen,  and  I.  Kushiro,  Anionic 
constitution  of  silicate  melts  quenched  at  1  atm 
from  Raman  spectroscopy:  Implications  for 
the  structure  of  igneous  melts,  Science,  208, 
1371-1373.,1980. 


Partitioning  of  Fluorine  and  Chlorine 

between  Apatite  and  Non-Silicate  Fluids 

at  High  Pressure  and  Temperature 


James  Brenan 


mineral  as  a  useful  indicator  of  halogen 
activities  in  the  geologic  environment.  Prior 
experimental  work  has  focused  on  calibrat- 
ing apatite  as  a  monitor  of  halogen  activi- 
ties in  the  low  pressure(f>)-temperature(7T) 
hydrothermal  environment  (Korzhinskiy, 
1981;  Latil  and  Maury,  1977).  Recently 
developed  solution  models  for  apatite,  flu- 
ids and  melts  now  provide  a  basis  for  theo- 
retical prediction  of  the  halogen  chemistry 
of  apatite  coexisting  with  these  phases  over 
a  somewhat  broader  range  of  P  and  T 
(Candela,  1986;PiccoliandCandela,  1991; 
Tacker  and  Stormer,   1989;  Zhu  and 
Sverjensky,  1991).   Although  such  work 
represents  an  important  contribution  to  the 
interpretation  of  the  halogen  chemistry  of 
natural  apatite,  its  usefulness  may  be  lim- 
ited to  the  overall  low  P-T  range  in  which 
experiments  were  performed  for  both  cali- 
bration and  development  of  solution-model 
databases.  In  order  to  exploit  apatite  as  a 
monitor  of  the  halogen  chemistry  of  fluids 
or  melts  in  the  high  P-T  environment,  ex- 
perimental determination  of  apatite  chem- 
istry at  these  conditions  is  a  requisite.  This 
report  describes  the  results  of  experiments 
aimed  at  this  goal  with  specific  emphasis 
on  measurements  of  the  distribution  of 
fluorine  and  chlorine  between  apatite  and 
non-silicate  fluids  (H2O  +  dissolved  salts  ± 
CO2,  carbonate  melt)  at  P-T  conditions 
appropriate  to  the  lower  crust  and  upper 
mantle. 


The  presence  of  fluorine  and  chlorine  as 
essential  structural  constituents  of  apatite, 
combined  with  the  widespread  occurrence 
of  apatite  amongst  diverse  parageneses, 
underscores  the  potential  utility  of  this 


Experimental  Technique 

Owing  to  the  difficulties  associated  with 
analyzing  fluids  and  low-viscosity  melts 


62 


CARNEGIE  INSTITUTION 


quenched  from  high  P-T experiments,  par- 
tition coefficients  (D-values;  wt  %  concen- 
tration in  apatite/wt  %  concentration  in 
fluid  or  melt)  were  determined  by  mass 
balance.  The  overall  strategy  was  therefore 
to  perform  experiments  by  (1)  encapsulat- 
ing known  quantities  of  finely-powdered, 
natural  apatite  (previously  well-character- 
ized in  terms  of  fluorine  and  chlorine  abun- 
dances) and  fluid,  or  apatite  and  melt,  in 
welded  Pt  containers,  (2)  equilibrating  this 
mixture  for  2-4  days  at  high  pressure  and 
temperature  (950-1050°C,  1.0-2.0  GPa) 
with  a  solid-media,  high-pressure  appara- 
tus, then,  (3)  analyzing  the  run-product 
apatite  for  F  and  CI  using  an  electron  mi- 
croprobe.  Fluids  were  added  to  these  ex- 
periments as  distilled  H2O  or  as  aqueous 
solutions  of  HC1  (1.8  and  5.3  wt  %),  NaCl 
(4, 15,  and  25  wt  %),  Na2CC>3  (4  and  15  wt 
%)  and  NaOH  (4  wt  %).  Water-carbon 
dioxide  mixtures  were  produced  by  weigh- 
ing in  distilled  H2O  with  silver  oxalate. 
The  carbonate  melt  used  in  these  experi- 
ments was  a  mixture  of  high-purity  carbon- 
ates with  the  stoichiometry  80  wt  %  dolo- 
mite: 20  wt  %  Na2CCh.  Water  was  added  to 
melt-bearing  experiments  to  yield  5-15  wt 
%  in  the  melt.  The  apatite  starting  material 
consisted  of  hand-picked  fragments  (free 
of  inclusions  and  visible  alteration)  of  gem- 
quality  crystals  that  originate  at  Durango, 
Mexico.  Microprobe  analysis  of  points  on 
25  different  apatite  fragments  yielded  aver- 
age concentrations  for  F  and  CI  of  3.57  (± 
0.13,  lo)  and  0.38  (±  0.06,  la)  wt  %, 
respectively.  A  rastered  beam  was  used  to 
minimize  mineral  degradation  during  mi- 
croprobe analysis,  and  no  loss  of  F  or  CI  x- 
ray  intensity  was  detected  even  for  data 


acquisition  times  exceeding  4  minutes;  stan- 
dard analytical  conditions  were  a  45  nA 
sample  current  at  15  kV  accelerating  volt- 
age. 

Due  to  the  low  solubility  of  apatite  in 
C02-H20-NaCl  and  dilute  HC1  solutions 
(i.e.,<l  wt%;Ayers  and  Watson,  1991),  no 
correction  for  apatite  dissolution  was  ap- 
plied to  calculated  partition  coefficients 
involving  these  compositions.  Based  on 
the  measurements  of  Ayers  and  Watson 
(1991),  a  solubility  of  3  wt  %  apatite  was 
used  in  partition  coefficient  calculations 
for  5.3  wt  %  HCl-bearing  experiments  and, 
combining  the  data  of  Baker  and  Wyllie 
(1990)  with  that  of  Watson  (1980),  a  solu- 
bility of  12.5  wt  %  apatite  was  used  to 
calculate  D-values  for  experiments  involv- 
ing carbonate  melt. 


100 1 — ■ — 1 — ■ — 1 — ■ — 1 — ■ — ' — ■ — > — ■ — ' — ■ — ' — 

^         if-A   - 2.2-2.6  Fluorme 

'=3 


.01 


.001 


*«*  rwv 

ft    TA  A      AA 

Chlorine 


m      D 


03  n 


2         4         6         8        10       12       14 
Wt  %  CI 


Fig.  31.  Apatite/fluid  partition  coefficients  for 
fluorine  (closed  symbols)  and  chlorine  (open  sym- 
bols) as  a  function  of  total  chlorine  abundance. 
Data  are  from  experiments  at  2.0  GPa  and  1050°C 
and  pertain  to  H20-bearing  fluids  with  and  with- 
out added  HC1  or  NaCl  (Circles,  triangles  and 
squares  refer  to  experiments  with  H2O,  H2O-HCI 
and  H20-NaCl,  respectively).)  The  numbered 
contours  for  the  fluorine  partitioning  data  refer  to 
the  total  fluorine  contents  of  these  experiments 
(rev  =  reversal). 


GEOPHYSICAL  LABORATORY 


63 


Results  and  Discussion 

Inspection  of  apatite  run  products  re- 
vealed substantial  grain-growth  during  the 
course  of  an  experiment:  samples  starting 
out  as  <10  Jim,  angular  grains  underwent 
coarsening  to  produce  subhedral  to  euhedral 
material  of  from  15  jum  to  mpre  than  100 
Jim  grain  size.  The  amount  of  grain  growth 
was  dependent  on  fluid  composition  and 
experiments  containing  HC1  or  Na-carbon- 
ate  solutions  or  carbonate  melt  produced 
the  most  abundant  large  grains.  Multiple 
microprobe  analyses  across  individual  apa- 
tite grains  from  run-products  showed  no 
evidence  of  compositional  zoning. 

Figures  31  and  32  portray  D-values  for 
fluorine  and  chlorine  as  functions  of  total 
chlorine  and  fluorine  concentration  mea- 
sured in  experiments  with  H20-NaCl-HCl 
fluids  at  1050  °C  and  2.0  GPa.  (Note  that  F 
concentrations  were  varied  by  adjusting 
the  ratio  of  apatite  to  fluid;  CI  contents 
were  varied  by  changing  the  CI  concentra- 
tion in  added  solutions.)  Average/) -values 
for  CI  from  experiments  containing  H2O  or 

100 


10 


0) 

a.  1 

(0 


.01 


.001 


Fluorine 

rev 

V 

*• 

m 

r 

Chlorine 

*»    A  ^-   reV 

v 

1 
4& 

*HCI 

D           O 

T 

H 

D 

-. 

2  NaCI 

, 

. 1 

: 

Oh2o 

1  2 

wt%F 


Fig.  32.  Apatite/fluid  partition  coefficients  for 
fluorine  and  chlorine  as  a  function  of  total  fluorine 
abundance  (symbols  as  in  Fig.31).  Data  are  from 
the  same  experiments  described  in  Fig.  31. 


HCl-solutions  are  -0.1  and  average  values 
obtained  from  aqueous  NaCl-bearing  ex- 
periments are  -0.015;  no  dependence  on 
either  absolute  F  or  CI  concentration  was 
seen  for  either  of  these  values.  Consistent 
with  previous  work  (Korzhinskiy,  1981; 
Latil  and  Maury,  1977),  D-values  for  F  are 
well  in  excess  of  those  measured  for  CI,  and 
are  generally  similar  at  a  given  total  CI 
concentration  (for  similar  total  F  contents), 
regardless  of  the  mode  of  CI  addition.  Par- 
tition coefficients  for  F  exhibit  no  system- 
atic dependence  on  total  CI  concentration 
(Fig.  31),  but  values  do  systematically  de- 
crease as  a  function  of  increasing  F  concen- 
tration (Fig.  32).  Results  of  a  reversal 
experiment  in  which  apatite  was  first  equili- 
brated in  H2O,  then  in  1.8%  HC1,  gave 
good  agreement  with  D-values  measured 
in  forward  experiments.  Average  partition 
coefficients  for  F  and  CI  in  basic  solutions 
(aqueous  NaOH  or  Na2C03)  are  ~5  and 
<0.02,  respectively. 

Figure  33  portrays  fluorine  and  chlorine 
D-values  as  a  function  of  pressure  for  ex- 
periments with  aqueous  HC1  or  NaCI  at 
1050°C.  D- values  for  CI  are  invariant  with 
pressure,  regardless  of  fluid  composition, 
as  is  the  D  for  F  in  the  HCl-bearing  fluid. 
The  F  partition  coefficient  with  aqueous 
NaCI  at  1.0  GPa  is,  however,  -10  times 
higher  than  values  determined  at  2.0  GPa 
for  similar  total  F  abundances.  D-values 
for  CI  at  950°C  and  2.0  GPa  (Fig.  33)  are 
identical  to  values  measured  at  1050°C 
whereas  F  partition  coefficients  are  4-10 
times  higher. 

Fluorine  and  chlorine  partition  coeffi- 
cients generally  decrease  and  increase,  re- 
spectively, as  a  function  of  the  mole  frac- 


64 


CARNEGIE  INSTITUTION 


100 


10 

p 

=3 

5=      1 
0) 

■«— » 
TO 


Fluorine 


.01 


.001 


950°C 


HCI 


Chlorine 
A — 


950°C 


1.0  1.5  2.0 

Pressure  (GPa) 


2.5 


Fig.  33.  Apatite/fluid  partition  coefficients  for 
fluorine  (closed  symbols)  and  chlorine  (open  sym- 
bols) as  a  function  of  pressure  for  experiments  at 
950  and  1050°C  (1050°C  data  unlabelled).  Data 
are  from  experiments  with  2.2-2.5  wt  %  total 
fluorine  (variable  chlorine  contents)  with  added 
HCI  (circles)  or  NaCl  (triangles). 

Hon  of  CO2  [X(C02>]  in  H2O-CO2  mix- 
tures (Fig.  34).  In  terms  of  D-values  for  CI, 
this  relation  is  greatly  altered  for  experi- 
ments in  which  CI  was  added  as  solutions 
with  >15  wt  %  NaCl  (Fig.  35;  F  partition 
coefficients  are  relatively  unaffected  by 


100 


10  ■ 


■g 

"=3 


0)       . 
♦J 

03 
Q. 

Si 


.01 


1          ■ 

> 
> 

• 

• 

• 

■ 

• 

fluorine 

r 

O 

chlorine 

* 

<D 

<D       ; 

> 

) 

■ 

— 1__ 

■ 

■ 

■ 

0.0 


0.2  0.4  0.6 

Mol  Frac  C02 


0.8 


Fig.  34.  Apatite/fluid  partition  coefficients  for 
fluorine  (closed  symbols)  and  chlorine  (open  sym- 
bols) as  a  function  of  the  mole  fraction  of  CO2  for 
experiments  involving  CO2-H2O  mixtures.  Data 
are  from  experiments  run  at  1050°C  and  2.0  GPa. 


Q. 

i  -1 

O 

SI 

O 
Q 


.01 


i 


H20 


XC02  =  0.3-0.5 
5%  HCI 


4%  NaCl 


15%NaCI 
25%  NaCl 


1  2  3 

Total  CI  (wt  %) 

Fig.  35.  Apatite/fluid  partition  coefficients  for 
chlorine  as  a  function  of  total  chlorine  concentra- 
tion. Data  are  from  experiments  involving  CO2- 
H2O  fluids  with  mole  fractions  of  CO2  =  0.3-0.5. 
Data  labels  refer  to  experiments  with  no  added 
chlorine  (H2O;  CI  is  from  the  apatite  starting 
material)  or  in  which  CI  was  added  as  5  wt  %  HCI 
or  4-25  wt  %  NaCl. 

these  fluid  composition  variations).  As 
illustrated  in  Fig.  35,  similar  D-values  for 
CI  (i.e.,  -0.25)  were  obtained  from  experi- 
ments in  which  CI  was  added  as  either  the 
apatite  starting  material,  4  wt  %  NaCl,  or 
5.3  wt  %  HCI.  Experiments  in  which  CI 
was  added  as  1 5  or  25  wt  %  NaCl,  however, 
exhibit  marked  drops  in  the  chlorine  D- 
value  (i.e.,  to  -0.065  and  -0.03,  respec- 
tively). Inasmuch  as  CI  partition  coeffi- 
cients were  found  to  be  independent  of  total 
CI  content  for  C02-free  fluids,  these  results 
may  be  somewhat  surprising.  The  overall 
reduction  in  the  CI  partition  coefficient 
could,  however,  be  accounted  for  if  an 
H20-rich  fluid  evolved  as  a  result  of  fluid 
unmixing  in  the  runs  with  NaCl-rich  com- 
positions. The  lower  chlorine  D-  value  found 
for  the  experiment  with  25  wt  %  NaCl, 
compared  to  that  with  15  wt  %  NaCl,  may 
therefore  suggest  a  higher  proportion  of 
this  H20-rich  fluid  in  the  former  experi- 
ments. Figure  36  portrays  fluid  composi- 


GEOPHYSICAL  LABORATORY 


65 


NaCI 


10 


Fig.  36.  Bulk  composition  of  fluids  in  the  system 
C02-H20-NaCl  (wt  %)  for  experiments  involving 
CO2-H2O  fluids  in  which  H26  was  added  as  4-25 
wt  %  NaCI  solutions  (see  caption  to  Fig.  3  5  for 
more  details).  The  tie-lines  and  phase  field  bound- 
ary are  schematic  but  meant  to  be  consistent  with 
the  observed  partitioning  relations  (see  text). 

tions  for  the  experiments  involving  mixed 
CO2-H2O  fluids  with  4,  15  and  25  wt  % 
NaCI  plotted  in  the  C02-H20-NaCl  ter- 
nary system.  Also  shown  in  this  figure  is  a 
topology  for  the  two-phase  field  that  is 
consistent  with  the  above  observations. 
Although  no  previous  experimental  mea- 
surements at  high  P  and  Thave  been  made 
with  regard  to  the  extent  of  immiscibility 
for  C02-H20-NaCl  fluids ,  results  at  low  P- 
T  conditions  (Popp  etal.,  this  Report)  indi- 
cate that  the  compositional  range  of  the 
two-fluid  field  in  this  system  can  be  exten- 
sive. Experiments  employing  synthetic 
fluid  inclusions  are  now  in  progress  in  an 
attempt  to  confirm  this  interpretation  of  the 
partitioning  data. 

Partition  coefficients  for  fluorine  and 
chlorine  between  apatite  and  carbonate  melt 
(obtained  at  1050°C,  2.0  GPa)  as  a  function 
of  total  wt  %  chlorine  are  shown  in  Fig.  37. 
Fluorine  D-values  are  lower  than  those 
measured  for  aqueous  fluids  (i.e.,  -1.5  vs 
>5,  respectively),  whereas  chlorine  parti- 
tion coefficients  (-0.07)  are  similar  to  val- 
ues measured  for  experiments  with  H2O 
and  HCl-bearing  solutions.  Partition  coef- 


0) 

E 


o 


Q-    1 


.01 


Molten  Carbonate  (Dol80:NaCarb20) 
5-1 5  wt  %  H20 


rev 


rev 


m 


m 


Fluorine 


Chlorine 


12  3  4  5 

Wt  %  CI 

Fig.  37.  Apatite/carbonate  melt  partition  coeffi- 
cients for  fluorine  (squares)  and  chlorine  (circles) 
as  a  function  of  total  chlorine  concentration  (ob- 
tained in  experiments  run  at  2.0  GPa  and  1050°C). 
Data  pertain  to  experiments  with  5-15  wt  %  H2O 
in  the  melt  phase. 


ficients  for  both  F  and  CI  show  little  varia- 
tion with  total  CI  content  (all  experiments 
had  F  abundances  of  -2.0-2.3  wt  %).  A 
reversal  experiment  that  involved  a  two- 
hour  preheating  of  the  sample  at  1300°C 
and  2.0  GPa  (thus  completely  dissolving 
the  apatite  into  the  melt),  then  a  slow, 
isobaric  cooling  to  the  final  run  condition 
of  1050°C  (and  holding  there  for  48  hours), 
yielded  identical  D-values  as  in  forward 
experiments. 


Implications  for  the  Fluorine 

and  Chlorine  Content  of 

Upper-Mantle  Fluids 

Testimony  bearing  on  the  action  of  flu- 
ids in  the  upper  mantle  may  be  found  in 
certain  suites  of  ultramafic  xenoliths  that 
contain  evidence  for  mineral  replacement 
by  volatile-bearing  phases  (O'Reilly  and 


66 


CARNEGIE  INSTITUTION 


100 


meg  aery  sts 
A      H20-HCI 

A      CO 3  Melt 

C-bearing  (xenoliths) 
□  C02-H20-HCI 
■      C03  Melt 


1  10 

Wt%F 


Fig.  38.  Calculated  fluorine  and  chlorine  contents 
of  high  P-T  fluids  based  on  the  compositions  of 
mantle-derived  apatites.  Fluid  compositions  were 
calculated  using  the  partition  coefficients  mea- 
sured in  this  study  for  fluids  capable  of  dissolving 
appreciable  amounts  of  apatite  (i.e.,  H2O-HCI  (± 
CO2)  fluids  or  carbonate  melt).  Apatite  composi- 
tions from  mantle  xenolith  parageneses  were  ob- 
tained from  Wassetal.  (1980),  Smith etal.  (1981) 
and  Exley  and  Smith  (1982).  Carbon-bearing 
apatite  compositions  (xenolith  paragenesis)  are 
from  O'Reilly  and  Griffin  (1988)  and  apatite 
megacryst  compositions  were  obtained  from 
Hervig  and  Smith  (1981)  and  Irving  and  Frey 
(1984). 


Griffin,  1988)  or  preserve  textures  indica- 
tive of  a  free  vapor  (Kovalenko  etal.,  1987) 
or  both.  Associated  with  these  petrographic 
indications  for  the  infiltration  of  fluids  are 
elevated  concentrations  of  elements  that 
are  typically  present  at  only  low  levels  in 
mantle  rocks  (e.g.,  Ba,  Cs,  Sr  and  the  rare- 
earth  elements).  Analyses  of  the  fluorine 
and  chlorine  content  of  apatites  that  occur 
as  a  minor  phase  in  some  such  fluid-altered 
rocks,  combined  with  the  results  presented 
here,  may  thus  provide  constraints  on  the 
halogen  abundances  of  some  mantle  meta- 
somatic  agents. 

Inasmuch  as  D-values  for  fluorine  and 
chlorine  were  found  to  depend  on  fluid 
composition,  accurate  estimates  of  mantle 
fluid  halogen  contents  will,  therefore,  be 
contingent  on  a  judicious  choice  of  parti- 


tion coefficients.  Based  on  results  of  the 
apatite  solubility  experiments  of  Baker  and 
WyUie(1990)andAyersandWatson(1991), 
the  non-silicate  fluids  capable  of  transport- 
ing the  most  significant  quantities  of  apa- 
tite are  molten  carbonate  and  HCl-bearing 
aqueous  solutions.  By  applying  the  D- 
values  measured  for  those  compositions  (at 
1050°C,  2.0  GPa)  to  analyses  of  apatites 
from  mantle  parageneses  (i.e.,  present  as 
interstitial  grains  in  xenoliths  or  as 
megacrysts;  see  caption  to  Fig.  38  for  data 
sources),  fluorine  and  chlorine  contents  of 
fluids  that  may  have  coexisted  with  mantle 
apatites  were  calculated  and  are  portrayed 
in  Fig.  38.  Two  of  the  apatites  analyzed  by 
O'Reilly  and  Griffin  (1988)  contain  car- 
bon abundances  of -0.3  and  -0.9  wt  %  and 
thus  may  preserve  evidence  for  equilibra- 
tion with  a  carbon-bearing  fluid;  partition 
coefficients  for  a  CO2-H2O-HCI  fluid 
[X(C02)  =  0.3-0.5]  were  therefore  used  in 
place  of  the  aqueous  HC1  values. 

Based  on  the  observed  P-T  dependence 
of  D-values  found  in  this  study,  equilibra- 
tion of  apatite  with  fluids  at  lower  P-T 
conditions  than  that  assumed  for  these  cal- 
culations would  not  effect  calculated  CI 
abundances  but  calculated  F  contents  might 
represent  maximum  values.  Fluid  compo- 
sitions determined  in  the  above  manner 
have  minimum  Cl/F  ratios  of  >  1,  F  abun- 
dances <1  wt  %  and  chlorine  concentra- 
tions of  from  ~1  to  -20  wt  %.  The  overall 
Cl-rich  nature  of  some  calculated  fluid 
compositions  may  suggest  an  important 
role  for  chlorine-bearing  fluids  as  agents  of 
mass  transport  in  the  upper-mantle.  This 
result  is  in  concert  with  the  measurements 
of  Brenan  and  Watson  (1991),  who  ob- 


GEOPHYSICAL  LABORATORY 


67 


served  significantly  greater  partitioning  of 
trace  elements  into  aqueous  chloride  solu- 
tions (coexisting  with  olivine)  relative  to 
experiments  involving  pure  H2O.  In  addi- 
tion, both  Selverstone  et  al.  (1990)  and 
Philippot  and  Selverstone  (1991)  have  docu- 
mented the  occurrence  of  trace-element 
rich  brines  in  Alpine  eclogites,  and  thus 
chlorine-rich  fluids  may  also  play  a  role  in 
trace  element  mobilization  in  other  high- 
pressure  environments. 


References 

Ayers,  J.  C.  and  E.  B.  Watson,  Solubility  of 
apatite,  monazite,  zircon,  and  rutile  in 
supercritical  fluids  with  implications  for  sub- 
duction  zone  geochemistry,  Phil.  Trans.  R. 
Soc.  Lond.  A,  in  press,  1991 

Baker,  M.  B.  and  P.  J.  Wyllie,  High-pressure 
solubility  of  apatite  in  carbonate-rich  melt 
(abstr),  EOS,  70,  1394,  1990. 

Brenan,  J.  M.  and  E.  B.  Watson,  Partitioning  of 
trace  elements  between  olivine  and  aqueous 
fluids  at  high  P-T  conditions:  Implications  for 
the  effect  of  fluid  composition  on  trace  ele- 
ment transport,  Earth  Planet.  Sci.  Lett.,  in 
press,  1991. 

Candela,  P.  A.,  Toward  a  thermodynamic  model 
for  the  halogens  in  magmatic  systems:  an 
application  to  melt-vapor-apatite  equilibria, 
Chem.  Geoi,  57,  289-301,  1986. 

Exley,  R.  A.  and  J.  V.  Smith,  The  role  of  apatite  in 
mantle  enrichment  processes  and  in  the 
petrogensis  of  some  alkali  basalt  suites, 
Geochim.  Cosmochim.  Acta,  46, 1375-1384, 
1982. 

Hervig,  R.  L.  and  J.  V.  Smith,  Dolomite-apatite 
inclusion  in  chrome-diopside  crystal,  Bellsbank 
kimberlite,  South  Africa,  Amer.  Mineral.,  66, 
346-349. 

Irving,  A.  J.  and  F.  A.  Frey,  Trace  element  abun- 
dances in  megacrysts  and  their  host  basalts: 
Constaints  on  partition  coeffcients  and 
megacryst  genesis,  Geochim.  Cosmochim. 
Acta,48,  1201-1221,  1984. 


Korzhinskiy,  M.  A.,  Apatite  solid  solution  as 
indicators  of  the  fugacity  of  HC1°  and  HF°  in 
hydrothermal  fluids,  Geochem.  Int.,  18,  44- 
60,  1981. 
Kovalenko,  V.  I.,  I.  P.  Solovova,  I.  D.  Ryabchikov, 
D.  A.  Ionov,  O.  A.  Bogatikov  and  V.  B. 
Naumov,  Fluidized  C02-sulphide-silicate 
media  as  agents  of  mantle  metasomatism  and 
megacrysts  formation:  evidence  from  a  large 
druse  in  a  spinel-lhezolite  xenolith,  Phys.  Earth 
Planet  Int.,  45,280-293,1987. 
Latil,  C.  and  R.  Maury,  Contribution  a  l'etude  des 
echanges  d'ions  OH*,  CI"  et  F"  et  de  leur 
fixation  dans  les  apatites  hydrothermales,5«//. 
Soc.  Fran.  Mineral.  Cristall.,  100,  246-250, 
1977. 
O'Reilly ,  S.  Y.  and  W.  L.  Griffin,  Mantle  metaso- 
matism beneath  western  Victoria,  Australia: 
1.  Metasomatic  processes  in  Cr-diopside 
lherzolites,  Geochim.  Cosmochim.  Acta,  52, 
433-447,  1988. 
Philippot,  P  and  J.  Selverstone,  Trace-element- 
rich  brines  in  eclogite  veins:  implications  for 
fluid  composition  and  transport  during  sub- 
duction,  Contrib.  Mineral.  Petrol.,  106,  417- 
430,  1991. 
Piccoli,  P.  M.  and  P.  A.  Candela,  The  mathemati- 
cal modeling  of  the  halogen  composition  of 
the  mineral  apatite  during  first  and  second 
boiling  (abstr),  EOS,  72,  312,  1991. 
Smith,  J.  V.,  J.  S.  Delaney,  R.  L.  Hervig,  and  J.  B. 
Dawson,  Storage  of  F  and  CI  in  the  upper 
mantle:  geochemical  implications,  Lithos,  14, 
133-147,  1981. 
Tacker,  R.  C.  and  J.  C.  Stormer,  A  thermodynamic 
model  for  apatite  solid  solutions,  applicable  to 
high-temperature  geologic  problems,  Amer. 
Mineral,  74,  877-888,  1989. 
Wass,  S .  Y. ,  P.  Henderson  and  C.  J.  Elliott,  Chemi- 
cal heterogeneity  and  metasomatism  in  the 
upper  mantle:  Evidence  from  rare  earth  and 
other  elements  in  apatite-rich  xenoliths  in  ba- 
saltic rocks  from  eastern  Australia,  Phil.  Trans. 
R.  Soc.  Lond.  A,  297,  333-346,  1980. 
W'atson,  E.  B.,  Apatite  and  phosphorous  in  mantle 
source  regions:  an  experimental  study  of  apa- 
tite/melt equilibria  at  pressures  to  25  kbar, 
Earth  Planet.  Sci.  Lett.,  51,  322-335,  1980. 
Zhu,  C.  and  D.  A.  Sverjensky,  A  set  of  consistent 
thermodynamic  properties  for  fluorapatite, 
hydroxylapatite  and  chlorapatite  (abstr),  EOS, 
72,  145,  1991. 


68 


CARNEGIE  INSTITUTION 


Investigation  of  Fluid  Immiscibility  in 

the  System  H2O-NACL-CO2  Using  Mass 

Spectrometry  and  Microthermometry 

Techniques  Applied  to 

Synthetic  Fluid  Inclusions 

Robert  K.  Popp?  John  D.  Frantz,  and 
Thomas  C.  Hoering 

Popp  and  Frantz  (1990)  described  the 
use  of  synthetic  fluid  inclusions  to  define 
the  limits  of  fluid  miscibility  in  the  system 
H20-NaCl-C02.  Inclusions  were  trapped 
in  quartz  prisms  contained  in  fluids  with 
sodium  chloride  and  carbon  dioxide  con- 
tents up  to  1 8  and  50  wt  %,  respectively,  at 
temperatures  of  500°,  600°,  and  700°C  and 
pressures  of  1000, 2000,  and  3000  bar.  The 
presence  of  two  different  types  of  inclu- 
sions within  a  single  sample  was  inter- 
preted as  evidence  of  immiscibility  for  a 
particular  fluid  composition,  temperature, 
and  pressure.  Samples  that  entrapped  im- 
miscible fluids  exhibited  both  (1)  inclu- 
sions containing  a  salt  crystal  in  addition  to 
a  vapor  bubble  and  (2)  inclusions  without 
salt  crystals,  but  with  a  much  larger  bubble 
than  in  the  first  case.  The  former  type 
inclusion  represents  the  quenched  high- 
density  (i.e.  NaCl-rich,  C02-poor)  phase, 
whereas  the  latter  type  represent  the 
quenched  low-density  (i.e.  C02-rich,  NaCl- 
poor)  phase.  Inclusions  that  formed  in  ex- 
periments in  which  the  fluid  was  in  the 
miscible  field  contained  only  a  single  type 
that  did  not  contain  NaCl  crystals  and  had 
proportionally  the  same  ratio  of  bubble  to 


Department  of  Geology,  Texas  A&M  Univer- 
sity, College  Station,  Texas  77843 


total  inclusion  volume.  This  report  de- 
scribes the  use  of  mass  spectrometric  tech- 
niques and  microthermometric  measure- 
ments to  define  more  precisely  the  location 
of  the  solvus  separating  the  one -phase  and 
two-phase  regions  in  T-P-X  space,  and  to 
define  the  chemical  compositions  of  the 
coexisting  high-density  and  low-density 
fluid  phases. 


Mass  Spectrometry 

A  Finnigan  4500  quadrupole  mass  spec- 
trometer was  modified  (Fig.  39)  for  the 


Fig.  39.  Modification  of  Finnigan  4500  mass 
spectrometer.  See  text  for  details. 

analyses  of  individual  and  groups  of  fluid 
inclusions  using  the  general  technique  of 
Barker  and  Smith  (1986),  as  modified  by 
Frantz  et  al.  (1989).  A  crushed  sample  of 
the  quartz  prism  (-20  mg)  was  placed  in  a 
silica  tube  (B)  surrounded  by  a  resistance 
heater  (A).  The  tube,  sealed  at  the  outer 
end,  was  inserted  into  a  flight  tube  (C) 
extending  through  the  existing  vacuum  lock 
of  the  mass  spectrometer.  The  tube  is 
designed  to  stretch  out  the  arrival  time  at 
the  ion  source  of  the  pulses  of  released 
gases  resulting  from  inclusion  decrepita- 
tions. The  flight  tube  connects  with  a  modi- 
fied electron  beam  cup  (G)  and  focuses 
most  of  the  released  gas  molecules  into  the 
path  of  the  electron-beam  where  they  are 


GEOPHYSICAL  LABORATORY 


69 


390 


Temperature,  °C 

400 


Temperature,  °C 


410 
— I 


408 


418 


428 


H,0* 


_K_»i , — lJ-a 


h;5jJXm^^ 


—      COfe 


*K~* 


jJv. 


—       C02 


.wjs~ 


'      I      ■ 


J I 1_ 


J I I I 1_ 


J I I I I i_ 


21000  21500  22000  22500 

Scan  Number 


i i i i i i i i i i i i i i i i ■'■■ 

23000  23500  24000  24500  25000 

Scan  Number 


Fig.  40.  Mass  spectrograms  showing  H20+  and  C02+  intensities  for  fluid-inclusion  decrepitations  from 
(a)  a  quartz  sample  equilibrated  at  500°C,  1000  bar  with  a  fluid  of  composition  NaCl7.8-(C02)25.2- 
H2C>67.o(wt  %)  and  (b)  a  quartz  sample  equilibrated  at  500°C,  1000  bar  with  a  fluid  of  composition 
NaCl9.4-H2O80.6-(CO2)i0.0  (wt  %).  The  spectrogram  in  Fig.  40a  demonstrates  the  existance  of  two  types 
of  inclusions  having  different  CO2/H2O  ratios;  the  spectrogram  in  Fig  40b,  the  existence  of  one  type  of 
inclusion. 


ionized.  These  ions  are  then  accelerated 
and  separated  by  the  quadrupole  mass  filter 
(D)  and  detected  by  the  ion  multiplier  de- 
tector (E).  A  cryogenic  pump  (F),  chilled 
with  liquid  nitrogen,  is  used  to  reduce  the 
water  background  at  mass  18.  Because  the 
mixtures  of  low  molecular  weight  gases 
analyzed  in  this  study  are  simple  compared 
to  the  organic  compounds  often  analyzed 
with  the  instrument,  the  selected  ion  moni- 
toring mode  of  the  INCOS  2300  data  sys- 
tem of  the  mass  spectrometer  was  used. 
Only  the  selected  masses  of  interest  were 
measured.  A  total  scan  time  of  28  millisec- 
onds for  two  masses  was  utilized. 

The  mass  spectrometer  was  used  to  ana- 
lyze the  ratio  of  CO2  and  H2O  in  the  fluid 
inclusions  by  monitoring  channels  corre- 
sponding to  masses  18  (H2O4")  and  44 
(CO24").  Approximately  45,000  scans  were 
collected  as  the  samples  were  heated  from 
200  to  600 °C  at  5  °C  per  minute.  Decrepi- 
tation of  the  inclusions  generally  occurred 
between  300  and  573  °C  (the  latter  being 
the  approximate  temperature  of  the  oc-p 
transition  for  quartz).  A  spectrogram  from 


the  analysis  of  inclusions  from  an  experi- 
ment at  composition  NaO7.8-CO225.2~ 
H2O67.O  equilibrated  at  500°C  and  1000 
bar  is  shown  in  Fig.  40a.  The  2000  scans 
shown  in  the  figure  correspond  to  a  rise  in 
the  sample  temperature  from  390°  to  410°C 
over  a  time  period  of  approximately  240 
seconds.  The  peaks,  which  correspond  to 
the  decrepitations  of  individual  and  mul- 
tiple fluid  inclusions,  cover  a  range  varying 
from  10  to  15  spectrometer  scans.   Indi- 
vidual peaks  represent  only  nanogram 
amounts  of  H2O  and  CO2.  The  arrival 
times  (scan  number)  and  shapes  of  the  CO2 
and  H2O  peaks  are  quite  similar,  though  the 
sensitivity  for  water  tends  to  be  somewhat 
less  than  that  of  carbon  dioxide  due  to  the 
tendency  of  water  molecules  to  absorb  on 
the  surface  of  the  flight  tube.  Fluid  immis- 
cibility  clearly  exists,  as  evidenced  by  the 
large  variability  in  the  area  ratios  of  CO2 
and  H2O.  In  the  case  of  fluid  of  composi- 
tion NaCl9.4-CO2i0.0-H2O80.6  equilibrated 
at  500°C  and  1000  bar  (Fig.  40b),  a  single 
fluid  phase  was  present  at  the  experimental 


70 


CARNEGIE  INSTITUTION 


O 
0 


Mean  =  0.406 
Std  Dev  =  0.039 


40      60      80     100 


Area  %,  C02 


40      60      80     100 


Area  %,  C02 


NaCI 


H20     10     20    30     40     50 


CO- 


Fig.  41.  Histograms  of  the  frequency  of  the  area  %  CO2  [areaco2(/areaco2+areaH20)]  from  mass 
spectrograms  resulting  from  decrepitation  of  fluid  inclusions  All  five  samples  were  equilibrated  with 
a  series  of  compositions  at  500°C,  2000  bar  in  the  single-phase  fluid  compositional  region,  as  shown 
in  the  ternary  diagram  in  the  lower  right. 


run  conditions,  because  the  area  ratios  of 
CO2  and  H2O  are  nearly  identical  for  all  the 
peaks. 

Five  samples  equilibrated  at  500°C  and 
2000  bar  containing  less  than  5.2  wt  % 
NaCI  with  varying  CO2  contents  were  se- 
lected for  calibration  of  the  spectrometer. 


Based  on  the  optical  detection  of  only  a 
single  inclusion  type  in  each  sample,  at  was 
concluded  that  the  five  samples  have  trapped 
miscible  fluids.  For  all  five  samples,  areas 
under  corresponding  CO2  and  H2O  peaks 
resulting  from  decrepitations  were  com- 
puted, using  the  INCOS  2300    software 


GEOPHYSICAL  LABORATORY 


71 


10      20    30      40     50     60 

wt  %  co2 

Fig.  42.  Calibration  curve  showing  the  relation 
between  area  %  CO2  measured  by  mass  spectrom- 
etry and  wt  %  CO2  of  the  equilibrated  fluid  for  the 
samples  shown  in  Fig.  41.  The  squares  represent 
the  mean  values  of  the  measured  area  percents 
with  the  brackets  indicating  one  standard  devia- 
tion. 

standard  with  the  Finnigan  4500  spectrom- 
eter. Histograms  of  area  %  CO2  measured 
for  the  peaks  are  shown  for  the  five  samples 
in  Fig.  41.  Fig.  42  shows  the  mean  values 
for  all  five  samples,  with  their  correspond- 
ing standard  deviations,  plotted  against  the 
concentrations  of  CO2  (in  wt  %)  initially 
added  to  the  experimental  charges.  The 
second-order  quadratic  least  squares  fit  of 
these  data  was  then  used  as  the  calibration 
curve  to  define  the  wt%  CO2  in  inclusions 
grown  in  the  two-phase  region,  for  which 
C02-contents  were  unknown. 

Eight  samples  equilibrated  at  500°C  and 
1000  bar  demonstrate  the  results  obtained 
for  inclusions  grown  in  the  immiscible 
field  (Fig.  43).  Histograms  labelled  1  and  2 
are  from  samples  that  trapped  miscible 
fluids  so  that  their  mean  C02-contents  cor- 
respond closely  to  the  original  CO2  con- 
tents of  the  fluid,  denoted  by  the  vertical 
dashed  lines.  The  other  six  histograms, 
however,  correspond  to  samples  that  trapped 
immiscible  fluids.  In  histograms  3  through 
6,  the  intervals  exhibiting  the  highest  fre- 


quency are  generally  at  much  lower  values 
of  wt  %  CO2  than  the  initial  bulk  composi- 
tion, but  some  more  C02-rich  intervals 
contain  smaller  populations.  The  highly- 
populated  intervals  of  low  wt  %  CO2  rep- 
resent the  concentration  of  cabon  dioxide 
in  the  sodium  chloride-rich,  high-density 
fluid  phase.  In  the  case  of  histogram  7,  for 
which  the  original  bulk  fluid  composition 
lies  extremely  close  to  the  carbon  dioxide- 
rich  limb  of  the  solvus,  the  interval  of 
highest  frequency  lies  in  the  region  of  the 
original  bulk  composition,  with  less  popu- 
lated intervals  lying  at  lower  CO2  concen- 
trations. In  this  case,  the  highly  populated 
intervals  correspond  to  the  concentration 
of  carbon  dioxide  in  the  C02-rich,  low- 
density   fluid  phase.      Histogram   8 
demonstates  a  case  in  which  highly  popu- 
lated intervals  extend  from  the  original 
bulk  composition  down  to  those  represent- 
ing quite  low  concentrations  of  CO2.  Be- 
cause the  two  immiscible  phases  are  likely 
to  be  intimately  intermixed  rather  than  sepa- 
rated into  a  single  high-density  and  a  single 
low-density  phase  within  the  capsule 
(Ramboz  et  al.,  1982;  Zhang  and  Frantz, 
1989),  many  inclusions  trap  varying  pro- 
portions of  the  two  fluids;  thus,  a  range  of 
inclusions  with  properties  intermediate 
between  the  two  end  members  is  com- 
monly observed  in  a  single  sample  from  the 
immiscible  field.  The  less-populated  inter- 
mediate intervals  result  from  decrepita- 
tions of  inclusions  containing  mixtures  of 
the  two  fluids  and  from  simultaneous  de- 
crepitations of  groups  of  vapor-rich  and 
liquid-rich  inclusions. 

The  use  of  mass  spectrometry  measure- 
ments resulted  in  an  expanded  two-phase 


72 


CARNEGIE  INSTITUTION 


0     20    40     60     80   100 

Wt%C02 


NaCI 
/ 


l^£v 


0     20    40    60     80   100 

Wt%C02 


H2Q    10    20     30     40     50     C02 


Fig.  43.  Histograms  of  the  frequency  of  the  values  of  wt  %  CO2  from  mass  spectrograms  resulting  from 
decrepitation  of  fluid  inclusions  grown  at  500°C,  1000  bar.  The  values  of  area  %  CO2  were  converted 
to  wt%  CO2  using  the  calibration  curve  shown  in  Fig.  42. 


NaCI  + 
Solution 


20     30     40      50     60 

Wt  %  NaCI 

Fig.  44.  Solubility  diagram  (in  wt  %)  of  NaCI  in 
NaCl-H20  solutions  as  a  function  of  temperature 
(data  from  Linke,  1958,  1965). 


field  relative  to  that  obtained  from  only  the 
optical  identification  of  the  two  inclusion 
types.  That  is,  the  mass  spectrometry  tech- 
nique has  the  increased  precision  neces- 


sary to  identify  inclusions  of  more  than  one 
composition  in  some  samples  where  only 
one  inclusion  type  has  been  identified  op- 
tically. In  addition,  the  technique  permits 
the  CO2  content  of  the  high-density  fluid  to 
be  determined.  The  concentration  of  CO2 
in  the  high-density  fluid  is  taken  to  be  that 
of  the  highest  frequency  intervals  at  rela- 
tively low  weight  %  CO2  (Fig.  43). 


Microthermometry 

In  order  to  determine  the  positions  of 
tie-lines  connecting  the  compositions  of 
coexisting  immiscible  fluids,  the  concen- 
tration of  NaCI  in  the  high-density  fluid 
inclusions      was      estimated      from 


GEOPHYSICAL  LABORATORY 


73 


500  °C 
2000  bar 


500  °C 
1000  bar 


H20     10     20     30    40     50 


£Q  H20     10     20     30    40     50         C02 


NaCI 


NaCI 


600  °C 

1000  bar      __  f 

60 


H~0     10     20     30    40    50 


700  °C 

1 000  bar  60 


10 


CO         H2°    10    20    30    40    50 


CO. 


Fig.  45.  Compositions  of  coexisting  immiscible  fluids.  Each  tie  line  represents 
the  compositions  of  the  coexisting  fluid  as  determined  for  the  sample  denoted 
by  the  filled  circle  through  which  they  pass.  See  text  for  further  details. 


microthermometry  measurements  made 
using  a  heating  stage  supplied  by  Fluid 
Inc.,  Denver,  Colorado,  In  the  phase  dia- 
gram shown  in  Fig.  44,  the  NaCI  content  of 
a  given  bulk  composition  is  known  if  the 
temperature  of  the  univariant  boundary 
between  the  "Solution"  field  and  the  "NaCI 
+  Solution"  field  is  known  (i.e.,  if  the 
temperature  of  NaCI  melting  is  known). 
The  presence  of  CO2  in  the  fluid  inclusions 
might  affect  the  temperatures  obtained  from 
the  phase  relations  in  the  C02-free  system. 
However,  the  results  of  the  mass  spectrom- 


etry measurements  described  above  sug- 
gest that  the  bulk  concentration  of  carbon 
dioxide  in  the  high-density  inclusions  is 
relatively  low.  In  addition,  most  of  that 
carbon  dioxide  is  contained  in  the  vapor 
phase  (i.e.,  the  bubble)  at  the  temperatures 
of  NaCI  melting.  Therefore,  only  very 
small  CO2  concentrations  must  be  con- 
tained in  the  liquid  phase,  and  thus  the 
effect  of  CO2  on  the  temperature  of  NaCl- 
melting  is  considered  insignificant. 

The  melting  temperatures  of  NaCI  crys- 
tals in  the  high-density  inclusions  were 


74 


CARNEGIE  INSTITUTION 


determined  for  selected  samples  equili- 
brated at  500°C,  1000  and  2000  bar;  600°C, 
1000 bar;  and 700°C,  1000 bar.  Arelatively 
large  variation  in  the  melting  temperature 
was  observed  in  each  individual  sample, 
with  the  largest  frequency  occurring  at  the 
highest  temperatures.  The  lower-tempera- 
ture measurements  are  obtained  from  in- 
clusions that  trapped  mixtures  of  the  two 
immiscible  fluids,  and  therefore  contain 
lower  NaCl  concentrations  than  inclusions 
that  trapped  only  the  high-density  phase. 
With  knowledge  of    NaCl  contents,  ob- 
tained from  the  highest  NaCl  melting  tem- 
perature measured  in  a  given  sample,  and 
knowledge  of  the  CO2/H2O  ratios  mea- 
sured by  mass  spectrometry,  the  chemical 
composition  of  the  high-density  fluid  phase 
is  completely  defined.  Tie  lines  were  lo- 
cated between  the  coexisting  fluid  phases 
at  500°Cand2000bar,500°Cand  lOOObar, 
600°C  and  1000  bar,  and  700°C  and  1000 
bar  (Fig.  45).  To  construct  the  tie  lines,  a 
straight  line  was  drawn  from  the  composi- 
tion of  the  high-density  fluid  through  the 
bulk  composition  to  the  carbon  dioxide- 
rich  limb  of  the  solvus.  At  a  given  tempera- 
ture and  pressure,  the  slopes  of  the  tie  lines 
systematically  steepen  as  the  NaCl-H20 
binary  is  approached.  With  increasing  tem- 
perature, the  tie-lines  become  increasingly 
steeper,  as  a  result  of  greater  partitioning  of 
sodium  chloride  and  reduced  partitioning 
of  carbon  dioxide  between  the  two  immis- 
cible phases. 


Summary 
The  use  of  mass  spectrometry  in 


conjunction  with  the  more  routine  optical 
and  microthermometric  techniques  pro- 
vides an  improved  basis  on  which  to  sepa- 
rate fluid  inclusions  that  trapped  one -phase 
fluids  from  those  that  trapped  two-phase 
fluids.  In  addition,  the  technique  provides 
a  relatively  straightforward  method  to  de- 
fine tie  lines  between  the  compositions  of 
the  coexisting  high-density  and  low-den- 
sity fluids.  The  techniques  described  here 
should  be  applicable  to  other  systems  of 
petrologic  interest,  such  as  those  contain- 
ing the  volatile  species  CH4,  N2,  and  HC1. 

References 

Frantz,  J.D.,  Y.  Zhang,  D.  D.  Hickmott,  and  T.  C. 
Hoering,  Hydrothermal  reactions  involving 
equilibrium  between  minerals  and  mixed 
volatiles.  1 .  Techniques  for  experimentally  load- 
ing and  analyzing  gases  and  their  application  to 
synthetic  fluid  inclusions,  Chemical  Geol.,  76, 
57-70,  1989. 

Barker,  C.  and  M.  P.  Smith,  Mass  spectrometric 
determination  of  gas  in  individual  fluid  inclu- 
sions in  natural  minerals.  Anal.  Chem.,  58, 
1330-1333,  1986. 

Linke.  W.F.,  Solubilities  of  Inorganic  and  Metal- 
Organic  Compounds,  1.  Van  Norstrand, 
Princeton,  N.J.,  4th  ed.,  1958. 

Linke.  W.F.,  Solubilities  of  Inorganic  and  Metal- 
Organic  Compounds,  2.  Am .  Chem .  Soc . ,  Wash- 
ington D.C,  4th  ed.,  1965. 

Popp,  R.K.  and  J.D.  Frantz,  Fluid  immiscibility  in 
the  system  H20-NaCl-C02  as  determined  from 
synthetic  fluid  inclusions,  Annu.  Rep.  Director 
Geophys.  Lab.,  Carnegie  Instn.  Washington, 
1989-1990,43-47,  1990. 

Ramboz,  C,  M.  Pichavant,  and  A.  Weisbrod, 
Fluid  immiscibility  in  natural  processes:  Use 
and  misuse  of  fluid  inclusion  data,  II.  Interpre- 
tation of  fluid  inclusion  data  in  terms  of  immis- 
cibility, Chemical  Geol,  37,  29-48r  1982. 

Zhang,  Y.  and  J.  D.  Frantz,  Experimental  determi- 
nation of  the  compositional  limits  of  immisci- 
bility in  the  system  CaCl2-H20-NaCl  at  high 
temperatures  and  pressures  using  synthetic  fluid 
inclusions,  Chemical  Geol.,  74,289-308, 1989. 


GEOPHYSICAL  LABORATORY 


75 


The  Akermanite-Gehlenite-Sodium 

Melilite  Join  at  950°C  and  5  kbar  in  the 

Presence  of  CO2  +  H2O 

H.G.  Huckenholz,  H.S.  Yoder,  Jr.,  T. 
Kunzmann*  and  W.Seiberl* 

Melilites  are  primarily  solid  solutions 
between  akermanite  (Ca2MgSi20y),  so- 
dium melilite  (NaCaAlSi207),  and 
gehlenite  (Ca2Al2Si07).  High-grade  meta- 
morphism  of  impure  limestones  and  dolo- 
mites favors  the  crystallization  of  members 
of  the  akermanite  -  gehlenite  solid  solution 
series,  and  the  sodium  melilite  component 
is  generally  low.  During  the  decarbonation 
process,  CO2  and  H2O  play  an  active  role 
and  greatly  influence  the  metamorphic  as- 
semblages. On  the  other  hand,  igneous 
rocks,  usually  highly  undersaturated  and 
alkalic,  are  generally  enriched  in  the  so- 
dium melilite  component  and  are  close  to 
the  akermanite  -  sodium  melilite  join  near 
ak60ge  1  osm30  0^  Goresy  and  Yoder,  1 974). 
A  C02-rich  fluid  is  probably  involved  in 
the  upper  mantle  melting  process,  and  is 
thought  by  some  to  be  responsible  for  the 
formation  of  melilitite  magma  as  well  as 
nephelinite  and  kimberlite  magmas. 

Experimental  data  on  melilite-C02-H20 
are  available  for  the  endmember  akermanite 
(Yoder,  1968,  1973,  1975;  Huckenholz  et 
al.,  this  Report),  gehlenite  (Huckenholz 
and  Yoder,  1974;  Huckenholz,  1977; 
Hoschek,  1974)  and  for  the  akermanite  - 
gehlenite  solid  solution  series  (Huckenholz 

Mineralogisch-Petrographisches  Institut, 
Ludwig-Maximilians  Universitat,  D-8000 
Miinchen  2,  Germany. 


et  al.,  1990).  In  order  to  elucidate  the 
crystallization  behavior  of  ternary  melilites 
and  their  relationship  to  adjoining  phases, 
an  experimental  study  of  the  isothermal, 
isobaric  section  at  950°C  and  5  kbar  of  the 
akermanite  -  gehlenite  -  sodium  melilites 
was  conducted  in  the  presence  of  H2O  and 
CO2  in  both  Washington  and  Munich. 


Experimental  procedure 

Thirty  crystalline  melilite  compositions, 
prepared  by  J.F.  Schairer,  from  the 
akermanite  -  sodium  melilite  and  the 
akermanite  -  gehlenite  -  sodium  melilite 
joins  (Schairer  and  Yoder,  1964;  Schairer  et 
al.,  1967)  were  used  in  the  experiments. 
Ten  crystalline  melilite  compositions  were 
also  available  from  the  akermanite  - 
gehlenite  join  (Huckenholz  et  al.,  1990) 
and  (pure)  sodium  melilite  from  Yoder 's 
(1973)  experimental  study   on  that 
endmember  composition.  Water  was  added 
in  excess  to  the  samples  for  the  melilite- 
H2O  experiments,  whereas  oxalic  acid 
dihydrate  (H2C204»2H20)  served  as  a 
source  of  CO2 + H2O.  Oxalic  acid  dihydrate 
produces    an   initial  X(C02)    [CO2/ 
(H2O+CO2)]  of  0.5.  Higher X(C02)  were 
obtained  by  adding  calcite  +  oxalic  acid 
dihydrate  to  the  sample,  and  lower X(C02) 
were  generated  by  a  mixture  of  oxalic  acid 
dihydrate  +  water.    The  weight  ratio  of 
oxalic  acid  dihydrate  to  sample  was  about 
1:2-3.  When  carbonation  of  sample  took 
place,  that  is,  formation  of  calcite  and  sea- 
polite,  a  final  X(C02)  of  about  0.30  ±  0.03 


76 


CARNEGIE  INSTITUTION 


Table  1 1.  Liquid  compositions 


Liquid  + 

H20 

Liquid  + 

CO2  +  H2O 

Sm  100-1* 

Sm  100-2 

Sm8 

SmlOO 

Sm5 

SmlO 

Sm67 

Si02 

43.40 

42.57 

44.44 

48.01 

48.60 

50.03 

48.80 

AI2O3 

22.98 

22.08 

(23.00) 

21.30 

23.16 

20.27 

(21.70) 

Cat) 

0.03 

0.03 

0.50 

0.06 

0.51 

0.54 

0.50 

15.59 

15.65 

(15.60) 

9.30 

8.55 

9.18 

(9.10) 

Na20 

8.18 

9.25 

8.40 

11.49 

11.11 

10.02 

10.53 

totals 

90.18 

89.58 

(91.94) 

90.16 

91.93 

90.04 

(90.63) 

mel 

48.4 

53.1 

47.9 

39.6 

23.9 

29.1 

27.4 

(ak) 

- 

- 

6.6 

0.4 

3.6 

3.8 

3.4 

(ge) 

27.2 

19.5 

24.8 

7.4 

13.4 

10.7 

12.2 

(Sm) 

21.2 

33.6 

16.5 

31.8 

6.9 

14.6 

11.8 

jd 

23.2 

36.0 

26.8 

39.2 

55.5 

31.3 

44.4 

ab 

28.4 

10.9 

25.3 

21.2 

20.6 

39.6 

28.2 

X(CQ2)V 

0.0 

0.0 

0.0 

0.72 

0.63 

0.54 

0.40 

Compositions  of  glasses  from  950°C  and  5  kbar  experiments.  SmlOO- 1,  sodium  melilite 
(starting  from  glass);  Sm  100-2,  sodium  melilite  (starting  from  crystalline  sodium  melilite); 
Sm8,  akermanitel0-gehlenite20- sodium  melilite70;  Sm5,  akermanite5-sodium  melilite95; 
SmlO,  akermanite  10- sodium  melilite90;  Sm67}  akermanite67- sodium  melilite33.  Molecu- 
lar proportions  are  (ak)  Ca2MgSi207;  (ge),  Ca2Al2SiC>7;  (Sm),  NaCaAlSi207;  jd, 
NaAlSi206;  ab,  NaAlSi308-  **  Numbers  in  parentheses  are  estimates. 


Table  12.  Compositions  of  crystalline  materials 


Liquid  +  H20 

Liquid 

+  CO2  +  H2C 

> 

SmlOO 

Sm8 

SmlOO 

Sm5 

Sm64 

Sm67 

AklOO 

wo 

mel 

cc 

cc 

cc 

cc 

cc 

Si 

1.002 

1.716 

0.004 

0.019 

0.001 

0.001 

0.000 

Al 

0.002 

1.059 

0.001 

0.015 

0.000 

0.001 

0.000 

Mg 

0.001 

0.250 

0.000 

0.002 

0.003 

0.003 

0.002 

Ca 

0.995 

1.547 

0.992 

0.961 

0.996 

0.994 

0.997 

Na 

0.001 

0.432 

0.003 

0.001 
Liquid  +  C02  + 

0.001 
H20 

0.001 

0.001 

Sm5 

SmlO 

Sm20 

Sm67 

Sm64 

Sm67 

AklOO 

wo 

wo 

wo 

wo 

cpx 

cpx 

cpx 

Si 

0.991 

0.991 

1.001 

0.998 

1.918 

1.925 

1.996 

Al 

0.003 

0.007 

0.002 

0.002 

0.226 

0.190 

0.003 

Mg 

0.008 

0.024 

0.011 

0.014 

0.848 

0.876 

0.997 

Ca 

0.996 

0.972 

0.984 

0.984 

0.979 

0.989 

1.002 

Na 

0.002 

0.005 

0.001 

0.001 

0.029 

0.020 

0.002 

Compositions  of  wollastonite,  melilite,  calcite,  and  clinopyroxene  from  950°C  and  5  kbar 
experiments.  Compositions  in  cations  per  formula  units.  Bulk  compositions  as  in  Table  1 1 ; 
others  are:  Sm64,  akermanite64-sodium  melilite36;  Sm20,  akermanite20- sodium  meliliteso; 
aklOO,  akermanite  100. 


GEOPHYSICAL  LABORATORY 


77 


950"C-5  kbar 


Sm 
NaCaAISip7 


L+wo+v 


Ge 
Ca2AI2Si07 


Ak 


Sm 
NaCaAISi^ 


L+cc+V 

+mel+cc+wo+V 


Ge 
CaAI^SJaG; 


L+mel+cc+cpx+wo+V 


Ca^AgS\p7 


Fig.  46.  Ak-Ge-Sm-H20  isothermal-isobaric  sec- 
tion at  950°C  and  5  kbar.  Line  A -Bis  the  limit  of 
solid  solution  for  ternary  melilites  found  in  natural 
rocks  (El  Goresy  and  Yoder,  1974).  Abbrevia- 
tions are:  V,  fluid;  L.,  liquid;  mel,  melilite;  wo, 
wollastonite;  gar,  garnet;  and  cpx,  clinopyroxene. 
Symbols  refer  to  the  phase  assemblages  as  la- 
belled in  the  plane.  Circle  with  dot  is  the  melilite 
composition  analyzed. 

in  the  subsolidus  assemblages  resulted. 

In  hypersolidus  assemblages,  H2O  par- 
titions between  fluid  and  liquid.  For  the 
liquid,  an  X(C02)  between  0.05  and  0. 1  is 
assumed.  This  assumption  is  derived  from 
mass  balance  calculations  conducted  on  a 
glass  quenched  from  the  sodium  melilite 
endmember  composition  in  which  calcite 
is  the  only  liquidus  phase,  and  from  CO2  + 
H2O  solubility  in  albite  melts  (Kadik  and 
Eggler,  1974).  The  CO2  partitions  pre- 
dominantly between  fluid  and  calcite  + 
scapolite.  The  "final"  X(C02)  that  results 
from  X(C02)(fluid)  +  X(C02)(liquid)  can 
be  calculated,  and  is  about  0.3  in  most  of 
the  hypersolidus  experiments. 

Experiments  were  carried  out  by  means 
of  two  internally  heated  gas-media  appara- 
tus in  Washington  and  in  Munich.  The  run 
duration  was  24  hours  in  each  case. 
Reversibility  of  experiments  has  not  as  yet 


Fig.  47.  Ak-Ge-Sm-H20-CC>2  isothermal-iso- 
baric section  at  950°C  and  5  kbar  at  an  X  (CO2)  of 
about  0.3.  Abbreviations  as  Fig.  46;  other  is  scap, 
scapolite.  Symbols  refer  to  the  phase  assemblages 
as  labelled  on  the  diagram.  Square  with  dot  is  the 
phase  assemblage  at  the  beginning  of  melting. 
Line  A  -  B  is  the  position  of  the  X(CC>2)  versus 
composition  plot  of  data  for  Fig.  48B. 


been  demonstrated  for  this  preliminary  re- 
port. The  run  products  were  examined  by 
optical  methods,  x-ray  powder  diffraction, 
and  for  10  compositions,  by  microprobe 
analyses.  Composition  of  glasses,  calcite, 
wollastonite,  clinopyroxene,  and  melilite 
are  given  in  Tables  11  and  12. 


Experimental  Results 

The  experimental  data  obtained  are 
presented  in  two  isothermal-isobaric  (950C 
-  5  kbar)  sections  of  the  ak  (akermanite)-ge 
(gehlenite)-Sm  (sodium  melilite)  join  with 
H2O  and  with  CO2  +  H2O  as  fluids  (Figs. 
46  and  47).  The  two  ternary  joins  are  each 
part  of  the  cc  (calcite)-di  (diopside)-CaTs 
(Ca-Tschermak's  component)-jd  (jadeite) 
tetrahedron,  which  in  turn  is  a  four-compo- 
nent volume  of  the  Na20-CaO-MgO- 


78 


CARNEGIE  INSTITUTION 


950"C-5  kbar 


scap  (mei)  4cc*cpx  tsp.V 


B 

■                           T                          T                       -1 1 1- 

1 

|  950'C-5  kbar  | 

- 

cpx.cc 
+wo+V 

■ 

scap+cc* 
sp<?)+c<x+V 

scap+cc+cpx*sp(?)+V 

0*5 

* 

" 

mel+cc           ^ 

\ 

\                         mal+cpx+cc+V 

\        ■ 

■ 

- 

\    . 

md+V 

■ 

10            20 

30 

40            50 

60 

70 

80 

90             A 

/Sm15) 

mol% 

(Ak85/Sm15) 

Al203-Si02-C02-H20  system.  Phases 
determined  are  fluid  (V),  liquid  (L)  calcite 
(cc,CaC03),  wollastonite  (wo,  CaSi03), 
clinopyroxene  [cpx, 

(Ca,Na)(Mg,Al)(Si,Al)206],  garnet  [gar, 
(Ca,Mg)3Al2Si30i2L  scapolite  [scap, 
Ca3(Ca,Na)iAl5-6Si6-7024(C03)]  and 
melilite  [mel,  (Ca,Na)2(Mg,Al) 
(Si,Al)207].  Compositions  of  calcite, 
clinopyroxene,  and  melilite  plot  in  the  cc- 
di-CaTs-jd  system,  whereas  wollastonite, 
garnet,  scapolite,  and  the  liquid  lie  outside. 
Chemographical  correlation  of  spatial  phase 
assemblages  from  the  multicomponent  sys- 
tem Na20-CaO-MgO-Al203-Si02-C02- 
H2O,  however,  cannot  always  be  appropri- 
ately illustrated. 

The  isothermal-isobaric  section  of  Fig. 
46  displays  the  phase  assemblages  encoun- 
tered in  experiments  with  H2O  as  the  fluid 
phase.  The  assemblages  are  outlined  by 
dashed  curves  that  are  thought  to  be  inter- 
faces of  multiphase  volumes  intersecting 
the  pseudo-ternary  plane.  The  maximum 
extent  of  melilite  solid  solution,  equivalent 


V?: 


950-C-5kbar 


,     V+cryilaU+LxMd 


cpx 

scap 


± 

Lcfjd.c/yilals  ±  cak.«e -' 


Y  ■ '  1H9  *ec  «cpx  jx  .3  - 

me)(1)  .v 


o.ot 

SM 


70 


60 


SO 
mol% 


Fig.  48.  Melilite  composition  versus  AXCO2)  .A. 
Melilite  composition  versus  X(C02)  plot  of  data 
along  the  join  akermanite  -  gehlenite.  Abbrevia- 
tions as  in  Figs.  46  and  47,  others  are  mei,  meionite; 
cor,  corundum;  sp,  spinel.  Symbols  refer  to  the 
phase  assemblages  as  labeled  on  the  diagram.  B. 
Melilite  composition  versus  X(CC>2)  plot  of  data 
along  line  A  (Ak85/Sml5)  -B  (Ge  85/Sm  15)  of 
Fig.  47.  Melilite  composition  from  above  and 
below  that  line  are  projected.  Abbreviations  as 
Figs.  46  and  47;  symbols  refer  to  the  phase  assem- 
blages labelled.  C.  Melilite  composition  versus 
X(C02)  plot  of  data  along  the  join  akermanite  - 
sodium  melilite.  Abbreviations  as  in  Figs.  46  and 
47.  Mel  (2)  lies  off  the  plane  in  the  gehlenite  -  rich 
portion  of  Fig.  47.  Vertical  solid  lines  with  dots 
are  tie  lines  between  fluid  and  liquid  in  composi- 
tions along  akermanite  -  sodium  melilite.  Sym- 
bols in  the  subsolidus  portion  refer  to  phase  assem- 
blages as  labeled  on  the  diagram. 


GEOPHYSICAL  LABORATORY 


79 


to  the  H20-saturated  solidus,  is  bound  by 
akermanite58      sodium      melilite42; 
akermanite26gehlenite3 1  sodium  melilite43 
(microprobe  data,  Table  12;  circle  with 
dot);  and  by  gehleniteyo  sodium  melilite3o 
compositions.  The  H20-saturated  melilite 
solidus  is  close  to  the  extent  of  melilite 
solid  solution  found  for  natural  melilites  by 
El  Goresy  and  Yoder  (1974).  Solid  phases 
in  the  hypersolidus  portion  of  the  plane 
coexist  with  a  water-saturated  liquid. 
Glasses  quenched  from  sodium  melilite  ioo 
as  well  as  akermaniteio  gehlenite20  so- 
dium melilite70  were  analyzed  by  electron 
microprobe.    Their  compositions  (Table 
11)  contain  the  melilite  components  (=  ak  + 
ge  +  sm)  on  the  order  of  about  50  mol  %  but 
also  jadeite  (NaAlSi206)  and  albite 
(NaAlSi308).    The  total  of  constituents 
analyzed  of  the  glasses  run  about  89-91%; 
the  remainder  is  believed  to  be  H2O  dis- 
solved in  the  melt.    The  liquidus-phase 
wollastonite  was  also  analyzed  (Table  12), 
and  is  very  close  to  wollastonite  with  minor 
solid  solution,  if  any.   No  nepheline  was 
detected  in  the  run  products,  neither  by 
optical,  x-ray  analysis,  nor  microprobe  in- 
spection. 

The  isothermal-isobaric  melilite  sec- 
tion at  950°C  and  5  kbar  with  CO2  +  H2O 
as  a  fluid  is  depicted  in  Fig.  48.  For  simplic- 
ity, the  ternary  plane  is  averaged  forX(C02) 
of  about  0.3.  The  bounding  melilite  solid 
solution  on  the  CO2 + H20-saturated  solidus 
was  assumed  to  be  akermanite20- 
gehlenite50-sodium  melilite30.  The  appar- 
ent boundaries  of  the  phase  assemblages 
mapped  in  the  plane  tend  to  converge  about 
this  composition. 

Further  indication  of  this  special  melilite 


composition  can  be  deduced  from  the  iso- 
thermal-isobaric, X(C02)  vs.  melilite  com- 
position plot.  Figure  48A  displays  such  a 
relationship  for  akermanite-gehlenite.  The 
solid  solution  is  bound  by  decomposition 
of  akermanite  +  CO2  to  diopside  +  calcite 
at  0.12±0.02  and  by  gehlenite  +  CO2  to 
meionite  +  calcite  +  corundum  at  X(C02) 
0.30+0.02  (Huckenholz  et  al.,  1990; 
Huckenholz  and  Seiberl,  1990),  respec- 
tively. The  breakdown  assemblages  of 
melilite  +  scapolite  -1-  calcite  +  corundum  + 
spinel  (?)  in  the  gehlenite-rich  portion  and 
that  of  melilite  +  clinopyroxene  +  calcite  in 
the  akermanite-rich  portion  indicate  a  maxi- 
mum melilite  solid  solution  of  about 
akermanite  1 5 -gehlenite85,  which  appears 
to  decompose  at  about  X(C02)  =  0.33  to 
scapolite  +  calcite  +  corundum  +  spinel  (?). 

A  C02-dependent  compositional  maxi- 
mum is  also  indicated  when  the  relations 
are  considered  along  line  A  -B  on  the 
akermanite-gehlenite-sodium  melilite  plane 
parallel  to  akermanite-gehlenite  at  about 
sodium  melilite  15.  In  the  section  displayed 
in  Fig.  48B,  the  melilite  -1-  CO2  stability  has 
increased  to  about  0.45  AXCO2) .  No  stable 
melilite  was  found  in  the  subsolidus  assem- 
blage of  scapolite  +  calcite  +  clinopyroxene 
+  spinel  (?)  at  0.5,  and  above  the  scapolite 
is  a  solid  solution  between  meionite 
[Ca4Al6Si6024(C03)]  and  the  (theoreti- 
cal) carbonate-marialite  endmember 
[Na3CaAl3Si9024(C03)].  At  950°C  and  5 
kbar  its  composition  in  (scapolite  +  calcite) 
-  bearing  assemblages  is  restricted  to  an 
equivalent  an-contentof  0.75  (Huckenholz 
and  Seiberl,  1990). 

Phase    relations    along    the    join 
akermanite-sodium  melilite  as  a  function 


80 


CARNEGIE  INSTITUTION 


of  X(C02)    are  shown  in  Fig.48C.  The 
extent  of  solid  solution  toward  sodium 
melilite  is  limited  to  akermanite70-sodium 
melilite30  and  an  X(C02)  of  0. 1 2,  which  is 
at  the  breakdown  of  akermanite  +  CO2  to 
diopside  +  calcite.  Melilites  labelled  mel 
(2)  occur  within  the  phase  assemblages 
generated  by  the  breakdown  of  akermanite 
+  CO2,  but  do  not  lie  in  the  pseudo-binary 
akermanite-sodium  melilite  plane.   Their 
composition  is  located  in  the  gehlenite-rich 
portion  of  the  akermanite-gehlenite-sodium 
melilite  join  as  shown  in  Fig.  47,  with 
X(C02)  up  to  about  0.45.  From  the  bound- 
ing melilite  solid  solution  of  the  fluid- 
saturated  solidus  toward  the  sodium  melilite 
endmember,  hypersolidus  phase  assem- 
blages   of    V+mel+cc+cpx+scap+L, 
V+mel+cc+cpx+wo+L,  V+mel+cc+wo+L, 
V+mel+cc+L  and  V+cc+L  are  traversed  by 
the  pseudo-binary  join.    Liquid  with  an 
assumed  X(C02)(L)  of  0.07  coexists  with 
fluid  having  X(C02)(V)  of  0.45  in  the 
akermanite-rich  portion  of  the  join.    To- 
ward the  sodium  melilite  endmember  com- 
position, the  amount  of  the  liquid  increases 
thereby  consuming  increasing  amounts  of 
H2O,  which  in  turn  results  in  raising  the 
X(C02)  of  the  fluid. 

Quenched  glasses  from  sodium 
melilite  100,  akermanite5 -sodium  melilite95, 
akermaniteio-sodium  melilite^,  and 
akermanite67-gehlenite33  bulk  composition 
have  been  analyzed  by  microprobe  (Table 
11).  They  exhibit  (calculated  molecular)  jd 
(NaAlSi2C>6)  and  ab  (NaAlSi308)  compo- 
nents in  addition  to  a  large  amount  of 
ternary  melilite  (ak+ge+sm).  In  contrast, 
glasses  quenched  in  the  presence  of  a  (pure) 


H2O  fluid  are  lower  in  the  jd  and  ab  com- 
ponents but  contain  a  ternary  mel-compo- 
nent  >  47%.  Analyzed  calcite  (Table  12) 
contains  minor  amounts  of  MgO  and  Na20; 
wollastonite  displays  minor  solid  solution 
toward  the  sodium  melilite  (<1%)  and 
akermanite  (2-4%)  components. 
Clinopyroxenes  were  analyzed  from  the 
akermanite64-sodium  melilite36, 
akermanite67-sodium  melilite33,  and 
akermaniteioo  bulk  compositions.  Ex- 
pressed as  endmembers,  they  reduce  to 
diopside87-CaTsio  jadeite3,  diopside89- 
CaTs9  jadeite2,  and  diopside99.7-CaTs<o.i 
jadeite<o.2,  respectively. 


Reference 

El  Goresy,  A.,  and  H.  S.  Yoder,  Jr.,  Natural  and 
synthetic  melilite  compositions.  Carnegie 
Instn.  Washington,  Year  Book,  73,  359-371, 
1974. 

Hoschek,  G.,  Gehlenite  stability  in  the  system 
CaO-Al203-Si02-H20-C02-  Contr.  Min. 
Petr.,  47,  245-254,  1974. 

Huckenholz,  H.  G.,  Gehlenite  stability  relations  in 
the  join  Ca2Al2SiC>7  -  H2O  up  to  10  kbar. 
NJb.  Miner,  Abh.,  130,  169-186,  1977. 

Huckenholz,  H.  G.,  and  H.  S.  Yoder,  Jr.,  The 
gehlenite-H20  and  \vollastonite-H2O  systems. 
Carnegie  Instn.  Washington,  Year  Book,  73, 
440-443,  1974. 

Huckenholz,  H.G.,  A.  Wassermann,  and  K.  T. 
Fehr,  Stability  and  phase  relations  of  gehlenite- 
akermanite  solid  solutions  in  the  presence  of  a 
H20-C02-fluid.  International  Symposium  of 
Experimental  Mineralogy,  Petrology  and  Geo- 
chemistry, Edinburgh,  UK,  p.  17,  terra  ab- 
stracts, 2,  1990. 

Huckenholz,  H.  G.,  and  W.  Seiberl,  Stability  and 
phase  relations  of  carbonate  scapolite  solid 
solutions  under  the  PT-regime  of  the  deeper 
crust.  Third  International  Symposium  of  Ex- 
perimental Mineralogy,  Petrology  and  Geo- 
chemistry, Edinburgh,  UK,.  P.  17;  terra  ab- 
stracts 2  1990. 

Kadik,  A.  A.,  and  D.  H.  Eggler,Melt-vapor  rela- 
tions on  the  join  NaAlSi308-H20-C02: 
Carnegie  Instn.  Washington,  Year  Book,  74, 
479-484,  1974. 


GEOPHYSICAL  LABORATORY 


81 


Schairer,  J.  R,  and  H.  S.  Yoder,  Jr.,  The  join 
akermanite  (Ca2MgSi207)  -  soda  melilite 
(NaCaAlSi207)  .  Carnegie  Instn.  Washing- 
ton, Year  Book,  63,  89-90,  1964. 

Schairer,  J.F.,  H.  S.  Yoder,  Jr.,  and  C.  E.  Tilley, 
The  high-termperature  behavior  of  synthetic 
melilites  in  the  join  gehlenite-soda  melilite- 
akermanite,  Carnegie  Instn.  Washington,  Year 
Book,  65,  217-226  1967. 

Yoder,  H.  S.,  Jr.,  Akermanite  and  related  melilite- 
bearing  assemblages.  Carnegie  Instn.  Wash- 
ington, Year  Book,  66,  p47 1-477,  1968. 

Yoder,  H.  S.  Jr.,  Melilite  stability  andparagenesis. 
Fortschr.  Mineral,  v.50,  140-173,  1973. 

Yoder,  H.S.,  Jr.,  Relationship  of  melilite- bearing 
rocks  to  kimberlite:  a  preliminary  report  on  the 
system  akermanite-C02.  Proc.  Internat. 
kimberlite  Conf.,  Cape  Town.  Phys.  Chem. 
Earth,  9,  883-894,  1975. 


Merwinite  Stability  and 

High-Temperature  Phase  Relations 

in  the  Presence  of  CO2  +  H2O. 

H.  G.  Huckenholz,  H.  S.  Yoder,  Jr.,  and 
W.  Seiberl 


monticellite +melilite-bearing  assemblages 
in  high-temperature  calc-silicate  rocks  oc- 
curring as  inclusions  in  pyroxenites  from 
the  critical  zone  of  the  eastern  Bushveld 
Complex. 

Phase  equilibria  studies  on  the  join 
CaMgSi206-CaC03-C02  of  the  CaO- 
MgO-Si02-C02  system  restrict  the 
merwinite  +  C02  stability  to  high-tem- 
perature but  low-pressure  conditions.  The 
reaction  akermanite  +  calcite  <=>  merwinite 
+  CO2  (step  1 1  of  the  decarbonation  series 
of  Bowen,  1940)  was  studied  by 
Shmulovich  (1969)  and  Walter  (1963a,b, 
1965)  at  low  pressure  but  high  tempera- 
ture. Merwinite  +  CO2  crystallizes  from 
akermanite  +  calcite  assemblages  at  tem- 
peratures of  <  1065°C  and  pressures  of  < 
0.5  kbar.  Merwinite  +  CO2,  however,  did 
not  crystallize  from  diopside  +  2  calcite 
assemblages  between  950°C  and  975°C  at 
1  kbar  (Yoder,  1975). 


Merwinite  [Ca3Mg(Si04)2]  was  discov- 
ered and  named  by  Larsen  and  Foshag 
(1921)  from  high-grade  metamorphosed 
carbonaceous  rocks  at  Crestmore  near  Riv- 
erside, California.  At  Crestmore  (Burnham, 
1959;  Walter,  1965)  and  other  localities 
(e.g.,  Scawt  Hill,  Northern  Ireland,  Tilley, 
1929;  Ardnamurchan,  western  Scotland, 
Agrell,  1965;  Christmas  Mountains,  Big 
Bend  region,  Texas,  Joesten,  1974), 
merwinite  is  mainly  associated  with  cal- 
cite, spurrite,  monticellite,  melilite,  and 
also  with  larnite.  Recently,  Wallmach  et  al. 
(1989)    described    merwinite    from 

Mineralogisch-Petrographisches  Institut, 
Ludwig-Maximilians  Universitat,  D-8000 
Munchen  2,  Germany. 


Experimental  Methods 

In  order  to  elucidate  the  stability  of 
merwinite  in  the  presence  of  CO2  +  H2O  at 
pressures  >  0.5  kbar,  an  experimental  study 
of  merwinite  and  its  high-temperature  phase 
relations  was  conducted.  Stability  and  phase 
relations  of  merwinite  with  akermanite, 
diopside,  calcite,  liquid,  and  fluid  were 
studied  for  pure  CO2  in  the  1-10  kbar 
pressure  range  at  temperatures  between 
900°  and  1450°C.  In  addition,  isobaric 
temperature  versus  CO2  +  H2O  relations 
were  investigated  at  1  and  3  kbar  and  at 
temperatures  between  700°  and  1 200°C. 

Mixtures  of  crystalline  materials  were 
used  in  all  cases  for  the  experiments.  They 
consisted  of: 


82 


CARNEGIE  INSTITUTION 


03 
_Q 
J*: 

<D 

13 
CO 
CO 

0 


12 
11 

ioh 

9 
8 

7 
6 
5 
4 
3 
2 
1 


0 

800 


0  o  /•/•/•    •/«> 


900 


1000  1100  1200 

Temperature,  °C 


1300 


1400 


Fig.  49.  Pressure-temperature  diagram  for  merwinite  +  CO2,  akermanite  +  calcite,  and  diopside  +  calcite 
compositions.  Heavy  lines  are  univariant  reaction  curves;  light  lines  are  restricted  reaction  curves;  short 
dashed  curve  is  the  compositional  singularity  for  di  +  2  cc  =  L;  I\,  I2,  invariant  points;  Si,  52  singular 
points.  Abbreviations  for  phases  are:  mer,  merwinite  (Ca3MgSi20g);  ak,  akermanite  (Ca2MgSi2(>7); 
di,  diopside  (CaMgSi2C>6);  cc,  calcite  (CaCC>3);  L,  Liquid.  Symbols:  Solid  triangles  (1)  diopside  + 
calcite  =  akermanite  +  CO2  (Shmulovich,  1969;  Walter,  1963);  solid  triangles  (2)  akermanite  +  calcite 
=  merwinite  +  CO2  (Shmulovich,  1969;  Walter,  1963);  open  hexagons,  di  +  cc  from  mixture  B,  C,  and 
Y  (for  composition  see  text);  open  squares,  akermanite  +  CO2  from  mixture  B,  C,  and  Y  as  well  as 
Yoder's  data  (1975);  open  diamonds,  merwinite  +  CO2  from  mixture  B,  C,  and  Y;  solid  diamonds,  mer 
+  L  +  CO2  from  mixtures  B,  C,  and  Y;  solid  squares,  ak  -1-  L  and  ak  +  L  +  CO2  from  mixture  B,  C,  and 
Y;  solid  pentagon  and  solid  circle  are  the  2  kbar  run  data  on  di  +  ak  +  CO2  =  L  (Yoder,  1975);  solid 
hexagons,  di  +  L  and  Di  +  L  +  CO2  from  mixture  C,  Y,  and  Yoder's  1975  run  data;  open  circles,  Liquid 
on  diopside  +  cc  -1-  CO2  compositions,  mixture  B,  C,  and  Yoder's  (1975)  run  data. 


(1).  merwinite  (crystallized  at  1200°C,  1 
atm),  mixture  A; 

(2).  akermanite  (crystallized  at  1050°C  at  1 
atm)  +  calcite  (Baker  Chemical  Com- 
pany, grade  C.R.)  equivalent  (on  a  mo- 
lar basis)  to  merwinite  +  CO2;  mixture 
B; 

(3).  natural  diopside  (Twin  Lakes,  Califor- 
nia; Smith,  1 966)  +  2  calcite  equivalent 


to  akermanite  +  calcite  +  CO2  or 
merwinite  -1-  2  CO2,  mixture  C;  and 

(4).  natural  2  wollastonite  (Willsboro,  N.  Y.) 
+  natural  dolomite  (Thornwood,  N. Y.) 
equivalent  to  akermanite  +  calcite  + 
CO2,  or  diopside  +  2  calcite,  or 
merwinite  +  2  CO2,  mixture  Y. 
Other  experiments  were  carried  out  on 

natural  rock  inclusions  from  the  upper  zone 


GEOPHYSICAL  LABORATORY 


83 


of  the  eastern  Bushveld  Complex  (locality 
Luipershoek,  Joubert,  1976).  Rocks  con- 
taining akermanite  +  diopside  (± 
monticellite),  Ji,  and  monticellite  + 
akermanite  (±  diopside),  J6,  were  collected 
by  H.  G.  Huckenholz  (October,  1990).  The 
synthetic  assemblages  merwinite  + 
monticellite  +  akermanite60-gehlenite40 
(mixture  E)  and  akermanite  +  monticellite 
+  calcite  (mixture  D)  were  also  studied  at  1 
and  3  kbar  in  the  presence  of  CO2  +  H2O. 
Experimental  data  were  obtained  by  means 
of  internally-heated,  gas-media  apparatus 
in  both  Washington,  D.  C,  and  Munich. 
Reversibility  of  the  experiments  was  en- 
sured by  the  direction  of  reaction  of  the 
different  crystalline  mixtures  listed  above. 


Merwinite  Relations  with  CO2 

Stability  of  merwinite  -1-  CO2  and 
merwinite  phase  relations  with  akermanite, 
diopside,  calcite,  liquid,  and  CO2  are  dis- 
played in  the  temperature  versus  pressure 
plot  in  Fig.  49.  With  the  data  of  Shmulovich 
(1969)  at  0.5  kbar  and  below  (see  also 
Walter,  1963a,b),  the  reaction  akermanite  + 
calcite  =  merwinite  +  CO2  increases  in 
temperature  from  1015°C  at  0.5  kbar  up  to 
1 1 82°C  at  1 .3  ±  0.2  kbar,  and  results  in  the 
invariant  assemblage  (h)  of  merwinite  + 
akermanite  +  diopside  +  calcite  +  liquid  + 
CO2  in  equilibrium.  Four  other  univariant 
curves, 

[ak]  merwinite  +  calcite  +  CO2  <=>  liquid, 

[mer]  akermanite  +  calcite  +  CO2 
<=>  liquid, 

[cc]  merwinite  +  CO2 

<=>  akermanite  -1-  liquid,  and 


[CO2]  liquid  +  merwinite 

<=>  akermanite  +  calcite, 

meet  at  that  invariant  point  I2.  The  reaction 
[ak]  was  bracketed  between  1150°C  and 
1200°C  at  P  =  1  kbar.  Its  position  at  about 
1180°C  is  fixed  due  to  the  (positive)  slope 
of  reaction  [mer]  and  by  the  run  at  1 200°C 
and  3  kbar  in  particular,  which  is  just  slightly 
above  the  solidus  with  a  phase  assemblage 
of  akermanite  +  calcite  +  liquid.  At  2  kbar 
and  temperatures  between  1200°  and 
1450°C,  merwinite +  CO2  does  not  crystal- 
lize from  akermanite  +  calcite  nor  from 
diopside  +  2  calcite  mixtures.  Thus,  the 
curves  for  the  two  univariant  reactions  of 
[cc]  and  [CO2]  must  pass  between  1.3  and 
2  kbar.  At  reaction  [cc],  merwinite  +  CO2 
tie  lines  are  interrupted  by  those  of 
akermanite  +  liquid  but  with  merwinite 
remaining  in  the  C02-absent  region  of  the 
merwinite  +  calcite  +  liquid  and  merwinite 
+  akermanite  +  liquid  assemblages. 

Because  of  chemographic  constraints, 
reaction  [CO2]  must  occur  on  the  high- 
pressure  side  of  reaction  [cc]  and  between 
1.3  and  2  kbar  as  well.  In  that  pressure 
range,  the  melting  curve  of  calcite  (Yoder, 
1 973)  intersects  the  reaction  [CO2]  at  about 
1350°C.  Calcite  in  the  assemblage  will 
melt  and  the  restricted  assemblage  of 
merwinite  +  akermanite  +  liquid  evolves 
from  S\.  At  higher  temperatures,  the  diop- 
side +  akermanite  +  CO2  solidus  and  the 
diopside  +  CO2  solidus  appear  in  the  diop- 
side +  akermanite  +  CO2  portion  of  the 
diopside  -1-  calcite  +  CO2  system.  The 
temperature  of  the  akermanite  +  diopside  + 
CO2  solidus  can  be  deduced  from  the  run 
data  of  Yoder  (1975)  on  diopside  +  calcite 
compositions  obtained  at  2  kbar. 
Akermanite  +  diopside  +  CO2  will  melt 
slightly  above  1400°C,  and  diopside  +  CO2 


84 


CARNEGIE  INSTITUTION 


(Rosenhauer  and  Eggler,  1975)  at  1415°C 
as  well. 

In  the  temperature  and  pressure  range 
studied,  merwinite  and  diopside  do  not 
coexist  in  the  presence  of  C02.  They  are 
separated  by  akermanite  +  calcite,  or  by 
akermanite  +  CO2  tie  lines.  Below  900°C 
there  is  only  a  very  narrow  CO2  pressure 
range  of  about  200  to  300  bar  where 
akermanite  +  CO2  is  formed  from  diopside 
+  calcite.    With  increasing  temperature, 
akermanite  +  CO2  is  stable  up  to  about  5 
kbar.    The  reaction  diopside  +  calcite  = 
akermanite  +  CO2  (step  8  of  Bowen's  de- 
carbonation  series)  was  studied  at  low  pres- 
sures by  Walter  (1963)  and  up  to  6  kbar  by 
Yoder  (1973),  who  exclusively  used  a  crys- 
talline mixture  of  (natural)  diopside  -1-  1 
calcite.  The  slope  of  the  reaction  curve  of 
Yoder  (1973),  drawn  at  the  first  appearance 
of  akermanite,  between  1  and  5.75  kbar  is 
about  58°C/kbar.  The  reaction  takes  place 
through  a  range  of  temperatures  50°-70°C 
wide  at  a  given  pressure,  presumably  be- 
cause of  possible  solid  solution  of  merwinite 
in  akermanite,  diopside  in  merwinite 
(Schairer  et  aiy  1967;  Yoder,  1973),  and 
minor  substitution  of  Ca  by  Mg  in  calcite. 
The  akermanite  +  diopside  +  calcite  region 
was  not  observed  in  the  present  study  when 
akermanite  +  calcite  and  diopside  +  2  cal- 
cite compositions  were  used.   The  newly 
crystallized  akermanite,  however,  contains 
tiny  inclusions  of  (relict)  diopside  that  are 
separated  from  calcite  by  the  akermanite 
host.  The  diopside  +  calcite  <=>  akermanite 
+  CO2  reaction  curve,  now  bracketed  by 
means  of  akermanite  +  calcite  and  diopside 
+  2  calcite    compositions,  has  a  revised 
slope  of  about  62°C/kbar.  It  terminates  at 
the  invariant  point  I\,  which  was  found  to 
be  located  at5.4±0.2kbar  and  1215°±5°C 
with  akermanite  +  diopside  +  calcite  + 


liquid  +  CO2  in  equilibrium. 

The  invariant  point  I\  is  the  locus  of  four 
other  reactions  occurring  clockwise: 

[CO2]  akermanite  +  calcite  +  diopside 
<=>  liquid, 

[ak]  calcite  +  diopside 

<=>  liquid  +  CO2, 

[cc]  diopside  +  liquid 

<=>  akermanite  +  CO2,  and 

[di]  akermanite  -1-  calcite  +  CO2  <=>  liquid. 

Separation  of  reaction  [C02]  from  reaction 
[ak]  was  not  possible  because  the  liquid 
field  may  be  located  very  close  to  or  even 
on  the  diopside  +  calcite  join.  Above  the 
temperature  of  reaction  [ak],  diopside  + 
liquid  +  C02  will  reach  the  diopside  +  2 
calcite  composition  (=  compositional  sin- 
gularity), and  any  further  increase  in  tem- 
perature will  move  the  liquid  -1-  C02  tie 
lines  along  diopside  +  calcite  toward  the 
join  akermanite  -1-  CO2,  that  is,  becoming 
compositionally  equivalent  to  diopside  +  1 
calcite.  That  particular  composition  is 
reached  in  S2  at  13 10°C  and  about  5.3  kbar 
located  on  reaction  [cc],  diopside  +  calcite 
=  akermanite  +  CO2,  from  which  the  two 
restricted  reactions  diopside  +  calcite  = 
liquid  (higher  pressure  limb)  and  liquid  = 
akermanite  -1-  CO2  (lower  pressure  limb) 
will  occur. 


Merwinite  Relations  With  CO2  and  H2O 

Experiments  on  the  stability  of 
merwinite  in  the  presence  of  C02  +  H2O 
were  conducted  at  1  and  3  kbar  (Figs.  50A 
and  50B).  The  merwinite  +  V  stability  field 


GEOPHYSICAL  LABORATORY 


85 


1300 


1200 


1100 


O 

°-  1000 
0 

|  900 
CD 

Q. 

E 

i®  800 


700 


600 


P=1  kbar 


(1)KUSHIRO&YODER,1964 

(2)  WALTER,  1965 

(3)YODER,1973 

(4)  HUCKENHOLZ  etal.,  1990 


0.0        0.1        0.2        0.3       0.4        0.5        0.6       0.7       0.8        0.9       1.0 

XcOo 


1300 


_  .  -  995' 


1195* 
1170' 


1090' 
1065' 


Fig.  50a  and  50b.  Isobaric  temperature  -X(<X>2)  [X((X>2)  =  CO2ACO2  +  H2O)]  diagram  for  1  kbar  (Fig. 
50A)  and  3  kbar  (Fig.  50B).  Abbreviations  for  phases  are:  mo,  monticellite  (CaMgSi04);  per,  periclase 
(MgO);  fo,  forsterite  (Mg2Si04)  spur,  spurrite  [Ca5Si208(C03)];  wo,  wollastonite  (CaSi03);  geh, 
gehlenite  (Ca2Al2Si07);  gro,  grossular  (Ca3Al2Si30i2);  cor,  corundum  (AI2O3);  mei,  meionite 
(Ca3  Al6Si6024*CaC03);  V,  fluid;  other  abbreviations  and  symbols  as  in  Fig.  49.  Letters  on  half-shaded 
symbols  or  symbols  with  dots  refer  to  experiments  on  mixtures  F,  D,  E,  Jj,  and  J6-  The  solidus  of  these 
compositions  (except  for  E)  refers  to  ak  +  di  +  CO2  (above  1 1 50°C).  Light  curves  and  light  dashed  curves 
are  decarbonation  reactions  (extrapolated)  from  Walter  (1963a.b;  1965)  h  to/6  are  not  fully  illustrated. 
The  geh  +  V  reaction  is  from  Huckenholz  et  al.  (1990). 


86 


CARNEGIE  INSTITUTION 


decreases  with  increasing  pressure  when 
the  C02  fluid  is  diluted  with  H2O.  At 
AXCO2)  =  1.00,  akermanite  +  calcite  = 
merwinite  +  V  are  restricted  to  a  tempera- 
ture of  1 1 20°C  at  1  kbar  and  1 1 82°C  at  1 .3 
kbar  (isobaric  invariant  point  I2).  From  I2 
the  reaction  akermanite  +  calcite  <=> 
merwinite + V  shifts  to  lower  AXCO2)  when 
the  pressure  increases  (light  dashed  curves 
in  Fig.  50A  and  50B).  At  810°C  and  at  1 
kbar  the  reaction  spurrite  -1-  monticellite  <=> 
merwinite  +  calcite  (Walter,  1965),  well 
displayed  at  Crestmore,  California 
(Burnham,  1959),  must  intersect  with 
akermanite  +  calcite  <=>  merwinite  +  V  at 
about  X(C02)  =  0.10  in  Fig.  50A  The 
resulting  invariant  point  labelled  I5  be- 
comes the  locus  of  the  reaction  of  spurrite 
+  akermanite  +  monticellite  <=>  merwinite 
+  V  and  akermanite  +  calcite  <=>  spurrite  + 
monticellite  +  V  (not  shown  in  Fig.  50A). 
With  decreasing  temperatures  spurrite  + 
akermanite  +  monticellite  <=>  merwinite  + 
V  intersects  monticellite  +  wollastonite  <=> 
akermanite  at  a  pressure  of  1  kbar  and  a 
temperature  as  low  as  700°C  and  atX(C02) 
=  0.07  (isobaric  invariant  point  Ie). 

Merwinite  +  calcite  +  V  will  melt  at  1 
kbar  over  almost  the  entire  range  of  X(C02) 
displaying  merwinite  +  calcite  +  liquid  as 
well  as  the  merwinite  -1-  liquid  +  V  assem- 
blage. Increasing  pressure  shifts  the 
merwinite  +  calcite  +  V  solidus  toward 
lower  X(C02)  and  lower  temperatures  as 
well.  At  3  kbar,  merwinite  +  calcite + V  will 
melt  atX(C02)  <  0.15  and  temperatures  < 
1150°C.  At  X(C02)  >  0.15,  akermanite  + 
calcite  +  V  forms  the  subsolidus  assem- 
blage, which  melts  between  1150°  and 
1 200°C  to  akermanite  +  calcite + liquid  and 
akermanite  +  liquid  +  V  assemblages.  The 
X(C02)  of  the  liquid  coexistent  with  fluid 


was  not  determined  during  the  course  of  the 
present  study. 

Melting  experiments  on  silicate  +  H2O 
+  CO2  systems  (Mysen,  1975;  Rosenhauer 
and  Eggler,  1975;  Kadik  and  Eggler,  1975) 
demonstrate,  however,  the  preference  of 
H2O  solubility  over  CO2  in  silicate  melts 
and,  therefore,  X(C02)  of  the  liquid  of 
about  0.05  was  assumed  for  the  tempera- 
ture and  pressure  conditions  studied.  The 
diopside  +  calcite  =  akermanite  +  CO2 
reaction  was  determined  for  3  kbar  at 
X(C02)  of  1 .00, 0.50  and  0.05  (Huckenholz 
et  ai,  1990;  this  study)  and  found  to  be 
1065°,  1010°,  and 740°C, respectively.  Cal- 
culations for  X(C02)  (Helgeson  et  ai, 
1978;  Holloway;  1977)  at  0.95,  0.10,  and 
0.05  yielded  temperatures  of  1059°,  846°, 
and  803  °C.  Experimental  data  for  1  kbar 
are  available  only  for  X(C02)  at  1 .00  and 
950°C  and  for  X(C02)  at  0.5  and  895°C 
(Huckenholz  etal.,  1990;  this  study).  They 
were  calculated  for  X(C02)  =  0.10  and 
0.20,  and  found  to  be  about  735°  and  770°C, 
respectively. 

The  diopside  +  calcite  <=>  akermanite  + 
V  reaction  is  intersected  by  the  fluid-  and 
calcite-absent  reaction  of  diopside  + 
monticellite  =  forsterite  +  akermanite  on 
the  basis  of  the  experimental  data  of  Kushiro 
and  Yoder  (1964)  and  of  Yoder  (1973).  The 
intersection  will  occur  at  1  kbar  at  880°C 
and  about  X(C02)  =  0.45  and  at  3  kbar  at 
about  920°C  and  X(C02)  =  0.20  ±  0.02, 
resulting  in  the  isobaric  invariant  point 
labeled  I4.  The  validity  of  the  diopside  + 
monticellite  =  forsterite  +  akermanite  reac- 
tion is  questioned  by  Helgeson  etal.  ( 1 978), 
who  calculated  reaction  temperatures  of 
about  1 300°C  for  1  kbar  and  about  1 350°C 
for  3  kbar.  It  should  be  noted,  however,  that 
the  anhydrous  melting  of  diopside  + 


GEOPHYSICAL  LABORATORY 


87 


akermanite  +  forsterite  assemblages  will 
occur  at  1357°C  at  1  arm  (Ricker  and 
Osborn,  1954).  A  hydrous  fluid  phase 
involved  in  the  melting  will  lower  the 
solidus  considerably,  well  below  the  tem- 
perature of  the  solid- solid  reaction  as  cal- 
culated by  Helgeson  et  al.  (1978).  The 
diopside  +  calcite  <=>  akermanite  +  CO2 
reaction  will  terminate  at  the  CO2-  and 
calcite-absent  reaction  of  monticellite  + 
wollastonite  <=>  akermanite  (Yoder,  1973; 
Huckenholz,  1990;  this  study).  The  result- 
ing isobaric  invariant  point  I4  for  1  kbar 
occurs  at  700°C  and  atX(C02)  =  0.1;  for  3 
kbar,  it  is  found  to  be  at  720°C  and  at 
X(C02)  =  0.05  ±  0.02. 


Discussion 

With  the  experimental  data  at  hand,  it 
has  been  demonstrated  that  merwinite  + 
CO2  is  stable  over  a  wide  range  of  tempera- 
ture but  is  restricted  to  pressures  below 
about  1.5  kbar.  Merwinite  +  V,  however, 
may  also  form  atX(C02)  as  low  as  0.05  and 
at  temperatures  as  low  as  700°C  from 
spurrite  +  monticellite  +  akermanite  and 
from  spurrite  -1-  monticellite  +  wollastonite 
assemblages  in  H20-rich  aureoles  of  car- 
bonaceous rocks  around  basic  intrusives. 
At  pressures  below  1 .3  kbar  merwinite  + 
calcite  +V  and  merwinite  +  V  assemblages 
reach  the  solidus  between  1 1 50°  and  1 200°C 
and  will  melt  at  any  X(C02)  up  to  1 .0.  At 
pressures  above  1 .3  kbar  the  merwinite  +  V 
stability  shifts  toward  lower  X(C02)  and 
akermanite  +  calcite  +  V  assemblages  reach 
the  solidus  and  are  subject  to  melting.  At  3 
kbar,  these  same  relationships  occur  at 
1 1 50°- 1 200°C  and  X(C02)  >  0. 1 5 . 

High-temperature  phase  assemblages 
occur  as  rock  inclusions  in  the  critical  zone 


of  the  Bushveld  Complex: 

( 1 )  merwinite + monticellite + melilite 
(equivalent  to  mixture  E), 

(2)  calcite  +  periclase  +  monticellite, 
and 

(3)  forsterite  +  periclase  + 
monticellite. 

A  different  set  of  high-temperature  phase 
assemblages  occur  as  rock  inclusions  in  the 
marginal  zone: 

(4)  akermanite  +  monticellite  +  cal- 
cite (equivalent  to  mixture  D), 

(5)  calcite  +  forsterite  +  monticellite, 

(6)  akermanite  +  diopside  + 
monticellite  (equivalent  to  inclusions 
Jl  and  J6),  and 

(7)  diopside  -1-  forsterite + monticellite. 
According  to  Wallmach  et  al.  (1989),  tem- 
peratures as  high  as  1300°  and  1200°C, 
respectively,  for  estimated  pressures  at 
about  1  kbar  (0.6  to  1 . 1  kbar)  and  2  kbar 
(1.1  to  2.4  kbar)  are  deduced.  Experimen- 
tal results  obtained  on  merwinite  +  V  as- 
semblages (mixtures  A,  B,  C,  and  Y), 
merwinite  +  monticellite  +  melilite  assem- 
blages (mixture  E),  akermanite  + 
monticellite  +  calcite  assemblages  (mix- 
ture D)  as  well  as  on  the  rocks  (Ji  and  J6) 
containing  akermanite  +  monticellite  +  di- 
opside are  clearly  above  their  solidi  when 
treated  at  1200°C  and  at  pressures  of  1  and 
3  kbar,  respectively,  with  X(C02)  between 
0.3  to  1.0.  Thus,  the  deduced  temperatures 
of  Wallmach  et  al.  (1989)  appear  to  be 
excessive. 


References 

Agrell,  S.  O.,  Poly  thermal  metamorphism  of  lime- 
stones at  Kilchoan,  Ardnamurchan,  Mineral. 
Mag.,  34,  (Tilley  vol.)  1-15,  1965. 

Bowen,  N.  L.,  Progressive  metamorphism  of  sili- 
ceous limestone  and  dolomite,  J.  Geoi,  48, 
225-274,  1940. 


88 


CARNEGIE  INSTITUTION 


Burnham,  C.  W.,  Contact  metamorphism  of  mag- 
nesian  limestone  at  Crestmore,  California, 
GeoL  Soc,  Amer.  Bull,  70,  879-920,  1959. 
Helgeson,  H.  D.,  J.  M.  Delany,  H.  W.  Nesbitt,  and 
D.  K.  Bird,  Summary  and  critique  of  the  ther- 
modynamic properties  of  rock-forming  min- 
erals, Amer.  J.  Sci.,  278-A,  1-229,  1978. 
Holloway,  J.  R.,  Fugacity  and  activity  of  molecu- 
lar species  in  supercritical  fluids,  in  Thermo- 
dynamics in  Geology,  D.  Fraser,  ed.,  161-181, 
1977. 
Huckenholz,  H.  G.,  A.  Wassermann,  and  K.  T. 
Fehr,  Stability  and  phase  relations  of  gehlenite- 
akermanite  solid  solutions  in  the  presence  of  a 
H2O-CO2  fluid,  Third  International  Sympo- 
sium Experimental  Mineralogy ,  Petrology,  and 
Geochemistry,  p.  81,  Edinburgh,  1990. 
Joesten,  R.  C,  Pseudomorphic  replacement  of 
melilite  by  idocrase  in  a  zoned  calc-silicate 
skarn,  Christmas  Mountains,  Big  Bend  Re- 
gion, Texas,  Amer.  Mineral.,  59,  694-699, 
1974. 
Joubert,  J.,  Gemetamorfoseerde  karbonaatins 
luitsels      in      die      bosone      van      die 
Bosveldstollingskompleks,     was     van 
Roossenekel,  Transvaal,  M.Sc.  thesis,  Univ. 
Pretoria,  Hillcrest,  South  Africa,  1976. 
Kadik,  A.  A.,  and  D.  H.  Eggler,  Melt- vapor  rela- 
tions on  the  join  NaAlSi30s-H20-C02, 
Carnegie  Instn.  Washington  Year  Book,  74, 
479-484,  1975. 
Kushiro,  I.,  and  H.  S.  Yoder,  Jr.,  Stability  field  of 
akermanite,  Carnegie  Instn.  Washington  Year 
Book,  63,  84-86,  1964. 
Larsen,  E.  S.,  and  W.  F.  Foshag,  Merwinite,  a  new 
mineral  from  the  contact  zone  at  Crestmore, 
California,  Amer.  Mineral.,  6,  143-148, 1921. 
Mysen,  B.  O.,  Stability  of  volatiles  in  silicate 
melts  at  high  pressure  and  temperature, 
Carnegie  Instn.  Washington  Year  Book,  74, 
454-478,  1975. 
Ricker,  R.  W.,  and  E.  F.  Osborn,  Additional  phase 
equilibrium  data  for  the  system  CaO-Mg02- 
Si02,  /.  Amer.  Ceram.  Soc,  37,  133-139, 
1954. 
Rosenhauer,  M.,  and  D.  H.  Eggler,  Solubility  of 
H2O  and  CO2  in  diopside  melt,  Carnegie 
Instn.  Washington  Year  Book,  74,  474-479, 
1975. 
Schairer,  J.  F.,  H.  S.  Yoder,  Jr.,  and  C.  E.  Tilley, 
Behavior  of  synthetic  melilites  in  the  join 
gehlenite  -  soda  melilite  -  akermanite,  Carnegie 
Instn.  Washington  Year  Book,  65,  217-226, 
1967. 
Shmulovich,  K.  I.,  Stability  of  merwinite  in  the 
system  CaO-MgO-Si02-C02,D^/./4c^.5d, 
USSR,  Earth  Sci.  Sect.,  184,  125-127,  1969. 
Smith,  J.  V.,  X-ray  emission  microanalyses  of 
rock-forming  minerals,  VI.  Clinopyroxenes 
near  the  diopside-hedenbergite  join,  /.  GeoL, 


74,  463-477,  1966. 

Tilley,  C  E.,  On  larnite  (calcium  orthosilicate,  a 
new  mineral)  and  its  associated  minerals  from 
the  limestone  xenoliths  in  the  eastern  Bushveld 
Complex,  Canad.  Mineral.,27, 509-523, 1989. 

Walter,  L.  S.,  Experimental  studies  on  Bowen's 
decarbonation  series,  I:  P-T  univariant  equi- 
libria of  the  "monticellite"  and  "akermanite" 
reactions,  Amer.  J.  Sci.,  261, 488-500,  1963a. 

Walter,  L.  S.,  Experimental  studies  on  Bowen's 
decarbonation  series  II:  P-T  univariant  equi- 
libria of  the  reaction:  forsterite  +  calcite  = 
monticellite  +  periclase  +  CO2,  Amer.  J.  Sci., 
261, 173-179,  1963b. 

Walter,  L.  S.,  Experimental  studies  on  Bowen's 
decarbonation  series  III:  P-T  univariant  equi- 
libria of  the  reaction  spurrite  +  monticellite  = 
merwinite  +  calcite  and  analyses  found  at 
Crestmore,  California,  Amer.  J.  Sci.,  263,  64- 
77, 1965. 

Yoder,  H.  S.,  Jr.,  Melilite  stability  and  paragen- 
esis,  Fortschr.  Mineral,  50,  140-173,  1973. 

Yoder,  H.  S.,  Jr.,  Relationship  of  melilite- 
bearing  rocks  to  kimberlite.  A  preliminary 
report  on  the  system  akermanite-C02,  Phys. 
Chem.  Earth,  9,  883-894,  1975. 


The  System  MG2S1O4-FE2S1O4 
at  Low  Pressure 

Hiroko Nagahara, Ikuo Kushiro*  and 
Bjorn  O.  Mysen 

Gas -solid  relationships  are  important 
when  we  consider  condensation,  evapora- 
tion, and  fractionation  of  the  solar  nebula, 
especially  in  regard  to  bulk  composition  of 
the  Earth  and  the  terrestrial  planets.  Gas- 
solid  relationships  of  minerals  are  different 
from  those  between  liquid-gas,  and  ther- 
modynamic data  are  insufficient  to  con- 
struct gas-solid  phase  diagrams.  Mysen 
and  Kushiro  (1988)  and  Kushiro  and  Mysen 
(1991)  measured  vapor  pressures  of  MgO, 


*  Geological  Institute,  University  of  Tokyo,  Hongo, 
Tokyo  113,  Japan 


GEOPHYSICAL  LABORATORY 


89 


Si02,  forsterite,  and  enstatite,  and  studied 
the  phase  relations  of  the  MgO-Si02  sys- 
tem. In  this  study,  we  have  measured  the 
vapor  pressure  of  fayalite,  and  based  on 
those  results  along  with  those  for  forsterite, 
the  gas-solid  phase  relationships  of  the 
olivine  system  as  a  function  of  the  Fe/ 
(Mg+Fe)  ratio  are  proposed. 

Vapor  pressure  measurements  of  fayalite 
were  made  using  a  Knudsen  cell  method 
similar  to  that  described  by  Mysen  and 
Kushiro  (1988).  The  starting  material  is  a 
single  crystal  of  fayalite,  about  2x2x5  mm 
in  size,  synthesized  with  the  Czochralski- 
pulling  method  by  H.  Mori  of  the  Univer- 
sity of  Tokyo.  The  crystal  was  powdered 
(1-10  Jim),  and  several  mg  of  the  sample 
was  placed  in  a  molybdenum  capsule  with 
two  2-mm  holes  drilled  in  sides.  The  ex- 
periments were  carried  out  in  vacuum  fur- 
naces in  the  University  of  Tokyo  and  the 
Geophysical  Laboratory;  the  two  furnace 
designs  are  nearly  identical  in  size  and 
design  (Mysen  and  Kushiro,  1 988).  Samples 
were  heated  at  a  rate  of  15-20°/min  from 
room  temperature  to  experimental  tem- 
peratures, which  ranged  from  1050°C  to 
1 175°C.  Run  durations  ranged  from  4  days 
at  1175°C  to  12  days  at  1075°C.  Total 
pressure  of  the  vacuum  chamber  was  4.0  x 
10-7  torr  (5.3  x  10"1(>  bar)  to  6.0  x  10-7  torr 
(7.9x10-9  bar). 

Experiments  in  molybdenum  capsules 
were  at  the  oxygen  fugacity  (fo2)  of  the 
M0-M0O2  buffer  to  ensure  that  fayalite  is 
stable.  The  M0-M0O2  buffer  is  about  1.5 
orders  of  magnitude  higher  than  the  iron- 
wiistite  (IW)  buffer,  and  1  to  1.5  orders  of 
magnitude  below  the  quartz-fayalite-mag- 
netite  (QFM)  buffer  (Mysen  and  Kushiro, 


1988).  Measured  weight  loss  (1-14  %)  was 
calibrated  against  vapor  pressure  by  the 
equation 


P    -    1 

1  m  — 


m 


dw 


Ac   dt 


v 


2kRT 
M 


(U 


where  Pm  is  the  vapor  pressure  of  a  sub- 
stance, A  is  the  area  of  the  orifice  of  the 
capsule,  c  is  the  clausing  factor,  dw  is  the 
weight  loss,  dt  is  duration,  M  is  the 
molecular  weight  of  the  effusing  vapor,  7is 
the  absolute  temperature,  and  R  is  the  gas 
constant  (Paule  and  Margrave,  1967).  The 
clausing  factor  for  the  cell  has  previously 
been  determined  for  Cu  and  Ag  by  Mysen 
and  Kushiro  (1988). 

The  residue  of  partial  evaporation  re- 
mained fayalite,  indicating  that  fayalite 
evaporates  congruently.  Since  forsterite 
evaporates  congruently  and  intermediate 
olivine  evaporates  stoichiometrically 
(Mysen  and  Kushiro,  1988;  Nagahara  et 
al.y  1988),  it  is  clear  that  olivine  evaporates 
stoichiometrically  regardless  of  the  Fe/Mg 
ratio.  Fayalite  heated  at  1175  °C  melted, 
but  fayalite  heated  at  1 160°C  did  not,  sug- 
gesting that  the  melting  point  is  about 
1 170°C  and  2  x  10-8  bar.  This  is  about  35°C 
lower  than  that  at  1  bar  (Bo  wen  and  Schairer, 
1935).  Lower  melting  temperatures  rela- 
tive to  melting  temperature  at  1  bar  have 
also  been  found  for  Si02  and  Mg2Si04  at 
low  pressures  (by  100°  and  200°C,  respec- 
tively) (Mysen  and  Kushiro,  1988).  Mysen 
and  Kushiro  (1988)  further  demonstrated 
the  presence  of  a  three-phase  region  be- 
tween that  of  solid  and  gas  and  that  of  liquid 
and  gas.  This  three  phase  region  (forsterite 
+  liquid  +  vapor)  implies  that  the  system 


90 


CARNEGIE  INSTITUTION 


1040 


1080 


1120 
T  (<>C) 


1160 


1200 


Fig.  5 1 .  Temperature  and  vapor  pressure  relation- 
ship of  fayalite.  Error  bar  represents  weight  mea- 
surement uncertainty. 


can  not  be  described  as  a  binary  (MgO- 
S1O2).  Therefore,  phase  relations  where 
liquid  is  present  will  not  be  discussed  in  the 
present  study. 

The  experimentally  determined  tem- 
perature and  vapor  pressure  (Pv)  relation- 
ships for  Fe2Si04  are  summarized  in  Fig. 
51.  The  relationship  is  further  shown  in 


c 


-12      ■ 


-16 


-20 


-24 


1/TxlO4  (1/K) 

Fig.  52.  Arrhenius  plot  for  vaporpressure  of  fayalite 
(this  work)  and  forsterite,  MgO,  and  Si02  (Mysen 
and  Kushiro,  1988).  Linear  regression  line  from 
the  fayalite  data  gives  In  />v=(-608±60)/Rr  + 
(273±9)/R  and  r  =  0.9976. 


Fig.  52  together  with  the  l/T  vs.  In  Pv  of 
Mg2Si04,  MgO,  and  Si02-  The  linear  re- 
gression curve  from  the  data  points  can  be 
expressed  in  the  equation  for  evaporation 


In  Pv  =  :Mv  +  ASv 
RT         R 


(2) 


where  A//v  and  ASV  are  enthalpy  and  en- 
tropy of  evaporation,  respectively.  From 
the  linear  regression,  A//v  is  608  ±  60  (1  o) 
(kJ/mol)  and  ASV  is  273  ±  9  (J/K-mol). 
These  values  are  similar  to  those  for 
forsterite  (640  ±  36  and  210±54,  respec- 
tively) at  the  same/02  (Mysen  and  Kushiro, 
1988). 

The  gas-solid  phase  diagram  was  drawn 
by  using  the  enthalpies  and  entropies  for 
evaporation  of  forsterite  and  fayalite  (Fig. 
53).  In  order  to  draw  the  phase  diagram, 
two  assumptions  were  made:  (1)  chemical 
equilibrium  is  achieved  between  crystals 
and  gas  in  a  Knudsen  cell,  and  (2)  both 
olivine  and  gas  are  ideal  solutions.  The 
assumption  (1)  can  be  valid.  The  residues 
are  sintered  homogeneous  fayalite  crystals 
regardless  of  experimental  duration.  Ac- 
cordingly, chemical  compositions  of  coex- 
isting gas  and  solid  have  been  uniform.  The 
assumption  (2),  ideality  of  the  olivine  solid 
solution  system,  has  been  shown  by  many 
investigators  (i.  e.,  Wood  and  Kleppa,  1981). 
Gas  can  be  treated  as  ideal. 

With  the  assumptions  made  above,  mol 
fractions  of  the  fayalite  component  in  gas 
and  solid  at  a  given  pressure  are  shown  by 
the  following  equations 


GEOPHYSICAL  LABORATORY 


91 


1800 


-" 1 • r 


^     1600 

CD 

■*-> 
(0 

i_ 

0) 
Q. 

E 

£      1400 


1200 


1800 


* 

1600 

<D 

k_ 

T 

■*-• 

CO 

O 

Q. 

F 

0) 

1400 

h- 

1200 


GAS 


OLIVINE  s.s.+  GAS 


I0"8bar 


J I l_ 


■ I I 1_ 


20  40  60  80  100 


"> 1 " r 


T r 


10"10  bar 


GAS 


_. 1 i_ 


1800 


^   1600 


1200 


-i 1 «" 


GAS 


10    bar 
j i i k 


20  40  60  80  100 

Fe/(Mg+Fe)x100 


20  40  60  80  100 

Fe/(Mg+Fe)x100 


Fig.  53.  Gas-solid  phase  diagrams  of  the  olivine 
system  at  10"8,  10A  and  10r°  bar. 


ln*-  = 
x 


and 


'_     AHFi 


\T0  ~fl 


In 


\-x_     AHFo 
1  -x 


R 


L 

To 


(3) 


(4) 


where  x  and  xy  are  the  Fe/(Fe+Mg)  ratios  of 
gas  and  solid,  respectively,  and  T0and  Tq 
are  vaporus  temperatures  for  fayalite  and 
forsterite,  respectively. 


The  calculated  vaporus  and  solidus 
curves  are  shown  in  Fig.  53.  The  conspicu- 
ous feature  of  the  figure  is  that  the  binary 
loop  is  quite  flat.  With  decreasing  pressure, 
the  vaporus  temperatures  for  both  forsterite 
and  fayalite  become  lower  and  the  loop 
becomes  more  flat.  The  figure  sows  that  the 
compositional  difference  between  coexist- 
ing solid  and  gas  is  extremely  large  com- 
pared to  that  between  solid  and  liquid  at  L 
bar  (Bo wen  and  Schairer,  1935).  There  is  a 
very  narrow  temperature  interval  over 
which  Fe -bearing  olivine  coexists  with  gas. 
Thus,  olivine  should  have  become  forsterite 


92 


CARNEGIE  INSTITUTION 


regardless  of  the  primary  composition  when 
it  was  heated  at  subvaporus  temperatures. 
Formation  of  forsterite  would  have  been 
kinetically  suppressed  when  equilibrium 
between  gas  and  solid  was  not  achieved. 
Hashimoto  (1990)  showed  that  forsterite 
evaporates  very  slowly  at  a  free  evapora- 
tion condition  (at  a  rate  one-tenth  of  that  in 
equilibrium);  that  is,  if  the  generated  gas 
was  removed  from  the  system  and  did  not 
equilibrate  with  the  solid,  formation  of 
forsterite  would  be  suppressed.  Another 
possible  factor  preventing  formation  of 
forsterite  is  cation  diffusion  in  solid 
forsterite,  which  depends  on  temperature, 
heating  duration,  and  the  grain  size  of  oli- 
vine. Although  the  Mg-Fe  inter-diffusion 
coefficient  in  olivine  is  largest  among  any 
other  elemental  diffusion  coefficients  in 
olivine  and  those  known  in  any  other  sili- 
cate minerals  (Freer,  1981),  short  heating 
and/or  large  grain  size  can  be  rate-limiting 
factors  for  formation  of  forsterite  as  partial 
evaporation  residue.  If  these  processes  were 
not  effective,  olivine  should  have  become 
forsterite  quickly  by  heating  at  subvaporus 
temperatures. 

The  phase  diagram  is  not  directly  appli- 
cable to  evaporation  in  the  solar  nebula; 
effects  of  other  components,  pressure,  and 
foi  should  be  evaluated.  Other  components, 
such  as  Al  and  Ca,  would  affect  the  abso- 
lute temperature  of  vaporus  and  solidus, 
but  would  not  change  significantly  the  shape 
of  the  diagram  because  of  much  smaller 
abundance  in  the  solar  nebula  of  these 
components  compared  with  Si,  Mg,  and  Fe 
(Anders  and  Ebihara,  1982).  The  pressure 
range  in  the  present  experiments  is  just 


applicable  to  the  solar  nebula.  Elemental 
abundances  of  Mg  and  Si  in  the  solar  sys- 
tem are  about  4  orders  of  magnitude  smaller 
than  that  of  H,  and  the  total  pressure  at  the 
midplane  of  about  3  A.U.  has  been  gener- 
ally   calculated  to  be  between  10-3  and 
10-5  bar  (Cameron,  1985;  Morfill  et  al., 
1985).  The  approximate  partial  pressure 
for  olivine  component  is  thus  107  to  10-9 
bar,  which  well  agree  with  the  pressure 
range  in  the  present  work.  Oxygen  fugacity 
condition  of  the  present  work  is  much  more 
oxidizing  than  that  estimated  for  the  solar 
nebula,  based  on  the  elemental  abundances 
of  the  solar  system.  However,  oxygen  fu- 
gacity as  high  as  the  present  work  has  been 
proposed  recently  for  the  formation  of  vari- 
ous chondritic  components  in  the  solar 
nebula  (Fegley  and  Palme,  1985;  Palme 
andFegley,  1990;  Weinbruclmtf/.,  1990). 
The  present  result  can  be,  thus,  applied 
almost  directly  to  the  conditions  for  forma- 
tion of  chondritic  components  in  the  solar 
nebula. 


References 

Anders,  E.,  and  M.  Ebihara,  Solar-system  abun- 
dances of  the  elements.  Geochim.  Cosmochim. 
Acta,  46,  2363-2380,  1982. 

Bowen,  N.  L.  and  Schairer,  J.  F.,  The  system 
MgO-FeO-Si02,  Amer.  Jour.  Sci.,  ser.  5,  29, 
151-217,  1935. 

Cameron,  A.  G.  W.,  Formation  and  evolution  of 
the  primitive  solar  nebula,  in  Protostar  and 
Planets II,  D.  C.  Black  and  M.  S.  Matthews,  eds., 
Univ.  Arizona  Press,  Tucson,  Arizona,  pp.  1073- 
1099,  1985. 

Fegley,  B.,  Jr.  and  H.  Palme,  Evidence  for  oxidiz- 
ing conditions  in  the  solar  nebula  from  Mo  and 
W  depletion  in  refractory  inclusions  in  carbon- 
aceous chondrites,  Earth  Planet.  Sci.  Lett.,  72, 
311-326,1985. 


GEOPHYSICAL  LABORATORY 


93 


Freer,  R.,  Diffusion  in  silicate  minerals  and  glasses: 
A  data  digest  and  guide  to  the  literature.  Contrib. 
Mineral  Petrol,  76,  440-454,  1981. 
Hashimoto,  A.  Evaporation  kinetics  of  forsterite 
and  implications  for  the  early  solar  nebula. 
Nature,  347,  53-55,  1990. 
Kushiro,  I.  and  B.  O.  Mysen,    in  Progress  in 
Metamorphic  and  Magmatic  Petrology  ,  L.  L. 
Perchuked,  Cambridge  Univ.  Press,  Cambridge, 
England,  411-433,  1991. 
Morfill,  G.  E.,  W.  Tscharnuter,  and  H.  Volk,  in 
Protostar  and  Planets  II,  D.  C.  Black  and  M.  S. 
Matthews,  eds.,  Univ.  Arizona  Press,  Tucson, 
Arizona,  pp.  493-533,  1985. 
Mysen,  B.  O.  and    I.  Kushiro,  Condensation, 
evaporation,  melting,  and  crystallization  in  the 
primitive  solar  nebula:  Experimental  data  in  the 
system  MgO-Si02-H2  to  1.0xl0-9barand  1870°C 
with  variable  oxygen  fugacity,  Amer.  Miner., 
73,  1-19,  1988. 
Nagahara,  H.,  I.  Kushiro,  B.  O.  Mysen,  and  H. 
Mori,  Experimental  vaporization  and  condensa- 
tion of  olivine  solid  solution,  Nature,  331,  516- 
518,  1988. 
Palme,  H.  and  B.  Fegley,  Jr.,  High  temperature 
condensation  of  iron-rich  olivine  in  the  solar 
nebula,  Earth  Planet.  Sci.  Lett.,  101,  180-195, 
1990. 
Paule,  R.  C.  and  J,  L,  Margrave,  Free-evaporation 
and  effusion  techniques,  in  The  Characteriza- 
tion of  High  Temperature  Vapors,  J.  L.  Margraves 
ed.,  Wiley,  New  York,  130-151,  1967. 
Weinbruch,  S.,  H.  Palme,  W.  F.  Muller,  and  A.  El 
Goresy,  FeO-rich  rims  and  veins  in  Allende 
forsterite:  Evidence  for  high  temperature  con- 
densation at  oxidizing  conditions,  Meteoritics, 
25,  115-125,  1990. 
Wood,  B.  J.  and  O.  J.  Kleppa,  Thermochemistry  of 
forsterite-fayalite  olivine  solutions,  Geochim. 
Cosmochim.  Acta,  45,  529-534,  1981. 


Fe3+,  Mg  Order-Disorder  in  Heated 

MGFE2O4:  A  Powder  XRD 

and  57Fe  Mossbauer  Study 

H.  St.  C.  O'Neill;  H.  Annersten,** 
and  D.  Virgo 

The  two  extreme  cation  distributions  in 
the  spinel  structure  are  the  so-called  nor- 
mal configuration  A  [Bii  O4  and  the  inverse 
configuration  B[AB]04  where  the  [  ]  refer 
to  the  octahedrally  coordinated  cations  and 
the  remaining  cations  are  in  tetrahedral 
coordination.  Disordered  configurations  of 
both  these  extreme  arrangements  can  be 
represented  as  A\.x  BX[AX  #2-x]04.  The 
degree  of  disorder  can  also  be  discussed  in 
terms  of  the  inversion  parameter  x,  defined 
as  the  fraction  of  B  cations  occupying  the 
tetrahedral  sites. 

A  general  thermodynamic  model  of  the 
cation  distribution  in  spinels  has  been  pro- 
posed by  O'Neill  and  Navrotsky  (1983, 
1984).  The  basic  tenet  of  this  model  is  that 
the  equilibrium  cation  distribution  is  re- 
lated to  the  free  energy  of  disordering  in  the 
following  way, 


•RT\A—^ U^SL 

\(l-x)(2-x)j    \    dx   Itj,, 


N 


CD 


where  AGd  is  the  change  in  the  non-con- 
figurational  free  energy  of  disordering. 

In  their  model,  the  free  energy  term 
AGd  was  shown  to  consist  of  an  enthalpy 


*Bayerisches  Geoinstitut,  Universitat  Bayreuth, 
Germany 

**  Department  of  Mineralogy  and  Petrology,  Insti- 
tute of  Geology,  University  of  Uppsala,  Sweden 


94 


CARNEGIE  INSTITUTION 


term,  AHd  which  takes  a  quadratic  depen- 
dence on  x  (AHd  =  OUC+  Pjc^).  There  was 
also  a  non  configurational  entropy  term 
ASd-  The  non-linear  nature  of  the  enthalpy 
term  implies  that  cation  site  preference  will 
depend  on  the  degree  of  inversion  of  the 
spinel  into  which  it  is  substituting.  Previ- 
ously it  had  been  proposed  (Navrotsky  and 
Kleppa,  1967)  that  the  site  preference  en- 
thalpies in  spinels  were  independent  of 
both  temperature  and  the  degree  of  order  in 
the  spinel  structure. 

The  experimental  basis  for  the  proposed 
model  of  non-linear  enthalpy  of  disorder- 
ing in  spinel  solid  solutions  is  limited 
(O'Neill  and  Navrotsky,  1983,  Nell  et  aL, 
1989).  Therefore,  the  present  ordering  and 
disordering  experiments  on  MgFe204  have 
been  specifically  designed  to  test  the  above 
thermodynamic  model.  Magnesio-ferrite 
is  an  ideal  composition  in  which  to  carry 
out  such  tests,  since  it  shows  a  relatively 
large  change  in  equilibrium  cation  distri- 
bution over  a  wide  temperature  range,  it  is 
stable  in  air,  and  the  degree  of  inversion  can 
be  determined  by  several  techniques  such 
as  x-ray  refinement,  and  57Fe  Mossbauer 
spectroscopy.  Data  in  the  literature  are 
viewed  as  unsatisfactory,  in  view  of  the 
possibility  either  that  the  samples  studied 
were  not  stoichiometric  or  that  the  cation 
site  populations  do  not  correspond  to  the 
annealing  temperatures. 


Experimental 

MgFe204  was  synthesized  in  air  using 
a  sodium  tungstate  flux.  The  oxide  compo- 


nents consisting  of  2  g  MgO,  4  g  Fe203, 20 
g  Na2W04,  and  2  g  WO3  were  melted  in  a 
Pt  crucible  at  1260  C  and  then  cooled  at 
6°C  per  hour  to  either  950°C  (first  batch)  or 
900°C  (second  batch),  at  which  tempera- 
tures the  melt  was  crystallized  for  approxi- 
mately 12  hours  before  the  final  cooling  to 
room  temperature.  The  sodium  tungstate 
flux  was  dissolved  in  warm  water.  Excess 
MgO  was  removed  in  dilute  nitric  acid. 
The  MgFe204  run  products  consisted  of 
euhedral,  octahedra  of  reddish-brown 
spinel,  -10  pxn  in  size. 

The  acid  cleaned  MgFe204  was 
analysed  by  ICP  and  by  two  different  elec- 
tron microprobe  systems,  one  at  the 
Bayerishes  Geoinstitut  and  the  other  at  the 
University  of  Uppsala.  The  MgO/Fe203 

values  were  respectively  1.014  ± 
0.019,1.014±0.009andl.028±0.015.All 

three  sets  of  analyses  indicate  that  MgFe204 
is  stoichiometric  within  the  analytical  un- 
certainty (ie.  within  2  standard  deviations). 
Ideal  stoichiometry  is  assumed  in  the  fol- 
lowing discussion  of  the  x-ray  and  57Fe 
Mossbauer  results. 

Both  disordering  and  ordering  experi- 
ments were  carried  out  in  vertical  quench 
furnaces  in  the  range  400-1250°C.  by  heat- 
ing 40-100  mg  of  the  starting  material  in 
platinum  capsules  welded  at  one  end  and 
crimped  at  the  other  end.  Temperatures 
were  controlled  to  ±0.5 °C,  and  run  dura- 
tions were  up  to  50  days.  The  heated 
samples  were  quenched  in  water.  Effective 
quench  times  calculated  by  extrapolation 
of  the  rate  studies  were  of  the  order  of  a  few 
milliseconds. 


GEOPHYSICAL  LABORATORY 


95 


Lattice  Parameter  Measurements  and 
Powder  XRD  Structural  Refinements 

Lattice  parameters  were  measured  us- 
ing CoA'ai  radiation  on  a  STOE  focusing 
diffractomer  with  a  curved  Ge  monochro- 
mator.  A  position-sensitive  detector  of  8° 
20  width  and  0.5°  data  interval  was  used  to 
collect  the  diffraction  data  over  the  range 
40-120°  2£  NBSSi  (0  =  5.43087)  was  used 
as  an  internal  standard.  Peak  positions 
were  determined  by  fitting  the  peak  profile, 
and  the  observed  20  's  were  adjusted  by 
using  a  linear  correction  based  on  the  six 
main  Si  peaks  in  the  two  regions  of  interest. 
The  spinel  lattice  parameter  a  was  then 
determined  by  a  weighted  least  squares 
refinement  of  the  positions  of  the  ten  most 
intense  spinel  peaks.  The  mean  internal 
estimate  of  one  standard  deviation  of  a  is 
calculated  to  be  ±  0.00015.  An  external 
estimate,  based  on  determinations  of  a  for 
samples  equilibrated  at  600, 650  and  700° C 
and  heated  for  times  that  exceed  that  re- 
quired for  equilibrium  gives  estimates  in 
theuncertainty  of0of±O.OOOll  toO.00013. 

The  XRD  structural  refinements  were 
measured  on  data  using  the  same 
diffractometer  but  with  Mo/foci  radiation 
and  over  the  range  in  20  from  7  to  80°.  Data 
were  collected  from  five  consecutive  scans 
and  the  scans  were  summed.  Crystal  struc- 
ture refinements  were  made  using  two  con- 
trasting methods  which  are  more  fully  dis- 
cussed by  O'Neill  et  al.  (1991a).  Here, 
results  for  full  profile  Rietveld-type  refine- 
ments using  the  program  DBW  (Wiles  and 
Young,  1987)  are  considered  (see  also 
O'Neill  et  al,  1991b)  The  structure  was 
refined  in  space  group  Fd3m    with  8(a) 


andl6(d)  cation  sites  and  32(e)  oxygen 
sites  all  fully  occupied.  The  refined  quanti- 
ties were  (1)  a  scale  factor,  (2)  the  Pearson 
VII  profile  shape  parameters  in  which  the 
component  m  could  vary  as  m  =  a  +  b/20 , 
(3)  an  asymmetry  parameter  for  peaks  with 
2q  <  34°,  (4)  a  20  zero  parameter,  (5)  Peak 
halfwidth  function  of  the  form  FWHM2  = 
UtmO  +  VtmO  +  W,  (6)  the  unit  cell 
constant  a  (7)  the  oxygen  positional  param- 
eter m  ,  (8)  the  inversion  parameter,  x,  and 
(9)  isotropic  temperature  factors  for  either 
cation  sites  or  oxygen,  or  both.  Refine- 
ments were  also  made  for  the  cases  where 
the  raw  data  background  was  either  in- 
cluded or  subtracted.  Bragg  and  total  pat- 
tern residuals  for  the  equilibrated  samples 
were  in  the  ranges  1 .97-3.63  and  2.89-4.57, 
respectively.  There  is  no  systematic  corre- 
lation between  these  residuals  and  the  de- 
gree of  inversion. 


57Fe  Mossbauer  Spectroscopy 

Conventional  transmission  spectra  were 
collected  in  a  512  multi-channel  analyzer 
operated  in  conjunction  with  a  constant 
accelerator  electro-mechanical  drive  unit. 
57Co  in  Rh  was  used  as  a  source.  A  calibra- 
tion spectrum  for  natural  iron  at  room  tem- 
perature was  simultaneously  recorded  at 
the  other  end  of  the  vibrator  unit.  The  two 
mirror-symmetric  spectra  typically  con- 
tained 0.6-0.8  x  10^  counts/channel,  and 
the  counts/channel  for  each  half  of  the 
spectral  data  were  averaged  before  analy- 
sis by  a  least-squares  fitting  program. 

The  room  temperature  spectra  of 
magnesio-ferrite  (Neel  point  around  600 


96 


CARNEGIE  INSTITUTION 


100.0 


o  99.0 


8  98.0 

jQ 

CO 

§  97.0 

c 

o 

w   ~~  ~ 

<i>  96.0 


95.0      - 


10 


•5  0  5 

Velocity,  mm/s 


10 


o 
w 
-Q 

CO 


cz 
CO 
c 
o 

CD 
DC 


100.0 


o    99.0 


98.0    - 


E    97.0    - 


96.0 
95.0 


-10 


0 
Velocity,  mm/s 


10 


Fig.54.  57Fe  Mossbauer  spectra  of  stoichiometric  MgFe204  measured  in  an 
external  field  of  4.5  Tesla.  The  absorption  patterns  labeled  A  and  B  refer  to  Fe3+ 
in  the/4  and£-sites,  respectively.  Upper  spectrum  is  MgFe204  heated  at  1050°C, 
35  mins.,  Absorber  temperature  =  171  K.  Lower  spectrum  is  MgFe2C>4  heated 
500° C,  8  days,  Absorber  temperature  =  189  K. 


K)  indicates  that  there  is  almost  complete 
overlap  of  the  magnetically  split  patterns 
due  to  Fe3+  in  the  tetrahedrally  and  octahe- 
drally  coordinated  sites.  Accordingly,  the 


spectral  data  were  collected  in  the  presence 
of  an  externally  applied  magnetic  field 
using  a  superconducting  magnet,  cooled  to 
4.2  K  to  produce  a  magnetic  field  of  4.5 


GEOPHYSICAL  LABORATORY 


97 


Tesla.  In  these  experiments,  the  external 
magnetic  field  is  orientated  parallel  to  the 
propagation  of  the  gamma  ray  and  thus  the 
intensity  of  the  transitions  AMi  =0  vanish; 
in  addition,  the  magnetic  hyperfine  field  at 
the  tetrahedral  site  increases  while  that  at 
the  octahedral  site  decreases.  Thus,  each 
six-line  pattern  of  the  298  K  spectra  of 
MgFe204  is  reduced  to  four  lines  in  the 
presence  of  the  external  field  (cf  Fig.  54). 
The  absorbers  were  made  by  mixing 
powdered  samples  of  equilibrated  and 
quenched  MgFe204  with  -100  mg  of 
thermo-setting  plastic  powder,  (cf  Virgo 
and  Hafner,  1969).  The  absorber  density 
was  7  mg  Fe/cm2-  The  absorber  was  cooled 
in  the  cryostat  to  temperatures  in  the  range 
1 2- 1 7 1 K;  this  procedure  minimized  differ- 
ences in  the  recoil-free  fraction  at  the  A  and 
B  sites. 

Least  squares  fitting  of  the  spectral  data, 
carried  out  using  lines  of  Lorentzian  shape, 
incorporated  equal  half- widths  of  each  four- 
line  pattern,  and  the  intensity  ratios  were 
independent  of  whether  the  line  intensities 
were  unconstrained  or  whether  the  mag- 
netically split  lines  1,  3,  4,  and  6  were 
constrained  in  the  ratio  3:1:1:3.  It  is  evi- 
dent from  Fig.  54  that  the  absorption  due  to 
Fe3+  in  the  octahedrally  coordinated  sites 
is  significant  broader  compared  to  that  for 
Fe3+  in  the  tetrahedrally  coordinated  sites. 
It  is  also  evident  from  Fig.  54  that  the  half- 
width  of  the  Z?-site  absorption  increases 
with  increasing  disorder.  For  the  sample, 
heat  treated  at  1050°C,  four  hyperfine  split 
patterns  were  refined  in  the  fit  of  Fe3+(#- 
site)  whereas  only  three  such  patterns  were 
required  for  the  remaining  heat-treated  and 
equilibrated  samples  in  order  to  obtain  sta- 


8.41 


< 
._-  8.40 

CD 
CD 

E 

cfl 

co  8.39 

Q. 

CD 
O 

'& 

CO 


8.38- 


8.37 


I        I        I        I        I       I       I        I 
_  MgFe204 

■  ■   ■  ■ 

■ 
■ 

—                                ■                                — 
■ 

■ 

■ 

■ 

■ 

_  ■                                                             _ 

i        i        i        i        i       i       i        i 

400 


600  800  1000  1200 

Temperature  of  anneal,  °C 


Fig.  55.  Lattice  parameter  (a)  measurements  on 
heat-treated  and  quenched  samples  of  stoichio- 
metric MgFe2C>4. 

tistically  meaningful  fits.  For  all  values  of 
x  only  a  single  hyperfine  pattern  was  fitted 
to  the  absorption  due  to  Fe3+  in  the  tetrahe- 
dral site. 


Approach  to  Equilibrium 

The  lattice  parameters  were  used  to 
monitor  the  approach  to  the  equilibrium 
cation  distribution  in  both  disordering  and 
ordering  experiments.  It  can  be  shown 
from  the  results  of  this  study  that  a  varia- 
tion of  ±0.0001  in  a  corresponds  to  a 
variation  in  x  of  about  ±0.001.  The  cell 
constants  for  samples  annealed  and  then 
quenched  at  50°  intervals  in  the  range  450- 
1250°C  are  shown  plotted  in  Fig.  55.  Equi- 
librium is  achieved  on  a  time  scale  of  about 
30  days  at  450°C  and  less  than  five  minutes 
at  temperatures  of  700°C  and  above.  For 
equilibrium  compositions  in  the  range  550- 
700° C,  equilibrium  was  verified  by  rever- 
sal experiments  whereby  the  equilibrium 
values  of  x,  at  fixed  temperature,  were 
determined  with  starting  material  that  was 


98 


CARNEGIE  INSTITUTION 


both  disordered  and  ordered  with  respect  to 
the  equilibrium  value  (O'Neil ,  1991c). 

The  data  plotted  in  Fig .  5  5  show  a  smooth 
increase  in  a  as  a  function  of  temperature, 
although  there  is  only  a  small  increase  in  a 
above  1050°C.  Itisofconcernatthesehigh 
temperature  that  the  rate  of  ordering  is  so 
fast  in  the  high  temperature  experiments 
that  the  equilibrium  distribution  of  Mg  and 
Fe3+  is  not  quenched-in.  There  is  also  the 
possibility  that  there  are  deviations  from 
the  stoichoimetric  composition  at  high  tem- 
perature. While  both  those  possibilities  are 
discussed  in  more  detail  by  O'Neill  et  al. 
(1991b),  it  is  significant  that  the  thermody- 
namic model  to  be  discussed  below  and 
established  from  a  fit  to  the  cation  distribu- 
tion on  samples  annealed  at  temperatures 
of  less  than  1000°C  does,  in  fact,  predict 
that  there  will  only  be  comparatively  small 
changes  in  ao  at  temperatures  above  1 100°C. 


of  the  B -values  with  the  degree  of  inver- 
sion. 

The  values  of  x  are  shown  plotted  against 
the  annealing  temperature  in  Figure  56.  As 
expected  there  is  a  smooth  change  in  the 
cation  distribution  from  nearly  inverse  at 
450°  C  to  a  more  random  configuration  at 
high  temperature. 

The  spectra  of  MgFe204  in  an  applied 
field  are  similar  to  that  reported  by  Sawatsky 
et  al.  (1969).  Significantly,  the  absorption 
patterns  due  to  Fe3+  in  the  distinct  crystal 
sites  are  well  resolved  (Fig.  54).  The  iso- 
mer shift  values  for  Fe3+  in  tetrahedral 
versus  octahedral  coordination  are  0.31- 
0.37mm/sec  and  0.43-0.49  mm/sec,  respec- 
tively. 

The  broadening  of  the  Z?-site  absorption 
relative  to  the  A -site  absorption,  noted  pre- 
viously, can  now  be  similarly  explained  as 
for  other  inverse  spinels  (Sawatsky  et  al., 


Cation  Distributions  in  MgFe204:  XRD 
and  57 Fe  Mossbauer  data. 

The  structural  refinements  in  which 
separate  isotropic  temperature  factors  were 
refined  for  each  cation  site  and  for  oxygen 
gave  the  lowest  values  for  RBragg  and  Rf 
although  these  statistical  criteria  were  only 
slightly  improved  over  a  model  wherein  an 
average  temperature  factor  for  both  cation 
sites  and  for  oxygen  are  refined.  The  values 
for  the  temperature  factors  deceased  in  the 
order  Btet  <  B0ct  <B  oxygen,  and  the  mean 
values  are  Btet  =  0.304  ±  0.029  B0ct  = 
0.360  ±  0.023  and  B0Xygen  =  0.507 
±0.03 1  A2.  There  is  no  systematic  variation 


0.90 

I 

-    1% 

1             1 

1             1 

1        1        1 

S    XRD 

o  Mossbauer 

0.80 

O^i 

- 

n  ir\ 

1 

1             1 

o 
1          1 

i          i     .     T 

400 


600 


800 
Temperature, 


1000 


1200 


Fig.  56.  A  comparison  of  thermodynamic  models 
for  cation  disordering  in  MgFe204  a)  a  two  term 
model  with  no  excess  non-configurational  en- 
tropy of  disorder  (dashed  curve)  and  b)  three  term 
model,  including  an  excess  entropy  term  (unbro- 
ken curve).  The  XRD  data  are  plotted  with  one 
standard  deviation  error  bars;  uncertainty  in  the 
Mossbauer  data  is  0.008  to  0.0 1 2  and  is  not  shown 
for  clarity.  The  three  XRD  data  from  1 100  to 
1200°C  (open  symbols)  were  not  included  in  the 
regressions,  because  of  the  possibility  of  a  small 
oxygen  deficiency  in  these  samples. 


GEOPHYSICAL  LABORATORY 


99 


1969).  In  the  inverse  MgFe204,  each  tetra- 
hedral  site  is  surrounded  by  twelve  nearest 
Fe3+  neighbors.  On  the  other  hand,  each 
octahedral  site  is  surrounded  by  only  six 
tetrahedral  Fe3+  neighbors.  It  is  proposed 
that  the  broadening  of  the  B-site  absorption 
lines  is  due  to  the  different  number  of 
probable  distributions  of  Fe3+  and  Mg  on 
the  six  nearest  neighbor  A -sites.  Quite  the 
opposite  effect  is  proposed  for  the  A-site 
Fe3+,  because  there  is  no  apparent  line 
broadening  at  this  site.  For  MgFe204  with 
x  ~  0.66,  the  statistical  probabilities  that  the 
B  sites  have  distributions  of  6  Fe,  5  Fe  1 
Mg,  4  Fe  2Mg  and  3  Fe  3Mg  in  the  next 
nearest  neighbor  A  sites  are  0.24,  0.39, 
0.26,  and  0.09.  For  samples  that  are  nearly 
inverse  with  x  ~  1.0,  the  corresponding 
probabilities  are  0.78, 0.20, 0.02, 0.0.  Thus, 
the  increased  halfwidth  of  the  5-site  with 
increasing  disorder  (cf  Fig.  54)  is  reason- 
ably explained  in  terms  of  additional  hy- 
perfine  fields  due  to  the  next-nearest-neigh- 
bor effect. 

The  values  of  the  inversion  parameters 
calculated  from  the  area  ratios  as  deter- 
mined from  the  absorption  spectra  are  com- 
pared with  the  values  from  the  Rietveld 
refinements  in  Fig.  57.  The  mean  differ- 
ence in  the  values  of  x  is  0.0056  with  a 
standard  deviation  of  0.0102.  Thus,  the 
5 1  Fe  Mossbauer  data  are  in  excellent  agree- 
ment with  the  XRD  data,  although  it  is 
noted  that  this  agreement  would  become 
less  satisfactory  if  fully  ionized  atom  scat- 
tering factors  were  used  in  the  structural 
refinement  or  if  isotropic  temperature  fac- 
tors had  not  been  separately  refined  for 
both  cation  sites,  or  if,  in  fact,  only  a  single 
temperature  factor  was  refined. 


Thermodynamic  model. 

The  equilibrium  values  of  x  in  the  tem- 
perature range  450-1050°C  determined 
from  the  XRD  Rietveld  refinements  and 
the  57pe  Mossbauer  data  have  been  fitted 
to  the  expression 


-ln#  =  ccMS-Fe3+  +  2p;t, 


(2) 


where  K  is  the  distribution  coefficient 
(Navrotsky  and  Kleppa,  1967).  The  values 
of  x  were  weighted  according  to  their  stan- 
dard deviation,  namely  ±0.008  to  ±0.012 


Comparison  of  Mossbauer  and  XRD 


0.90 


To 
a> 

jD 

<n 

o    0.80 


0.70 


1:1  line 


0.70  0.80 

x  (Rietveld  refinements) 


0.90 


Fig.  57.  Comparison  of  the  powder  XRD  determi- 
nations of  x  from  the  Rietveld  refinements  with  the 
corresponding  values  determined  from  the  57Fe 
Mossbauer  spectra.  Data  are  plotted  with  one 
standard  deviation  error  bars. 


for  Mossbauer  data  and  ±0.004  for  the 
XRD  data.  The  results  of  the  fit  for  this  two- 
term  model,  shown  plotted  in  Fig.  56,  are 
aMg-Fe3+=  26.6  ±  0.4,  p  =  -21.7  ±  0.3  kJ/ 
mol  and  X^v  =  1.95.  There  is  an  improve- 
ment in  the  fit  if  an  additional  term  repre- 
senting a  nonconfigurational  entropy  of 
mixing  is  included.  Equation  (2)  becomes 

-In K  =  aMg-pe3+  -  7oMg-Fe3+  +2p;t ,  (3) 


100 


CARNEGIE  INSTITUTION 


with  a  Mg-Fe3+  =  16>9  +  2.5  kJ/mol  and 
aMg-Fe3+=  .2.67  ±  1.52  andX^  =  1.10.  It 
is  of  interest  in  the  latter  case  that  the  values 
of  x  for  the  heat  treated  samples  at  1100, 
1 150  andl250°C  are  in  agreement  with  the 
predicted  values  using  the  three-term  model. 
This  latter  result  is  significant,  of  course,  in 
terms  of  whether  the  high  temperature  cat- 
ion distributions  have  been  quenched  at  the 
respective  annealing  temperatures. 

In  the  formalism  of  the  thermodynamic 
model  proposed  by  O'Neill  and  Navrotsky 
[  1983, 1984;  equation  (1)]  the  nonconfigu- 
rational  entropy  term  inferred  above  is  taken 
to  refer  to  a  vibrational  energy  contribution 
and/or  the  effect  of  short-range  ordering 
(O'Neill  and  Navrotsky,  1983).  In  the 
latter  case  it  is  relevant  that  significant 
positional  disorder  of  Fe3+  and  Mg  on  the 
B  sub-lattice  is  inferred  from  the  57pe 
Mossbauer  spectra  (cf  Fig.  54). 

In  the  literature,  thermochemical  data 
required  to  evaluate  the  proposed  thermo- 
dynamic model  are  sparse.  The  enthalpy 
associated  with  the  change  in  cation  distri- 
bution in  MgFe204  in  the  temperature  range 
700-1200°C  has  been  measured  by  trans- 
posed-temperature-drop  calorimetry 
(Navrotsky,  1986;  Table  3).  The  experi- 
mental value  of  -5.5  kJ  does  not  agree  with 
the  value  of  —  1  kJ  predicted  from  the 
present  study  [equations  (2),  (3)].  How- 
ever, it  should  be  noted  that  the  stoichiom- 
etry  of  the  MgFe204  used  in  the  thermo- 
chemical measurement  was  not  specified. 
Further  evaluation  of  the  proposed  model 
for  cation  disordering  in  spinels  must  there- 
fore await  thermochemical  data  on  well- 
characterized  samples. 


References 


Navrotsky,  A.  and  O.  J.  Kleppa,  The  thermody- 
namics of  cation  distributions  in  simple  spinels. 
/.  Inorg.  Nucl.  Chem.,  29,  2701-2714,  1967. 

Navrotsky,  A.,  Cation-distribution  energetics  and 
heats  of  mixing  in  MgFe204-MgAl204} 
2nFe204-2nAl204  and  NiAl204-2nAl204 
spinels:  Study  by  high- temperature  calorim- 
etry. Amer.  Mineral,  71,  1160-1169,  1986. 

Nell  J.,  B.  J.  Wood,  T.  O.  Mason,  High-tempera- 
ture cation  distributions  in  Fe304-MgAl204- 
MgFe204-FeAl204  spinels  from 
thermopower  and  conductivity  measurements. 
Amer.  Mineral,  74,  339-351,  1989. 

O'Neill,  H.  St.  C,  W.  A.  Dollase,  and  C.  R.  Ross 
II,  Temperature  dependence  of  the  cation  dis- 
tribution in  nickel  aluminate  (NiAl204)  spinel: 
a  powder  XRD  study.  Submitted  to  Physics 
and  Chemistry  of  Minerals,  1991a. 

O'Neill,  H  St.  C,  H.  Annersten,  and  D.  Virgo  D., 
The  temperature  dependence  of  the  cation 
distribution  in  magnesioferrite  (MgFe204) 
from  powder  XRD  structural  refinements  and 
Mossbauer  spectroscopy.  Submitted  to  Amer. 
Mineral,  1991b. 

O'Neill,  H.  St.  C.  The  rates  of  cation  order- 
disorder  in  MgFe204  and  MgAl204  spinels. 
In  preparation,  1991c. 

O'Neill,  H.  St.  C,  and  A.  Navrotsky,  Simple 
spinels:  Crystallographic  parameters,  cation 
radii,  lattice  energies,  and  cation  distribution. 
American  Mineralogist,  68,  181-194,  1983. 

O'Neill,  H.  St.  C,  and  A.  Navrotsky,  Cation 
distributions  and  thermodynamic  properties 
of  binary  spinel  solid  soultions.  Amer.  Min- 
eral, 69,  733-753,  1984. 

Sawatsky,  G.A.,  F.  Van  der  Wande,  and  A.  H. 
Morrish,  Mossbauer  study  of  several  ferri- 
magnetic  spinels.  Physical  Review,  187, 747- 
757, 1969. 

Virgo,  D.,  S.  S.  Hafner,  Fe^+-Mg  order-disorder 
in  heated  orthopyroxenes.  Mineral  Soc.  Amer. 
Spec.  Paper  2,  67 -87,  1969. 

Wiles,  D.B.,  R.  A.  Young,  A  new  computer  pro- 
gram for  Rietveld  analysis  of  X-ray  powder 
diffraction  patterns.  Journal  App.  Cry  stall., 
14,  149-151,  1987. 


GEOPHYSICAL  LABORATORY 


101 


Crystallography  —  Mineral  Physics 


Predicted  High-Pressure  Mineral 

Structures 

with  Octahedral  Silicon 

Robert  M.  Hazen  and  Larry  W.  Finger 

Silicates  are  the  most  common  minerals 
on  the  earth's  surface,  and  they  probably 
dominate  throughout  the  Earth's  mantle. 
Many  hundreds  of  silicate  structures  have 
been  determined  and  catalogued  (e.g., 
Liebau,  1985),  but  only  about  50  different 
structures  account  for  the  vast  majority  of 
all  crustal  silicates  (Smyth  and  Bish,  1988). 
A  common  feature  of  all  these  low-pres- 
sure mineral  structures  is  the  presence  of 
silicon  cations  exclusively  in  4-coordina- 
tion  [lYISi  by  oxide  anions. 

High-pressure  experiments  demonstrate 
that  all  common  crustal  silicates  undergo 
phase  transitions  to  new  structures  with  6- 
coordinated  silicon,  [VI]Si,  at  pressures 
between  8  GPa  (for  pure  Si02)  to  about  30 
GPa,  which  corresponds  to  the  pressure  at 
the  top  of  the  lower  mantle.  Mineral  physi- 
cists identify  silicon  coordination  number 
as  a  major  crystal  chemical  difference  be- 
tween the  crust  and  lower  mantle:  silicon  is 
virtually  all  four  coordinated  above  about 
250  km,  but  is  entirely  six  coordinated 
below  670  km.  The  earth's  transition  zone, 
on  the  other  hand,  is  marked  by  the  appear- 
ance of  a  group  of  high-pressure  silicates 
with  both  [IVlSi  and  tVIlSi.  The  stability  of 
these  minerals  is  apparently  confined  to  a 


rather  narrow  pressure  range  from  approxi- 
mately 10-30  GPa.  Within  these  limits, 
however,  are  silicate  structures  of  remark- 
able complexity  and  great  topological  in- 
terest. 

The  objectives  of  this  review  are  to 
tabulate  all  known  high-pressure  silicate 
structures  with  six  coordinated  or  mixed- 
coordinated  silicon,  to  identify  crystal 
chemical  systematics  among  these  struc- 
tures, and  to  predict  additional  high-pres- 
sure silicate  structure  types. 

There  are  only  a  dozen  known  high- 
pressure  structures  with  Si06  polyhedra 
(Table  13).  These  silicates  can  be  divided 
conveniently  into  two  groups.  Above  about 
25  GPa,  corresponding  to  the  Earth's  lower 
mantle,  all  silicates  studied  to  date  are 
observed  to  transform  to  one  of  seven  dense 
structures,  in  which  all  Si  is  6-coordinated. 
These  structures  —  rutile,  perovskite,  il- 
menite,  hollandite,  calcium  ferrite, 
pyrochlore,  and  K2NiF4 — are  well  known 
room-pressure  topologies  for  transition 
metal  oxides.  In  the  high-pressure  silicate 
isomorphs  silicon  occupies  the  octahedral 
transition  metal  site,  while  other  cations 
may  adopt  six  or  greater  coordination. 

At  pressures  between  about  10  and  20 
GPa  (in  the  Earth's  transition  zone)  a  sec- 
ond group  of  silicates  forms  with  mixed  4- 
and  6-coordination.  These  phases  include 
silica-rich  modifications  of  the  well  known 
garnet,  pyroxene,  and  wadeite  structures, 
as  well  as  complex  new  magnesium-bear- 


102 


CARNEGIE  INSTITUTION 


Table  13.  Compositions  and  calculated  densities  of  high-pressure  silicates  with  Si06  octahedra. 


Composition  (Mineral  Name)        Structure  Type 

p  calc* 

References 

(g/cm3) 

A.  High-Pressure  phases 

with  Si06 

groups  only 

SiC>2  (Stishovite) 

Rutile 

4.29 

Hill  etal.  (1983) 

CaSi03 

Cubic  Perovskite 

4.25 

Mao  etal.  (1989) 

MgSi03 

Ortho  Perovskite 

4.10 

Horiuchiertf/.  (1987) 

MgSi03 

Ilmenite 

3.81 

Horiuchi  eet  al.  (1982) 

ZnSi03 

Ilmenite 

5.25 

Ito  and  Matsui  (1974) 

KAlSi308 

Hollandite 

3.91 

Yamadaera/.  (1984) 

BaAl2Si208 

Hollandite 

5.3 

Reid  and  Ringwood  (1969) 

CaAl2Si208 

Hollandite 

3.9 

Madon  era/.  (1989) 

NaAlSi04 

Calcium  Ferrite 

3.91 

Yamada etal.  (1983) 

Sc2Si2Ov 

Pyrochlore 

4.28 

Reid  etal.  (1977) 

In22Si207 

Pyrochlore 

6.34 

Reid  etal.  (1977) 

Ca2Si207 

K2NiF4 

3.56 

Liu  (1978) 

B.  High-Pressure  phases  with  SiC>6  + 

Si04  groups 

MgSi03  (Majorite) 

Garnet 

3.51 

Angela  al.  (1989) 

MnSi03 

Garnet 

4.32 

Fujino etal.  (1986) 

Na(Mgo.5Sio.5)Si206 

Pyroxene 

3.28 

Angel  etal.  (1988) 

K2Si409 

Wadeite 

3.09 

Swanson  and  Prewitt  (1983) 

Mgi4Si5024 

Anhydrous  Phase  B  3.44 

Finger  etal.  (1991) 

Mgi2SUOi9(OH)2 

Phase  B 

3.37 

Finger  etal.  (1991) 

*p  calc  =  density  calculated  from  unit-cell  parameters  at  room  pressure  and  temperature. 


ing  phases  designated  "phase  B"  and  "an- 
hydrous phase  B."  [A  third  group  of  room- 
pressure  [VllSi  silicates,  detailed  by  Finger 
and  Hazen  (1991)  but  not  considered  here, 
includes  silicon  phosphates  with  relatively 
open  framework  structures.] 

The  twelve  high-pressure  structures 
listed  in  Table  13  display  a  wide  range  of 
linkages  between  Si  octahedra  and  other 
polyhedra.  In  stishovite,  hollandite,  and 
pyroxene  a  combination  of  edge-  and  cor- 
ner-sharing is  observed,  but  in  phase  B  and 
anhydrous  phase  B  each  Si  octahedron 
shares  all  12  edges  with  adjacent  Mg  octa- 
hedra. In  pyrochlore,  garnet,  and  wadeite 
the  Si  octahedra  form  part  of  a  comer- 


linked  framework,  but  additional  cations  in 
eight  or  greater  coordination  share  edges 
and  faces  with  the  octahedra.  Ilmenite  pre- 
sents yet  a  different  topology,  with  unusual 
face  sharing  between  Mg  and  Si  octahedra, 
as  well  as  corner  and  edge  sharing.  Despite 
the  differing  polyhedral  linkages,  the  size 
and  shape  of  Si06  polyhedra  are  similar  in 
all  twelve  high-pressure  compounds.  Poly- 
hedral volumes  at  room  pressure,  for  ex- 
ample, vary  by  only  about  ±  4%  from  an 
average  7.67-A3  value.  All  Si  octahedra  are 
close  to  regular  (i.e.,  distortion  indices  are 
small)  relative  to  the  range  observed  for 
many  divalent  and  trivalent  cation  octahe- 
dra. The  observed  tendency  of  silicon  to 


GEOPHYSICAL  LABORATORY 

Table  14.  Predicted  lVI^Si  structures  that  conform  to  more  than  one  criterion. 


103 


Formula 

Structure 

1* 

2* 

3* 

4* 

5* 

(MgSi)02(OH)2 

Diaspore 

X 

X 

Ga^SiOg 

- 

X 

X 

Ga4Si7O20 

- 

X 

X 

MgioSi30i6 

Aerugite 

X 

X 

CaSi205 

Sphene 

X 

X 

MgSi(OH)6 

Stottite/Gibbsite 

X 

X 

BaSU09 

Benitoite 

X 

X 

Fe2Si05 

Pseudobrookite 

X 

X 

*  Criteria  for  predicting  tVIlSi  structures: 

1.  Edge-sharing  octahedral  chains 

2.  Germanate  isomorphs 

3.  Ti,  Mn,  and  Fe  oxides 

4.  Substitution  of  (Mg  +  Si)  for  2A1 

5.  System  Mg-Si-O-H 


adopt  highly  regular  coordination  leads  us 
to  conclude  that  silicon  will  not  readily 
adopt  exotic  5-  or  7-coordination  or  highly 
distorted  4-  or  6-coordination  groups  at 
high  pressure  in  either  crystalline  or  amor- 
phous condensed  phases. 

Five  systematic  relations  among  the 
structures  in  Table  1 3  can  be  used  to  predict 
other  possible  high-pressure  silicates  (see 
also  Table  14).  Each  of  these  criteria  can  be 
used  to  predict  other  potential  [VI]  Si  phases. 


Fig.  58.  Relationships  among  the  rutile  (A), 
IrSe2(B),  ramsdellite(C),  hollandite  (D), 
psiomelane  (E),  and  a  hypothetical  composite 
structure  (F),  after  Bursill  (1979). 


1.  Three  structure  types  (rutile, 
hollandite,  calcium  ferrite)  are  formed  from 
edge-sharing  chains  of  silicon  octahedra. 
This  relation  points  to  other  likely  structure 
types,  all  of  which  incorporate  edge-shar- 
ing octahedral  chains  linked  to  adjacent 
strips  by  corner  sharing,  as  systematized  by 
Wadsley  (1964),  Bursill  and  Hyde  (1972), 
and  Bursill  and  Hyde  (1979).  Rutile  has 
single  chains,  leading  to  1  x  1  square  chan- 
nels, while  hollandite  and  calcium  ferrite 
have  double  chains,  yielding  larger  chan- 
nels. Many  similar  octahedral  chain  struc- 
tures, such  as  ramsdellite  (1x2)  and 
psilomelane  (2  x  3)  are  also  known  (Fig. 
58),  and  each  of  these  could  provide  a 
topology  suitable  for  silicon  in  6-coordina- 
tion. 

2.  Nine  of  the  twelve  known  [VllSi 
high-pressure  structure  types  were  first 
synthesized  as  germinates  at  lower  pres- 
sures. A  systematic  search  of  the  Inorganic 
Crystal  Structure  Data  Base  (ICSD,  FIZ 
Karlsruhe  distributors)  for  germanates  with 
[  vrlGe  in  systems  containing  the  additional 
cations  Na,  K,  Mg,  Fe,  Ca,  Al,  Ti,  Si,  and  P 


104 


CARNEGIE  INSTITUTION 


revealed  25  structure  types,  only  nine  of 
which  have  known  silicate  analogs  (Finger 
andHazen,  1991).  A  number  of  these  com- 
pounds, including  Fe4Ge209,  FesGe30i8, 
CaGe205,Ca2Ge70i6,Ca4Ge30io(H20), 
and  K2BaGegOi8,  are  good  prospects  for 
high-pressure  silicate  analogs. 

3.  All  seven  high-pressure  tVIlSi  struc- 
tures without  tetrahedral  Si  are  isomorphs 
of  room-pressure  oxides  with  trivalent  or 
tetravalent  transition  metals  (Ti,  Mn,  or  Fe) 
in  octahedral  coordination.  Structures  of 
other  binary  oxides  with  octahedral  tita- 
nium, manganese,  or  iron  may  also  repre- 
sent possible  topologies  for  mantle  miner- 
als. Particularly  relevant  in  this  context  are 
the  structures  of  CaSi205  with  the  sphene 
structure,  Fe2Si05  or  Al2Si05  with  the 
pseudobrookite  structure,  and  CaSi409  with 
the  benitoite  structure. 

4.  High-pressure  ilmenite,  garnet,  and 
pyroxene  forms  of  magnesium-bearing  sili- 
cates are  all  related  to  room-pressure  phases 
by  the  substitution  of  octahedral  Mg  and  Si 
for  a  pair  of  aluminum  cations.  Similar 
substitutions  might  occur  at  high  pressure 
in  several  other  common  rock-forming 
minerals,  including  kyanite,  staurolite, 
pseudobrookite,  lawsonite,  cordierite, 
clinozoisite,  gibbsite,  and  diaspore.  Note 
that  this  substitution  scheme  will  not  work 
for  many  common  aluminum-bearing  min- 
erals with  mixed  4-  and  6-coordinated  alu- 
minum. The  substitution  in  muscovite 
[K[VI]Al2[IV](AlSi3)Oio(OH2)],  for  ex- 
ample, would  yield  the  magnesian  mica 
celadonite,K[VI](MgAl)[lV]Si4Oio(OH)2, 
in  which  all  Si  is  tetrahedrally  coordinated. 
Octahedral  Al,  thus,  must  constitute  more 


than  two-thirds  of  all  aluminum  to  produce 
a  [^1] Si  phase  by  the  substitution  2A1  — > 
(Mg  +  Si). 

5 .  Finger  and  Pre  witt  ( 1 990)  documented 
the  close  structural  relations  among  a  num- 
ber of  hydrous  and  anhydrous  magnesium 
silicates,  and  used  those  systematics  to 
propose  several  as  yet  unobserved  struc- 
tures, including  high-pressure  hydrous 
phases  with  octahedral  silicon.  They  rec- 
ognized that  several  known  phases,  includ- 
ing chondrodite,  humite,  forsterite,  phase 
B,  and  anhydrous  phase  B,  are  members  of 
a  large  group  of  homologous  magnesium 
silicates  that  can  be  represented  by  the 
general  formula: 

m[Mg4«+2[][VJSi2«08«(OH)4]Mg6/2+4- 
2mod(n,2)[YIlSin+mod(n,2)OSn+4 

where  mod{n,2)  is  the  remainder  when  n  is 
divided  by  2.  Finger  and  Prewitt  (1990) 
examined  cases  where  n  =  1 ,2,3,4,  °°  and  m 
=  1,2,  oo.  Structures  with  octahedral  silicon 
result  for  all  cases  where  m  is  not  infinity. 
Of  special  interest  is  the  proposed  structure 
of  "superhydrous  phase  B,"  a  compound 
predicted  by  the  logical  progression  from 
Mgi4Si5024  (anhydrous  phase  B)  to 
Mgi2Si40i9(OH)2  (phase  B)  to 
MgioSi30i4(OH)4.  Gasparik  (1990)  sug- 
gested that  an  as  yet  unanalyzed  hydrous 
magnesium  silicate  synthesized  at  1 8.6  GPa 
and  1600  °C  possesses  this  structure,  and 
further  studies  on  that  material  are  in 
progress. 

Several  structure  types  appear  to  follow 
two  of  the  five  very  different  systematic 
trends.    These  structures,  therefore,  de- 


GEOPHYSICAL  LABORATORY 


105 


mand  further  study.  Of  special  interest  to 
earth  scientists  are  CaSi20s  with  the  titanite 
structure,  Fe2SiOs  with  the  pseudobrookite 
structure,  and  Mg  ioSi30 16  with  the  aerugite 
structure.  Each  of  these  phases,  or  their 
isomorphs  with  other  cations  replacing  Ca, 
Mg,  and  Fe,  might  be  represented  in  the 
Earth's  mantle.  In  fact,  Stebbins  and 
Kanzaki  (1991)  mention  the  existence  of 
titanite-type  CaSi20s  in  some  of  their  run 
products,  though  identification  of  this  phase 
was  provisional. 

Also  worthy  of  further  study  are  the 
proposed  hydrous  phases  MgSi02(OH)2 
and  MgSi(OH)6,  which  are  isomorphs  of 
diaspore  and  stottite,  respectively.  Such  li- 
nen phases  would  be  expected  to  occur 
only  locally  in  the  earth's  deep  interior,  but 
their  presence,  integrated  over  the  Earth's 
volume,  could  represent  a  major  respository 
of  water. 

Most  common  rock-forming  cations, 
including  Na,  Mg,  Fe,  Ca,  Mn,  Al,  Ti,  and 
Si,  are  small  enough  to  fit  into  the  tetrahe- 
dral  or  octahedral  interstices  of  a  close- 
packed  oxygen  net.  However,  the  presence 
of  many  other  cations,  including  H,  B,  K, 
Rb,  Pb,  rare  earths,  and  U,  could  disrupt  the 
close-packed  array  and  lead  to  other,  as  yet 
unrecognized,  structure  types.  The  gal- 
lium and  barium  silicates,  Ga4SiOs, 
Ga4SiyO20,  and  BaSi409,  are  just  three  of 
the  dozens  of  possible  new  [VlJSi  struc- 
tures likely  to  be  observed  as  high-pressure 
investigations  extend  beyond  the  traditional 
rock-forming  elements.  These  structures 
are  not  likely  to  play  a  significant  role  in 
mantle  mineralogy,  but  they  will  provide  a 
more  complete  understanding  of  the  crys- 
tal chemistry  of  octahedral  silicon. 


Conclusions 

Is  the  earth's  deep  interior 
mineralogically  simple?  Are  there  only  a 
few  dominant  structure  types,  or  is  there  an 
unrecognized  complexity  in  the  crystal 
chemistry  of  octahedral  silicon?  There  are 
hundreds  of  different  crustal  silicates  with 
PYlSi,  but  only  a  dozen  high-pressure  [VI]Si 
structures  have  been  produced.  This  dis- 
parity may  reflect  the  relatively  small  num- 
ber of  high-pressure  studies,  but  it  also 
arises,  at  least  in  part,  from  the  nature  of 
oxygen  packing.  Numerous  crustal  sili- 
cates, from  the  commonest  minerals  quartz 
and  feldspar  to  the  dozens  of  zeolites  and 
other  framework  silicates,  possess  open, 
low-density  topologies  with  correspond- 
ingly loose  packing  of  oxygen.  There  are 
no  obvious  limits  to  the  variety  of  silicates 
based  on  irregular  oxygen  packing. 

Volume  constraints  imposed  by  high 
pressure,  however,  favor  structures  with 
approximately  close-packed  oxygens. 
These  restrictions  on  anion  topology  re- 
duce the  number  of  possible  cation  con- 
figurations as  well,  and  it  is  thus  antici- 
pated that  the  number  of  different  struc- 
tural topologies  in  the  earth's  deep  interior 
will  be  much  smaller  than  at  the  surface. 
Dense,  close-packed,  and  for  the  most  part 
high-symmetry  structures,  such  as  those 
represented  by  the  seven  known  topologies 
with  all  tvrlSi,  will  predominate.  Never- 
theless, within  these  restrictions  there  ex- 
ists opportunity  for  considerable  structural 
diversity  owing  to  three  factors  -  reversible 
phase  transitions,  cation  positional  order- 
ing, and  modularity,  particularly  based  on 


106 


CARNEGIE  INSTITUTION 


different  close-packed  layer  stacking  se- 
quences. This  potential  diversity  is  only 
hinted  at  by  the  known  phases. 

Several  of  the  known  high-pressure 
types,  including  perovskite,  K2NiF4,  and 
pyrochlore,  can  adopt  numerous  structural 
variants  based  on  slight  changes  in  lattice 
distortions  and  cation  distribution.  The 
perovskite  structure,  in  particular,  can  un- 
dergo dozens  of  phase  transitions  based  on 
octahedral  tilting,  cation  ordering,  cation 
displacements,  and  anion  defects  (Megaw, 
1973;  Hazen,  1988).  We  must  study  pro- 
posed mantle  phases  at  the  appropriate 
conditions  of  pressure  and  temperature  to 
document  the  equilibrium  structural  varia- 
tions. 

Close  packing  of  oxygen  leads  to  modu- 
lar structures,  with  certain  features  (e.g., 
edge-sharing  octahedral  chains  of  rutile; 
the  double  chains  of  hollandite;  the  comer- 
sharing  octahedral  sheets  of  perovskite;  the 
face-sharing  topology  of  ilmenite)  that  can 
link  together  in  many  ways  to  form  ordered 
superstructures  of  great  complexity.  Such 
complexity  was  recognized  by  Wadsley 
( 1 964)  and  Bursill  and  Hyde  ( 1 979)  in  their 
descriptions  of  modular  rutile-hollandite- 
Ga203  structures,  and  it  is  realized  in  the 
homologous  series  including  phase  B,  an- 
hydrous phase  B,  and  several  other  struc- 
tures. Phase  B,  for  example,  is  based  on 
oxygen  close  packing,  yet  it  has  40  inde- 
pendent atoms  in  its  asymmetric  unit  to 
yield  one  of  the  most  complex  ternary 
silicates  yet  described.  Variations  on  the 
phase  B  structure  could  be  based  on  chang- 
ing the  relative  number  and  position  of  the 
two  different  structural  layers,  by  introduc- 


ing other  types  of  layers,  or  by  staggering 
layers  to  produce  clino-  and  ortho-type 
structures  as  observed  in  other  close -packed 
systems,  for  example,  in  the  biopyriboles 
as  describedbyThompson(1978)  and  Smith 
(1 982).  The  structure  could  be  further  com- 
plicated by  element  ordering  among  the  17 
different  cation  sites  as  Al,  Fe  Ti,  Mn,  and 
other  elements  enter  the  structure  in  a  natu- 
ral environment . 

The  study  of  octahedrally-coordinated 
silicon  is  still  in  its  infancy,  yet  clear  trends 
are  beginning  to  emerge  from  the  scattered 
data  on  diverse  structures  and  composi- 
tions. It  is  now  evident  that  while  silicate 
perovskite  may  be  the  predominant  phase 
in  the  Earth's  lower  mantle,  a  number  of 
other  dense  silicate  phases  will  compete  for 
elements  such  as  K,  Ba,  Ca,  and  Al.  It 
appears  that  the  earth's  transition  zone  will 
display  the  varied  mineralogy  of  mixed 
tVIlSi  and  PV] Si  silicates,  including  some 
of  the  most  complex  structures  known  in 
the  mineral  kingdom.  And  it  is  certain  that 
a  detailed  understanding  of  the  mantle  must 
await  studies  of  these  fascinating  phases  at 
temperatures  and  pressures  appropriate  to 
the  Earth's  dynamic  interior. 


References 

Angel,  R.  J.,  T.  Gasparik,  N.  L.  Ross,  L.  W.  Finger, 
C.  T.  Prewitt,  and  R.  M.  Hazen,  A  silica-rich 
sodium  pyroxene  phase  with  six-coordinated 
silicon,  Nature,  335,  156-158,  1988. 

Angel,  R.  J.,  L.  W.  Finger,  R.  M.  Hazen,  M. 
Kanzaki,  D.  J.  Weidner,  and  R.  C.  Liebermann, 
Structure  and  twinning  of  single-crystal 
MgSiC>3  garnet  synthesized  at  17  GPa  and 
1800°C,  Amer.  Mineral,  74,  509-512,  1989. 

Bursill,  L.A.  and  B.  G.  Hyde,  Structural  relation- 
ships between  P-gallia,  rutile,  hollandite, 


GEOPHYSICAL  LABORATORY 


107 


psilomelane,  ramsdellite,  and  gallium  titanite 
type  structures,  Acta  Cryst.,  B35,  530-538, 
1979. 
Bursill,  L.  A.,  and  B.  G.  Hyde,  Rotation  faults  in 
crystals,  Nature  (London) Phys.  Sci.,  240, 122- 
124,  1972. 
Finger,  L.  W.,  and  R.  M.  Hazen,  Crystal  chemistry 
of  six-coordinated  silicon:  A  key  to  under- 
standing the  Earth ' s  deep  interior,  Acta  Cryst., 
A47,  1991,  in  press. 
Finger,  L.W.  and  C.  T.  Prewitt,  Predicted  compo- 
sitions for  high-density  hydrous  magnesium 
silicates,  Geophys.  Res.  Lett.,  16,  1395-1397, 
1990. 
Finger,  L.  W.,  R.  M.  Hazen,  and  C.  T.  Prewitt, 
Crystal  structures  of  Mgi2Si40i9(OH)2  (Phase 
B)  and  Mgi4Si5024  (Phase  AnhB),  Amer. 
Mineral,  76,  1-7,1991. 
Fujino,  K.,  H.  Momoi,  H.  Sawamoto,  and  M. 
Kumazawa,  Crystal  structure  and  chemistry  of 
MnSi03  tetragonal  garnet,  Amer.  Mineral., ' 
71,  781-785,  1986. 
Gasparik,  T.  Phase  relations  in  the  transition  zone, 

/.  Geophys  Res.,  95,  15751-15769,  1990. 
Hazen,  R.  M.,  Perovskites,  Sci.  Amer.,  74-8 1 ,  June 

1988. 
Hill  R.J.,  M.  D.  Newton,  and  G.  V.  Gibbs,  A 
crystal  chemical  study  of  stishovite,  /.  Sol. 
State  Chem.,  47,  185-200,  1983. 
Horiuchi,  H.,  M.  Hirano,  E.  Ito,  and  Y.  Matsui, 
MgSiC>3  (ilmenite-type):  Single  crystal  x-ray 
diffraction  study,  Amer.  Mineral,  67, 788-793, 
1982. 
Horiuchi,  H.,  I.  Eiji,  andD.  J.  Weidner,  Perovskite- 
type  MgSiC>3:  Single-crystal  x-ray  diffraction 
study,  Amer.  Mineral,  72,  357-360,  1987. 
Ito,  E.,  and  Y.  Matsui,  High-pressure  synthesis  of 
ZnSiC>3  ilmenite,  Phys.  Earth  Planet.  Int.,  9, 
344-352,  1974. 
Liebau,  F.,  Structural  Chemistry  of  Silicates, 

Springer,  New  York,  1985. 
Liu,  L.,  High  pressure  Ca2SiC>4,  the  silicate 
K2NiF4-isotype  with  crystalchemical  and  geo- 
physical implications,  Phys.  Chem.  Minerals, 
3,  291-299,  1978. 
Madon,  M,  J.  Castex,  and  J.  Peyronneau,  A  new 
aluminocalcic  high-pressure  phase  as  a  pos- 
sible host  of  calcium  and  aluminum  in  the 
lower  mantle,  Nature,  342,  422-425,  1989. 
Mao,  H.K.,  L.  C.  Chen,  R.  J.  Hemley,  A.  P. 
Jephcoat,  Y.  Wu,  and  W.  A.  Bassett,  Stability 
and  equation  of  state  of  CaSi03-perovskite  to 
134  GPdiJ.Geophys.Res.,94, 17,889-17,894, 
1989. 
Megaw,  H.  D.,  Crystal  Structures:  A  Working 
Approach,  Philadelphia,  W.  B.  Saunders,  1973. 
Reid,  A.  F.,  and  A.  E.  Ringwood,  Six-coordinate 
silicon:  High  pressure  strontium  and  barium 
aluminosilicates  with  the  hollandite  structure, 
J.  Solid  State  Chem.  1,  6-9,  1969. 


Reid,  A.  F.,  C.  Li,  and  A.  E.  Ringwood,  High- 
pressure  silicate  pyrochlores,  SC2S12O7  and 
In2Si207,  Solid  State  Chem.,  20,  219-226, 
1977. 

Smith,  J.  V.,  Geometrical  and  Structural  Crystal- 
lography, Wiley,  New  York,  1982. 

Smyth,  J.  R.,  and  D.  L.  Bish,  Crystal  Structures 
and  Cation  Sites  of  the  Rock-Forming  Miner- 
als, Allen  and  Unwin,  Winchester,  Massachu- 
setts, 1988. 

Stebbins,  J.  F.,  and  M.  Kanzaki,  Local  structure 
and  chemical  shifts  for  six-coordinated  silicon 
in  high-pressure  mantle  phases,  Science,  251, 
294-298,  1991. 

Swanson,  D.K.,  and  C.  T.  Prewitt,  The  crystal 
structure  of  K2SiVISi3IVC>9,  Amer.  Mineral, 
65,581-585,1983. 

Thompson,  J.  B.,  Biopyriboles  and  polysomatic 
series,  Amer.  Mineral,  63,  239-249,  1978. 

Wadsley,  A.  D.,Non-Stoichiometric  Compounds, 
L.  Mandelcoin,  ed.,  p.  Ill,  New  York,  Aca- 
demic Press,  1964. 

Yamada,  H.,  Y.  Matsui,  andE.  Ito,  Crystal-chemi- 
cal characterization  of  NaAlSi04  with  the 
CaFe204  structure,  Mineral.  Mag.,  47,  177- 
181, 1983. 

Yamada,  H.,  Y.  Matsui,  andE.  Ito,  Crystal-chemi- 
cal characterization  of  KAlSi3Us  with  the 
hollandite  structure,  Mineral.  J.  (Japan),  12, 
29-34,  1984. 


SlMULTANOUS  HlGH  P-^ 

Diffraction  Measurements 

OF  (Fe,Mg)Si03-PEROVSKITE 

and  (Fe,Mg)0  Magnesiowustite: 

Implications  for  Lower 

Mantle  Composition 

Yingwei  Fei,  Ho-Kwang  Mao,  Russell  J. 
Hemley,  and  Jinfu  Shu 

(Fe,Mg)Si03-perovskite  and  (Fe,Mg)0 
magnesiowustite  are  likely  stable  phases  in 
the  Earth's  lower  mantle.  The  thermal  prop- 
erties of  those  phases  are  of  critical  impor- 
tance for  constraining  the  composition  of 
the  lower  mantle.  In  this  paper,  we  report 
simultaneous  high  P-T  synchrotron  x-ray 
diffraction  measurements  of  (Fe,Mg)Si03 


108 


CARNEGIE  INSTITUTION 


>n     Mw 


C/) 

c 

CD 


Mw200 


Mw220 


Au200 


Au111 

V 


'•'.  Au311 

\ 

'•  Au220    Mw311     :".* 


19.8  GPa,  310  K 


/ 


Energy,  keV 


Fig.  59.  Representative  energy-dispersive  x-ray  diffraction  spectra  (20  =  15°)  of  magnesiowustite 
(Mg.6Fe.4)0  at  three  different  P-T  conditions. 


perovskite  and  (Fe,Mg)0  magnesiowustite. 
These  data  provided  the  first  direct  mea- 
surements of  the  effect  of  pressure  on  the 
thermal  expansivity  of  these  minerals.  In 
the  discussion  we  also  examine  the  impli- 
cations of  these  results  for  the  composition 
and  mineralogy  of  the  lower  mantle. 


Experimental  Methods 

The  samples  used  in  the  experiments 
are  (Feo.4Mgo.6)0  magnesiowustite,  syn- 
thesized from  mixtures  of  Fe2C>3  and  MgO 


at  a  temperature  of  1573  K  and  an  oxygen 
fugacity  (fen)  of  lO10-8  bar  (Rosenhauer  et 
a/.,  1976),  and  (Feo.iMgo.9)Si03  perovskite, 
synthesized  from  synthetic  pyroxene  by 
laser  heating  at  40  GPa  in  diamond-anvil 
cell  (Mao  etal,  1991). 

The  experiments  were  carried  out  using 
a  high-temperature  diamond-anvil  cell, 
made  of  inconel,  and  synchrotron  radia- 
tion. A  nickel  alloy  (Rene  41)  gasket  with 
thickness  of  200  Jim  was  preindented  to  a 
pressure  of  17  GPa.  The  powder  sample 
was  placed  in  the  250-|im-diameter  sample 
chamber,  filling  less  than  one-third  of  the 


GEOPHYSICAL  LABORATORY 


109 


volume.  Gold  foil  and  ruby  grains  were 
placed  in  the  sample  chamber  as  pressure 
calibrants  at  high  temperature  (Anderson 
et  al. ,  1 989)  and  at  room  temperature  (Mao 
etal. ,  1 986),  respectively.  The  sample  cham- 
ber was  then  filled  with  neon  gas  at  200 
MPa  in  a  high-pressure  gas-loading  device, 
and  sealed  at  a  pressure  of  2  GPa.  The  neon 
served  as  a  quasi-hydrostatic  pressure  trans- 
mitting medium  over  the  pressure  range  of 
measurement. 

The  sample  was  heated  with  an  external 
platinum-wire  resistance  heater  (Mao  et 
al,  1991).  The  heater  was  placed  on  the 
cylinder  of  the  cell.  Temperatures  were 
measured  with  a  chromel-alumel  thermo- 
couple, while  pressures  were  determined 
by  measuring  the  lattice  parameter  of  gold. 
During  the  experiment,  pressure  usually 
decreases  with  increasing  temperature  at 
the  rate  of  about  5  GPa/100  K. 

Polychromatic  (white)  wiggler  synchro- 
tron x-radiation  at  the  National  Synchro- 
tron Light  Source,  Brookhaven  National 
Laboratory  was  used  for  the  energy-disper- 
sive x-ray  diffraction  measurements.  The 
diffraction  data  were  collected  with  an  in- 
trinsic germanium  solid-state  detector  at  a 
fixed  20  angle  of  15°(±  0.005°).  The  energy 
was  calibrated  by  using  known  energies  of 
x-ray  emission  lines  (Ka  and  Kp)  of  Mn, 
Cu,  Rb,  Mo,  Ag,  Ba,  and  Tb.  The  20  angle 
was  calibrated  by  collecting  the  diffraction 
pattern  of  platinum.  The  experimental  con- 
ditions were  optimized  by  considering  the 
x-ray  beam  size,  data  collecting  time,  and 
slit  size  for  the  detector.  With  a  60-|im 
beam  spot,  a  complete  diffraction  pattern 
of  magnesiowustite  with  reasonable  peak 
counts  can  be  obtained  in  about  five  min- 


utes. Figure  59  shows  three  typical  spectra 
of  magnesiowustite  with  internal  standard 
gold  collected  at  different  P-T  conditions. 
All  diffraction  patterns  show  at  least  four 
sharp  diffraction  lines,  111,  200,  220,  and 
311,  of  magnesiowustite.  For  theperovskite, 
the  diffraction  was  carried  out  using  mono- 
chromatic synchrotron  x-ray  and  film  meth- 
ods (Mao  et  al.,  1991).  These  techniques 
were  used  because  of  the  need  for  high 
resolution  to  resolve  multiplets  in  the  dif- 
fraction patterns  arising  from  the  or- 
thorhombic  distortion  of  the  perovskite. 


Results  and  Discussions 

Lattice  parameters  of  the 
magnesiowustite  were  determined  from 
diffraction  lines  111,  200,  220,  and  311  by 
using  a  peak-fitting  program.  For 
perovskite,  diffraction  peaks  (mainly  020, 
112,  and  200)  were  measured  by  both 
manual  and  computerized  film-reading 
methods.  The  experimental  results  are  plot- 
ted in  Figures  60  and  61 .  The  uncertainties 
in  pressure  result  from  the  measurements 
of  lattice  parameter  of  gold  which  was  used 
as  an  internal  high-pressure  calibrant  at 
high  temperature.  The  error  of  ±  0.0015  A 
in  the  lattice  parameter  of  gold,  which  is  the 
typical  measurement  uncertainty  in  the  ex- 
periments, corresponds  to  about  ±  0.30 
GPa  in  pressure. 

To  construct  a  P-V-T  equation  of  state 
from  a  combination  of  1-bar  thermal  ex- 
pansion data,  300-K  compression  data,  and 
simultaneous  high  P-T  volume  data,  we 
express  pressure  in  a  general  form 

/W)  =  /W300K)  +  />*,  (1) 


110 


CARNEGIE  INSTITUTION 


o 
o 

CO 
CL 

P 
> 


~l 

10  15 

Pressure,  GPa 


Fig.  60.  Volume  at  high  P-T  relative  to  the  300  K 
isotherm.  Experimental  data  are  from  Mao  et  al 
[1991]  (crosses);  Knittle  et  al,  [1986]  (solid 
squares);  and  Wang  et  al,  [1991]  (open  squares). 
The  lines  are  calculated  isotherms. 


o 
o 

CO 

a." 
> 


2.5-1 


2.0_ 


1.5- 


1.0- 


0.0- 


"3 

S93"*- 


600 


0,      „,,      4-  _^«76 


565 

5*3 


*"1      «• 

323  -}-31Q 


500K 


300K 


1^ 

20 


—T 
25 


- 1- 
30 


Pressure,  GPa 


Fig.  61.  Volume  at  high  P-T  relative  to  the  300  K 
isotherm.  Experimental  data  are  from  Fei  et  al 
[1991b]  (crosses);  and  Suzuki  [1975]  (solid 
squares).  The  lines  are  calculated  isotherms. 

where  P(V,300K)  can  be  expressed  by  the 
standard  third-order  Birch-Murnaghan  form 


P  =  h 

2 


4#-(#I-^-^fe 


(2) 


where  Kto  and  KTq  are  the  isothermal  bulk 
modulus  and  its  pressure  derivative  at  room 
temperature,  respectively.  The  thermal  pres- 
sure can  be  modeled  in  two  ways  as  dis- 
cussed below. 

By  using  the  thermodynamic  identity, 
the  thermal  pressure  can  be  calculated  by 
integrating  the  oKv  i.e., 


/>th  = 


f 

/300K 


aKjiXT 


(3) 


where  a  is  the  thermal  expansivity  and  KT 
is  isothermal  bulk  modulus  at  T  and  V  of 
interest.  To  parameterize  the  thermal  pres- 
sure, we  start  with  an  assumption  that  (BK^ 
dT)p  is  independent  of  temperature,  and 
obtain  (dKflT),  =  -2 .7 '(±0.3)  x  102  GPa/K 
for  magnesiowiistite  and  (dK^dT^p  =  - 
6.3(±0.5)  x  10 2  GPa/K  for  perovskite  by 
fitting  our  P-V-T  data  to  the  thermal  pres- 
sure model.  A  detailed  discussion  of  these 
results  is  given  in  Mao  et  al  ( 1 99 1 )  and  Fei 
etal  (1991b). 

The  Anderson-Griineisen  parameter  ST 
is  commonly  used  to  measure  the  change  in 
thermal  expansivity  at  high  P-T.  It  is  de- 
fined by  (Anderson,  1967) 


<5r  = 


3lna 


_      1 


lMT\ 


[dlnV  It      aKA  dT 


(4) 


The  parameters,  5r  for  magnesiowiistite 
and  perovskite  derived  from  our  data  are 
listed  in  Table  15.  They  decrease  with  in- 
creasing temperature  below  the  Debye  tem- 
perature and  approach  a  constant  value  at 
high  temperature.  The  ST  values  for 
magnesiowiistite  and  perovskite  are  4.3 
and  6.5,  respectively,  above  the  Debye 
temperatures. 


GEOPHYSICAL  LABORATORY 


111 


Table  15.  Temperature  variation  of  some  thermodynamic  parameters  for  magnesiowiistite  and 
perovskite 


Perovskite 

Magnesiowiistite 

T,K 

a(106) 

KT,  GPa 

ocKt 

dr 

q 

a(106)  tfr,GPa 

aKj 

Sr 

<7 

300 

21.90 

260.9 

57.15 

11.02 

7.48 

31.32 

157.0 

49.16 

5.49 

1.97 

400 

28.94 

254.6 

73.68 

8.55 

5.28 

35.87 

154.3 

55.33 

4.88 

1.57 

500 

33.00 

248.3 

81.94 

7.69 

4.50 

38.47 

151.6 

58.31 

4.63 

1.40 

600 

35.90 

242.0 

86.86 

7.25 

4.10 

40.31 

148.9 

60.00 

4.50 

1.30 

700 

38.24 

235.7 

90.11 

6.99 

3.86 

41.79 

146.2 

61.07 

4.42 

1.24 

800 

40.28 

229.4 

92.40 

6.82 

3.70 

43.07 

143.5 

61.79 

4.37 

1.19 

900 

42.15 

223.1 

94.04 

6.70 

3.58 

44.25 

140.8 

62.28 

4.34 

1.15 

1000 

43.92 

216.8 

95.21 

6.62 

3.50 

45.35 

138.1 

62.60 

4.31 

1.13 

1100 

45.62 

210.5 

96.02 

6.56 

3.45 

46.41 

135.4 

62.81 

4.30 

1.11 

1200 

47.27 

204.2 

96.52 

6.53 

3.42 

47.43 

132.7 

62.92 

4.29 

1.09 

1300 

48.89 

197.9 

96.74 

6.51 

3.41 

48.44 

130.0 

62.94 

4.29 

1.09 

1400 

50.48 

191.6 

96.71 

6.51 

3.41 

49.43 

127.3 

62.90 

4.29 

1.08 

1500 

52.05 

185.3 

96.45 

6.53 

3.43 

50.40 

124.6 

62.78 

4.30 

1.08 

1600 

53.62 

179.0 

95.96 

6.57 

3.47 

51.37 

121.9 

62.60 

4.31 

1.09 

1700 

55.17 

172.7 

95.26 

6.61 

3.52 

52.33 

119.2 

62.36 

4.33 

1.10 

The  thermal  contribution  can  also  be 
calculated  by  the  Mie-Gruneisen  relation, 

^h  =  Qect,  eD)  -  £(3ook,  eD)] 

where  E(T,QD)  is  the  harmonic  internal  en- 
ergy calculated  from  either  a  Debye  model 
or  a  single  Einstein  oscillator  model 
(Zharkov  and  Kalinin,  1971),  which  are 
equivalent  in  the  high-temperature  limit. 
The  Debye  temperature  6D  and  Griineisen 
parameter  /are  considered  to  be  functions 
of  volume  only:  y=  -dkiOJdlnV ,  with  the 
volume  dependence  of  y  given  by  q  =9lny 
/3lnK  The  model  parameters,  0D,  y,  and  q, 
can  be  obtained  by  fitting  the  experimental 
P-V-T  data  to  the  Mie-Griineisen  equation 
of  state. 

There  is  some  uncertainty  in  the  Debye 
temperature  Qm  for  perovskite.  As  a  result 
of  the  non-linear  character  of  the  fit,  the 
parameters  are  correlated  and  there  is  a 
trade-off  among  the  best-fit  values.  Figure 
61  illustrates  the  trade-off  between  y0  and  q 
for  values  of  Bm  ranging  from  725  to  1025 


K  for  the  silicate  perovskite.  The  circles 
indicate  the  best  fit  to  the  experimental  data 
when  both  y0  and  q  are  simultaneously 
optimized.  Notably,  the  q  of  3 . 3  (at  y0 = 1 .70 
and  6m  =  725  K)  for  perovskite  obtained  in 
this  analysis  is  considerably  higher  than 
many  other  materials  {e.g.,  q  =  0-1  is  com- 
monly found  in  shock- wave  studies).  How- 
ever, the  high  q  value  is  consistent  with 
high  <57-value  derived  from  an  independent 
method  of  analysis  (Mao  et  al. ,  1 99 1 )  using 
the  thermodynamic  relation 


q  =  8T  +  1 


KT 


(6) 


when  (dlnCydlnVOy,  =  0,  which  is  valid  at 
high  temperature. 


The  trade-off  between  the  values  of  K. 


TO 


and  Kn'  that  fit  the  static  compression  data 
has  been  examined  previously  (Mao  et  al., 
1 99 1 ).  That  work  showed  that  the  assump- 
tion that  Kn*  =  4  yields  Kw  =  26 1  GPa.  If  a 
lower  value  for  the  bulk  modulus  is  as- 
sumed, a  higher  value  for  K^  is  required  to 
be  consistent  with  the  static  compression 


112 


CARNEGIE  INSTITUTION 


data.  For  example,  adopting  Kw  =  247 
GPa,  as  obtained  from  Brillouin  scattering 
measurements  at  zero  pressure  by  Yeganeh- 
Haeri  et  al.  (1989),  requires  Kn*  =  5.5. 
Because  of  the  identity  relating  K^  and  q 
[equation  (6)],  it  is  useful  to  consider  the 
effect  of  a  higher  assumed  value  for  A^'  on 
the  thermal  properties.  A  fit  to  the  experi- 
mental data  with  the  assumption  of  Kn'  = 
5.5  yields  q  =  1. 8  at  y0 =1.70,  with  0D0  =  725 
K  (Fig.  62).  The  difference  in  densities 
calculated  for  perovskite  at  high  P  and  T 
with  Kw'  varied  over  this  range  increases 
with  pressure:  for  example,  at  60  GPa  K^ 
=  4  gives  0.8%  higher  density  than  that 
calculated  with  K^  =  5.5. 

The  parameter  values  for  calculating 
the  P-V-T  relations  of  perovskite  and 
magnesiowiistite  are  summarized  in  Table 
16.  Isotherms  calculated  with  the  Mie- 
Griineisen  relation  in  the  range  of  the  x-ray 
diffraction  measurements  are  shown  in  Fig- 
ures 62  and  63.  For  magnesiowiistite,  both 
equations  (3)  and  (5)  both  predict  consis- 
tent P-V-T  relations  up  to  140  GPa  and 
3000  K(Fei  etaL,  1991b).  For  perovskite, 
however,  equation  (3)  predicts  higher  vol- 
ume at  high  P-T  than  equation  (5)  (about 
0.5%  higher  at  60  GPa  and  2000  K).  These 
differences  can  result  in  large  uncertainties 
in  the  lower  mantle  composition. 


Fig.  62.  The  trade-off  between  y0  and  q  at  various 
given  Debye  temperature  0D0.  The  circles  indi- 
cate the  best  fit  to  the  experimental  data  when 
both  y0  and  q  are  simultaneously  optimized.  The 
solid  lines  are  obtained  when  Kw  =  261  GPa  and 
AYo  =  4  are  used.  The  dashed  lines  are  obtained 
when  K-ro  -  247  GPa  and KTq  =  5.5  are  used.  (See 
text  for  discussion.) 


A  detailed  comparison  of  the  densities 
calculated  from  P-V-T  equations  of  state 
presented  above  and  that  determined  by 
seismology  (e.g.,  PREM,  Dziewonski  and 
Anderson,  1981)  indicates  that  a  large 
perovskite  component  relative  to 
magnesiowusite  with  an  upper  mantle  Fe/ 
Mg  ratio  (Ring  wood,  1 975)  matches  PREM 
to  within  1  %  throughout  the  entire  lower 


Table  16.  Parameters  of  the  thermal  equations  of  state  of  perovskite  and  magnesiowiistite 


Parameters 


(Fe,Mg)SiC>3-perovskite 


(Fe,Mg)0-magnesiowustite 


Vo,  cm3/mol 

24A6+0MXFe 

11.25+1.02XFe 

Km  GPa 

261(4) 

160-7.5XFe 

Kto 

4 

4 

(dK^dT)P,  GPa/K 

-6.3(5)  x  lO-2 

-2.7(5)  x  lO"2 

(dKiftnv,  GPa/K 

-2.5(3)  x  10-2 

-0.2(2)  x  lO"2 

5r 

6.5(5) 

4.3(5) 

Oqo,  K 

725(25) 

500(20) 

7o 

1.70(5) 

1.50(5) 

q 

3.3(5) 

1.1(5) 

GEOPHYSICAL  LABORATORY 


113 


mantle  (see  Hemley  et  al.,  1991).  This 
implies  a  silica  enrichment  in  the  lower 
relative  to  the  upper  mantle.  The  conclu- 
sion is  dependent  on  the  geotherms  and  the 
thermoelastic  parameters.  Figure  63  shows 
the  Fe/(Mg+Fe)  and  Si/Mg  ratios  for  a  best 
fit  of  density  between  the  mineralogical 
model  and  PREM  in  the  top  portion  of  the 
lower  mande  (pressure  range  of  24-60  GPa). 
The  composition  trade-off  between  tem- 
perature and  density  is  demonstrated  by 
varying  the  adiabatic  temperature  at  670 
km  from  1 800  K  to  2000  K  and  its  gradient 
from  7  K/GPa  to  10  K/GPa.  Abest  fit  of  X 


Fe 


=  0.1  andXs.=  1.0  is  obtained  for  the  2000- 
K  adiabat.  The  Xsi  value  is  not  as  well 
constrained  by  the  density  analysis.  How- 
ever, the  bulk  modulus,  the  slope  of  density 
profile  divided  by  density,  provides  a  con- 
straint on  Si/Mg.  The  minimum  of  XSi  is 
always  close  to  1  (i.e.,  pure  perovskite 
composition)  for  the  equation  of  state  of 
perovskite  used  in  this  study.  The  mini- 
mum shifts  to  more  olivine-enriched 
composition  if  the  volume  dependence  of 
the  thermal  expansivity  of  perovskite  de- 
creases. 

Recently,  we  also  showed  that  the  cal- 
culated phase  relations  in  the  MgO-FeO- 
Si02  system  under  lower  mantle  condi- 
tions are  sensitive  to  the  volume  depen- 
dence of  the  thermal  expansion  coefficient 
of  component  phases,  with  the  value  for  the 
(Fe,Mg)Si03  perovskite  being  especially 
significant  (Fei  and  Hemley,  1991).  We 
find  that  <5rof  at  least  6  (q  of  2.9)  is  required 
to  match  the  observed  phase  boundary  for 
Mg2Si04  (spinel)  =  MgSi03  (perovskite) 
+  MgO  (periclase).  A  high  value  for  5j  is 
also  found  to  be  consistent  with  the  experi- 
mental phase  equilibrium  data  (Ito  and 
Takahashi,  1989)  and  element  partitioning 


0.20 


0.15 


0.10 


0.05  - 


decreasing  q 


q=2.3 
18O0K 

10K/GPa    9=3.3 
1800K 


(7=3.3 
2000K 
10K/GPa 


g=3.3 
10K/GPa    1800K 

7K/GPa 


0.00 


0.5 


0.6 


0.7  0.8 

Xsi 


0.9  1 .0 


Fig.  63.  Misfits  between  the  mineralogical  model 
and  seismic  data  for  density  as  functions  of  Xpe  and 
Xsi.  The  Fe/Mg  and  Si/Mg  ratios  for  a  best  fit  of 
density  are  calculated  by  varying  the  adiabatic 
temperature  at  670  km  from  1 800  K  to  2000  K  and 
its  gradient  from  7  K/GPa  to  10  K/GPa.  The  effect 
of  the  volume  dependence  q  for  perovskite  is  also 
indicated  by  the  arrow. 


data  (Fei  et  al.,  1991a)  in  the  system 
Mg2Si04-Fe2Si04.  The  stability  field  of 
perovskite  shifts  to  higher  pressure  and  the 
Mg-rich  region  with  decreasing  &r  value. 
A  lower  value  of  &r  would  give  rise  to  two 
complete  solid  solutions  between  spinel 
and  magnesiowustite  and  between 
perovskite  and  magnesiowustite,  in  contra- 
diction to  experimental  results  in  this  sys- 
tem. Increasing  XFe  and  temperature,  and 
decreasing  <5r,  are  shown  to  decrease  sig- 
nificantly the  stability  field  of  perovskite. 
The  primary  conclusions  of  this  study 
concern  the  top  of  the  lower  mantle  (from 
670  -  1000  km  depth).  Comparison  be- 
tween the  extrapolated  equation  of  state 
data  and  seismic  results  at  greater  depths 
indicates  that  the  compositional  arguments 
put  forth  here  do  apply  to  the  bottom  of  the 
lower  mantle;  that  is,  the  conclusion  that  a 
lower  mantle  assemblage  dominated  by 


114 


CARNEGIE  INSTITUTION 


perovskite  with  an^Fe  of -0.1  is  consistent 
with  the  seismic  data.  The  main  uncertain- 
ties in  this  lower  mantle  composition  model 
may  rise  from  both  temperature  and  pres- 
sure extrapolations  in  the  thermal  pressure 
models.  The  two  models  for  perovskite 
used  in  the  analyses  result  in  0.5%  differ- 
ence in  desity  which  accounts  for  about  2% 
difference  in  iron  content  by  fitting  to  the 
PREM  densities.  Uncertainties  in  Kj  can 
result  in  more  significant  changes  in  the 
composition  model,  especially  to  the  Si/ 
Mg  ratio  argument.  More  accurate  models 
of  the  composition  and  mineralogy  in  the 
deeper  portions  of  the  mantle  would  re- 
quire measurements  at  higher  P-T  condi- 
tions. 


References 

Anderson,  O.  L.,  Equation  for  thermal  expansivity 
in  planetary  interiors,  J.  Geophys.  Res.,  72, 
3661,  1967. 

Anderson,  O.  L.,  D.  G.  Isaak,  and  S.  Yamamoto, 
Anharmonicity  and  the  equation  of  state  for 
gold,  /.  Appl.  Phys.,  65,  1534-1543,  1989. 

Dziewonski,  A.  M.,  and  D.  L.  Anderson,  Prelimi- 
nary reference  earth  model,  Phys .  Earth  Planet. 
Interiors,  25,  297-356,1981. 

Fei,  Y.,  andR.  J.  Hemley,  Stability  of  (Fe,Mg)Si03- 
perovskite  in  the  lower  mantle,  Geophys.  Res. 
Lett.,  in  press,  1991. 

Fei,  Y.,  H.  K.  Mao,  and  B.  O.  Mysen,  Experimen- 
tal determination  of  element  partitioning  and 
calculation  of  phase  relations  in  the  MgO- 
FeO-Si02  system  at  high  pressure  and  high 
temperature,  /.  Geophys.  Res.,  96,  2157- 
2169,  1991a. 

Fei,  Y.,  H.  K.  Mao,  J.  Shu,  and  J.  Hu,  P-V-T 
equation  of  state  of  magnesiowustite 
(Mgo6Feo4)0,  Phys.  Chem.  Miner.,  submit- 
ted, 1991b. 


Hemley,  R.  J.,  Y.  Fei,  andH.  K.  Mao,  Constraints 
on  lowermantle  composition  from  P-V-T  mta- 
surements  of  (Fe,Mg)SiC>3-perovskite  and 
(Fe,Mg)0,  High  Pressure  Research  in  Min- 
eral Physics:  Application  to  Earth  and  Plan- 
etary Science,  edited  by  Y.  Syono  and  M.  H. 
Manghnani  eds.,  submitted,  1991. 

Ito,  E.,  and  E.  Takahashi,  Postspinel  transforma- 
tions in  the  system  Mg2Si04-Fe2Si04  and 
some  geophysical  implications,  /.  Geophys. 
Res.,  94,  10,637-10,646,  1989. 

Knittle,  E.,  R.  Jeanloz,  and  G.  L.  Smith,  The 
thermal  expansion  of  silicate  perovskite  and 
stratification  of  the  Earth's  mantle,  Nature, 
319,  214-216,  1986. 

Mao,  H.  K.,  J.  Xu,  and  P.  M.  Bell,  Calibration  of 
the  ruby  pressure  gauge  to  800  kbar  under 
quasihydrostatic conditions,/.  Geophys. Res., 
97,4673-4676,1986. 

Mao,  H.  K.,  R.  J.  Hemley,  Y.  Fei,  J.  F.  Shu,  L.  C. 
Chen,  A.  P.  Jephcoat,  Y.  Wu,  and  W.  A. 
Bassett,  Effect  of  pressure,  temperature,  and 
composition  on  lattice  parameters  and  density 
of  (Fe,Mg)SiC>3-perovskites  to  30  GPa,  J. 
Geophys.  Res.,  in  press,  1991. 

Ring  wood,  A.  E.,  Composition  and  Petrology  of 
the  Earth's  Mantle,  pp.  618,  McGraw-Hill, 
New  York,  1975. 

Rosenhauer,  M.,  H.  K.  Mao,  and  E.  Woermann, 
Compressibility  of  magnesiowustite 
(Feo  4Mgo  6)0  to  264  kbar,  Carnegie  Inst. 
Washington  Year  Book  ,  75,  513-515,  1976. 

Suzuki,  I.,  Thermal  expansion  of  periclase  and 
olivine  and  their  anharmonic  properties,  J. 
Phys.  Earth,  23,  145-159,  1975. 

Wang,  Y.,  D.  J.  Weidner,  R.  C.  Liebermann,  X. 
Liu,  J.  Ko,  M.  T.  Vaughan,  Y.  Zhao,  A. 
Yeganeh-Haeri,  and  R.  E.  G.  Pacalo,  Phase 
transition  and  thermal  expansion  of  MgSi03 
perovskite,  Science,  251,  410-413,  1991. 

Yeganeh-Haeri,  A.,  D.  J.  Weidner,  and  E.  Ito, 
Elasticity  of  MgSiC>3  in  the  perovskite  struc- 
ture, Science,  243,  787-789,  1989. 

Zharkov,  V.  N.,  and  V.  A.  Kalinin,  Equation  of 
State  for  Solids  at  High  Pressures  and 
Temperatures,  pp.  257,  Consultants  Bureau, 
New  York,  1971. 


GEOPHYSICAL  LABORATORY 


115 


High-Pressure  Crystal  Chemistry 

of  Iron-Free  wadsleyite, 

P-MG2S1O4 

Jinmin  Zhang,  Robert  M.  Hazen,  and 
Jaidong  Ko* 

Wadsleyite,  P-(Mg,Fe)2Si04,  may  be 
the  most  abundant  mineral  in  the  upper 
mantle  between  400  and  550  km,  and  the 
phase  transformation  of  olivine  to 
wadsleyite  may  explain  the  400-km  dis- 
continuity. Since  this  phase  was  first  found 
to  occur  at  high  pressure,  much  work  has 
been  done  to  relate  its  crystal  structure  and 
properties  to  geophysically  important  prob- 
lems. For  example,  the  01  and  even  02 
positions  have  been  considered  to  be  pos- 
sible sites  for  protonation;  therefore,  the 
wadsleyite  phase  may  be  a  repository  of 
water  in  the  mantle  (Smith,  1987;  Downs, 
1989). 

The  wadsleyite  structure  has  three  crys- 
tallographically  non-equivalent  octahedral 
sites,  M7,  Ml  and  M3.  Wadsleyite  in  the 
mantle  is  believed  to  contain  -10%  Fe2+, 
which  partitions  in  the  order  M3>M1>M2 
(Finger  et  al.,  in  preparation).  This  prefer- 
ence should  depend  on  the  volume,  the 
distortion  and  crystal  field  stabilization 
energy  (CFSE)  of  each  site.  The  character- 
istics of  these  cation  sites  in  an  iron-free 
phase  at  high  pressure  are  important  to 
understanding  the  distribution  of  Fe2+  in 
wadsleyite  in  the  mantle.  In  this  paper,  we 
report  the  result  of  structure  refinements  of 
P-Mg2Si04  at  five  pressures  up  to  4.84 


*  Department  of  Earth  and  Space  Science,  State 
University  of  New  York  at  Stony  Brook 


GPa,  with  particular  attention  to  the  prop- 
erties of  the  octahedral  sites.  This  study, 
together  with  work  in  progress  on  Fe-con- 
taining  wadsleyites,  will  help  to  define  the 
properties  of  this  phase  in  the  mantle. 

The  single  crystals  used  in  this  study 
were  synthesized  by  Jaidong  Ko  at  the 
High-Pressure  Laboratory  of  the  State  Uni- 
versity of  New  York  at  Stony  Brook.  The 
sample  was  produced  at  a  pressure  of  about 
16  GPa  and  a  temperature  of  1400°C. 

A  subequant  crystal  0.06  x  0.12  x  0.13 
mm  in  size  was  selected,  and  wasexamined 
optically  and  by  x-ray  diffraction  at  room 
conditions.  Three  intensity  data  sets  were 
collected  and  several  unit-cell  parameter 
measurements  were  made  on  this  crystal 
before  it  was  crushed  when  trying  to  in- 
crease pressure  after  the  run  at  2.88  GPa. 
Another  crystal  0.04  x  0.06  x  0.06  mm  in 
size  was  used  for  higher  pressure  measure- 
ments. 

The  selected  crystal  and  a  small  piece  of 
fluorite  crystal  were  mounted  in  a  dia- 
mond-anvil cell  designed  for  single -crystal 
x-ray  diffraction  studies  (Hazen  and  Fin- 
ger, 1982).  The  fluorite  crystal  was  used  as 
an  internal  standard  of  pressure,  following 
the  method  of  Hazen  and  Finger  ( 1 98 1 ).  An 
Inconel  750X  gasket  with  0.40  mm  diam- 
eter hole  was  centered  over  one  0.60-mm 
diamond  anvil;  the  crystal  was  then  affixed 
to  the  anvil  face  inside  the  hole  with  a  small 
dot  of  the  alcohol  insoluble  fraction  of 
vaseline  petroleum  jelly.  A  mixture  of  4:1 
methanol rethanol  was  used  as  the  hydro- 
static pressure  medium. 

All  x-ray  measurements  except  for  the  0 
and  1.16  GPa  intensity  data  collections 
were  performed  with  a  Picker  automated 


116 


CARNEGIE  INSTITUTION 


Table  17.  Crystallographic  data  for  iron-free  wadsleyite 


Pressure,  GPa 

Paramter 

0 

1.16 

1.81 

2.88 

4.84 

crystal  size,  mm 

0.06x0.12x0.13 

. 

_ 

_ 

0.04x0.06x0.06 

IH>  cm-1 

11.16 

11.25 

11.29 

11.35 

11.48 

Range  of  TO) 

0.92-0.94 

0.92-0.94  0.93-0.94 

0.93-0.94 

0.95-0.96 

RinP 

0.060 

0.069 

0.057 

0.057 

0.078 

Number  of  data 

307 

294 

303 

286 

239 

R  J3)  all  data 

0.056 

0.055 

0.056 

0.059 

0.060 

/?(4)  all  data 

0.089 

0.094 

0.082 

0.072 

0.114 

Number  F0>2gf 

236 

216 

244 

242 

157 

Rw 

0.055 

0.054 

0.055 

0.057 

0.058 

R 

0.068 

0.069 

0.061 

0.057 

0.074 

T  is  transmission  factor.  R{nt  is  residual  for  internal  agreement  of  symmetry 
equivalent  reflections.  Rw  =  (Luj(F0  -  Fc)2/ZwF02f-5  R  =  IXiF0\  -  IFCII/ZIIF0II 


four-circle  diffractometer  with  filtered  Mo- 
Koc  radiation  (A  =  0.7107  A).  The  intensity 
data  at  0  and  1.16  GPa  were  collected  with 
a  Huber  four-circle  diffractometer  with 
graphite  monochromated  Mo-Koc  radiation 
(X  =  0.7093  A),  but  the  cell  parameters 
measured  with  the  Picker  diffractometer 
were  used  for  the  structure  refinements  for 
the  purpose  of  consistency. 

Unit-cell  parameters  were  measured 
with  the  procedure  of  King  and  Finger 
(1979),  whereby  several  reflections  are 
measured  in  eight  equivalent  orientations. 
The  range  of  20  for  all  reflections  was  1 8- 
31°  in  order  to  avoid  systematic  errors  that 
result  from  comparing  angular  data  from 
different  ranges  (Swanson  et  al.,  1985). 

Intensities  were  measured  for  all  acces- 
sible reflections  in  a  hemisphere  of  recipro- 
cal space  with  sin0/A  <  0.7.  Corrections 
were  made  for  Lorentz  and  polarization 
effects,  crystal  absorption  by  the  diamond 
and  beryllium  components  of  the  pressure 


cell.  Digitized  step  data  were  integrated  by 
the  method  of  Lehmann  and  Larsen  ( 1 974). 
Refinement  conditions,  refined  isotropic 
extinction  coefficients,  refined  structural 
parameters,  and  isotropic  temperature  fac- 
tors are  given  in  Tables  17  and  18. 

As  observed  by  Hazen  et  al.  ( 1 99 1 ),  the 
c  axis  is  the  most  compressible,  with  pc  = 
2.32(4)  x  1 0-3  GPa- l .  The  a  and  b  axes  have 
compressibilities  of  1.72(14)  x  10-3  and 
1.71(3)  x  10"3  GPa-1,  respectively,  —  al- 
most completely  identical  with  the  results 
of  Hazen  et  al.  (1991),  who  reported  the 
value  of  1.73(3)xl0-3  for  the  a  and  b  axes 
and  2.39(3)  x  10"3  for  the  c  axis.  The 
stiffness  of  a  and  b  relative  to  c,  according 
to  Hazen  et  al.  (1991),  results  from  the 
pseudo-layering  of  the  structure,  with  Mg- 
octahedral  layers  parallel  to  (001 ),  and  cross 
linking  by  Si207  tetrahedral  pairs. 

The  pressure-volume  data  were  fitted 
with  program  VOLFIT  to  a  Birch- 
Mumaghan  equation  of  state  with  4  as  the 


GEOPHYSICAL  LABORATORY 


117 


Table  18.  Positional  and  equivalent  isotropic  thermal  parameters. 


Pressure,  GPa 

Parameter 

0 

1.16 

1.81 

2.88 

4.84 

Extn. 

Coef(10-4) 

0.37(5) 

0.38(5) 

0.26(5) 

0.31(6) 

0.26(6) 

Ml,  x,y,z 

0 

0 

0 

0 

0 

B 

0.62(10) 

0.66(11) 

0.55(8) 

0.45(7) 

0.54(18) 

M2,* 

0 

0 

0 

0 

0 

y 

1/4 

1/4 

1/4 

1/4 

1/4 

2 

0.9696(5) 

0.9694(6) 

0.9704(4) 

0.9697(4) 

0.9687(9) 

B 

0.52(8) 

0.52(8) 

0.43(6) 

0.49(6) 

0.47(13) 

M3,* 

1/4 

1/4 

1/4 

1/4 

1/4 

y 

0.1265(3) 

0.1264(3) 

0.1267(2) 

0.1267(2) 

0.1263(5) 

z 

1/4 

1/4 

1/4 

1/4 

1/4 

B 

0.91(6) 

0.91(7) 

0.70(5) 

0.70(5) 

0.58(10) 

Si,  x 

0 

0 

0 

0 

0 

y 

0.1203(2) 

0.1203(2) 

0.1200(2) 

0.1199(1) 

0.1202(3) 

z 

0.6166(2) 

0.6165(3) 

0.6167(2) 

0.6168(2) 

0.6172(4) 

B 

0.55(5) 

0.54(5) 

0.41(4) 

0.37(4) 

0.36(7) 

Ol.z 

0 

0 

0 

0 

0 

y 

1/4 

1/4 

1/4 

1/4 

1/4 

z 

0.2173(10)  0.2181(10) 

0.2187(8) 

0.2204(7) 

0.2210(14) 

B 

0.9(2) 

0.8(2) 

0.5(1) 

0.6(1) 

0.2(2) 

02,x 

0 

0 

0 

0 

0 

y 

1/4 

1/4 

1/4 

1/4 

1/4 

z 

0.7158(9) 

0.7167(9) 

0.7159(8) 

0.7158(7) 

0.7199(14) 

B 

0.5(1) 

0.5(2) 

0.5(1) 

0.5(1) 

0.3(2) 

03,* 

0 

0 

0 

0 

0 

y 

0.9889(5) 

0.9895(5) 

0.9900(4) 

0.9899(4) 

0.9900(8) 

z 

0.2564(8) 

0.2557(8) 

0.2554(6) 

0.2553(5) 

0.2538(15) 

B 

0.9(1) 

0.6(1) 

0.54(9) 

0.40(8) 

0.8(2) 

04,* 

0.2605(10)  0.2597(12) 

0.2629(10) 

0.2608(9) 

0.258(2) 

y 

0.1226(3) 

0.1225(4) 

0.1218(4) 

0.1227(2) 

0.1211(7) 

z 

0.9923(4) 

0.9933(5) 

0.9933(4) 

0.9925(3) 

0.9936(7) 

B 

0.66(7) 

0.71(8) 

0.63(7) 

0.58(6) 

0.8(1) 

assumed  pressure  derivative  (K9).  The  bulk 
modulus  is  162(6)  GPa  if  Vo  is  not  con- 
strained, comparable  to  the  value  160(3) 
GPa  reported  by  Hazen  et  al.  (1991). 

The  structure  refinement  of  Fe-free 
wadsleyite  by  Finger  et  al.  (in  preparation) 
was  used  as  the  initial  model  in  the  present 
work  except  that  isotropic  temperature  fac- 
tors were  used  in  place  of  anisotropic  ones 
in  their  study.  The  space  group,  Imma,  was 
confirmed  by  the  structure  refinement.  The 
scattering  factor  curves  for  Mg,  Si,  and  O 


are  those  of  neutral  atoms  in  International 
Tables  for  X-ray  Crystallography  (1974). 

The  RFINE6  program  (Finger  and 
Prince,  1975)  was  used  for  the  structure 
refinements.  The  calculation  converged 
quickly  to  the  R  values  listed  in  Table  17. 
The  extinction  coefficients,  atomic  coordi- 
nates, and  isotropic  temperature  factors  are 
listed  in  Table  18. 

Polyhedral  volumes  were  calculated 
using  the  program  VOLCAL,  written  by 
Finger  (in  Hazen  and  Finger,  1982);  qua- 


118 


CARNEGIE  INSTITUTION 


Table  19.  Polyhedral  parameters  for  iron-free  wadsleyite 

Pressure,  GPa 
Parameter 

0                  1.16              1.81                2.88 

4.84 

Ml,  Volume 

QE* 

AV 

VS 

M2,  Volume 

QE 

AV 

VS 

M3,  Volume 

QE 

AV 

VS 

Si,  Volume 

QE 

AV 

VS 

01,  VS 

02,  VS 

03,  VS 

04,  VS 


11.648(38) 
1.0053(10) 
15.83 
2.17 

11.880(38) 
1.0056(46) 
19.58 
2.09 

12.026(21) 
1.0070(18) 
23.11 
2.06 

2.276(10) 
1.0037(40) 
14.63 
3.84 

1.98 
2.01 
1.95 
2.04 


11.519(41) 
1.0045(11) 
13.62 
2.21 

11.779(42) 
1.0061(51) 
21.76 
2.11 

11.887(22) 
1.0065(20) 
21.74 
2.10 

2.294(11) 
1.0034(43) 
13.68 
3.79 

1.99 
2.02 
1.96 
2.04 


11.514(35) 
1.0049(10) 
15.52 
2.21 

11.959(35) 
1.0057(39) 
19.63 
2.07 

11.807(18) 
1.0061(15) 
20.10 
2.12 

2.254(9) 
1.0033(35) 
12.58 
3.90 

2.02 
2.03 
1.98 
2.08 


11.439(28) 
1.0046(8) 
14.43 
2.24 

11.768(30) 
1.0056(34) 
19.79 
2.12 

11.785(16) 
1.0058(13) 
18.79 
2.13 

2.254(8) 
1.0040(30) 
15.90 
3.89 

2.02 
2.05 
1.99 
2.09 


10.966(73) 
1.0044(21) 
13.35 
2.39 

11.554(70) 
1.0065(79) 
23.01 
2.18 

11.591(33) 
1.0057(31) 
18.72 
2.19 

2.298(20) 
1.0026(71) 
10.50 
3.79 

2.04 
2.07 
2.02 
2.09 


*QE  and  AV  stand  for  the  quadratic  elongation  and  angle  variance  parameters,  respec- 
tively, of  Robinson  et  al.  (1969).  VS  stands  for  the  valence  sum  of  Brown  (1981). 


dratic  elongations  and  angular  variances 
were  calculated  according  to  the  defini- 
tions of  Robinson  et  al.  (1971);  the  bond 
valence  sums  are  after  Brown  (1981).  The 
results  are  listed  in  Table  19. 

The  Si04  tetrahedral  volume  does  not 
decrease  significantly  with  pressure,  typi- 
cal of  all  high-pressure  studies  below  10 
GPa.  Of  the  three  Mg-0  octahedra,  M3  is 
the  largest  and  most  distorted  at  room  pres- 
sure (see  Table  19);  as  pressure  increases,  it 
becomes  smaller  and  perhaps  more  regular. 
Changes  of  the  Ml  and  M2  octahedra  are 
less  well  defined  (Fig.  64).  The  Ml  and  M2 
octahedra  become  smaller  with  increasing 


pressure,  but  there  is  little  change  in  the 
distortion. 

Least-square  fit  of  volume-pressure  data 
of  the  three  polyhedra  gives  average 
compressibilities  of  8.7(11),  4.3(11),  and 
6.9(6)  x  10'3  GPa"1,  respectively,  forM7, 
M2,  and  M3  sites,  corresponding  to  bulk 
moduli  of  116(15),  234(59),  and  145(12) 
GPa. 

Zhang  et  al.  (1991)  demonstrated  that 
the  unit-cell  volume  of  an  isostructural 
series  is  approximately  linearly  related  to 
the  bond  length  of  a  specific  cation-anion 
bond  as  long  as  other  bond  lengths  do  not 
change.  The  compression  of  wadsleyite  fits 


GEOPHYSICAL  LABORATORY 


119 


11.80 


V=  11.677(30)  -0.1 01  (13)P 

K  =  116(15)  GPa 
* 


i  i  i  i — i  i  i — i  i  i  i  i  i  i  i  i  i  i  i — i — i — i  i  i 


V=1 1.92(3)  -0.051(13)P 
I        K  =  234(59)  GPa 


11.60 

11.50 

£<      12.00 


i   i  i — i  i  i  i  i  i  i 


i  i  l  i — i  i  i  i — i— i- 


V=  11.998(16)  -0.083(7)P 
K  =  145(12)  GPa 


2.0  3.0  4.0 

Pressure,  GPa 


5.0 


Fig.  64.  Variations  of  the  volumes  of  the  three  M- 
O  octahedra  with  pressure. 

into  this  specification  because  the  Si-0 
bond  length  is  virtually  constant.  Figure 
65  shows  the  variation  of  the  unit-cell  vol- 
ume with  the  average  of  the  three  M-0 
bond  lengths;  the  linear  relationship  is  a 
clear  indication  that  the  contraction  of  the 
unit  cell  is  largely  due  to  the  shortening  of 
the  M-O  bonds.  This  linear  relationship  is 
also  true  when  Fe2+  substitutes  for  Mg2+  in 
the  wadsleyite  structure  and  thus  increases 
the  average  M-0  bond  length.  In  Fig.  66, 
using  the  data  of  Finger  et  al.  (in  prepara- 
tion), the  unit-cell  volume  of  wadsleyite 
with  different  Fe2+  contents  is  plotted  vs. 
the  average  M-0  bond  length.  Again  it 


520 


2.045         2.055         2.065         2.075         2.085 

Average  M-0  bond  length,  A 

Fig  65.  Linear  relationship  between  the  unit-cell 
volume  and  the  average  of  the  three  M-O  bond 
lengths. 


co 
© 

E 

O 

> 

© 

o 

i 

■*± 

"c 

ID 


■ 

546 

- 

Fe25 

•  S 

544 

• 

542 

Fe08v/ 

'         Fe16 

■ 

540 

•  yr 

• 

rqci 

x^» 

Te00 

2.078      2.081      2.084     2.087     2.090     2.093 

Average  M-0  bond  length,  A 

Fig  66.  Unit-cell  volume  as  a  function  of  the 
average  M-0  bond  length.  The  increases  of  the 
average  M-O  bond  length  and  the  unit-cell  volume 
are  due  to  the  substitution  of  Fe2+  for  Mg2+. 


suggests  that  the  increase  or  decrease  of  the 
unit-cell  volume  is  due  to  the  increase  or 
decrease  of  the  M-0  bond  lengths. 
We  reach  several  conclusions 
(1).  Measurements  of  unit-cell  param- 
eters at  10  pressures  give  the  axial 
compressibilities  of  wadsleyite,  which  are 
1.72(14),  1.71(3)  x  10-3,  and  2.32(4)  x  10- 
3  GPa-1,  for  a,  by  and  c,  respectively.  As- 
suming a  Birch- Murnaghan  equation  of 
state  with  K'  =  4,  the  bulk  modulus  is 
162(6)  GPa. 


120 


CARNEGIE  INSTITUTION 


(2).  The  atomic  coordinates,  bond 
lengths,  bond  angles,  and  polyhedral  vol- 
umes of  the  cation  sites  are  documented. 
The  bulk  moduli  of  the  three  octahedra  are 
116(15),234(59),andl45(12)GPaforM7, 
M2,  and  Mi,  respectively.  M3  is  the  octa- 
hedron for  which  the  volume  vs.  pressure 
data  points  are  the  least  scattered  and  is  also 
the  one  which  seems  to  become  more  regu- 
lar with  increasing  pressure.  The  other  two 
octahedra,  especially  M2 ,  have  bulk  moduli 
with  large  errors,  and  no  significant  varia- 
tion in  distortion  is  observed. 

(3).  The  unit-cell  volume  is  linearly 
related  to  the  average  M-0  bond  length,  an 
indication  that  the  decrease  of  the  unit-cell 
volume  is  due  to  the  shortening  of  the  M-0 
bonds. 


References 

Brown,  I.  D.,  The  bond-valence  method:  an  em- 
pirical approach  to  chemical  structure  and 
bonding,  in  Structure  and  Bonding  in  Crys- 
tals, O'Keefe  and  Navrotsky,  eds.,  2,  1-30, 
Academic  Press,  Boston,  1981. 

Downs,  J.  W.,  Possible  sites  for  protonation  in  (i- 
Mg2Si04  from  an  experimentally  derived  elec- 
trostatic potential,  Amer.  Mineral,  74,  1 124- 
1129,1989. 

Finger,  L.  W.,  andE.  Prince,  A  system  of  Fortran 
IV  computer  programs  for  crystal  structure 
computations,  US  National  Bureau  of  Stan- 
dard Technical  Note  854,  Washington  DC, 
1975. 

Hazen,  R.  M.,  and  L.  W.  Finger,  Calcium  fluoride 
as  an  internal  pressure  standard  in  high  pres- 
sure/high-temperature crystallography,/.  Ap- 
plied Cry  stallogr.,  14,  234-236,  1981. 

Hazen,  R.  M.,  and  Finger,  L.  W.,  Comparative 
Crystal  Chemistry,  Wiley,  New  York,  1982. 

Hazen,  R.  M.,  J.  Zhang,  and  J.  Ko,  Effects  of  Fe/ 
Mg  on  the  compressibility  of  synthetic 
wadsleyite:  0-(Mg \.x^x)2^0a  (jc<0.25), 
Phys.  Chem.  Minerals,  77,  416-419,  1991. 

International  Tables  of  X-ray  Crystallography, 
Vol.  IV,  Kynock  Press,  Birmingham,  1974. 


King,  H.  E.,  and  L.  W.  Finger,  Diffracted  beam 
crystal  centering  and  its  application  to  high- 
pressure  crystallography,  /.  Applied 
Crystallogr.,  12,  374-378,  1979. 

Lehmann,  M.  S.,  and  M.  K.  Larsen,  A  method  for 
location  of  the  peaks  in  step-scan-measured 
Bragg  reflections,  Acta  Cry  stallogr.,  A30, 580- 
584,  1974. 

Robinson,  K.,  G.  V.  Gibbs,  and  P.  H.  Ribbe, 
Quadratic  elongation:  A  quantitative  measure 
of  distortion  in  coordination  polyhedra,  Sci- 
ence, 772,567-570,  1971. 

Smith,  J.  R.,  p-Mg2SiC>4:  a  potential  host  for 
water  in  the  mantle?  Amer. Mineral.,  72, 1051- 
1055,  1987. 

Swanson,  D.  K.,  D.  J.  Weidner,  C.  T.  Prewitt,  and 
J.  J.  Kandelin,  Single-crystal  compression  of 
Y-Mg2Si04,  EOS,  66,  370,  1985. 

Zhang,  J.  M,  D.  N.  Ye,  and  C.  T.  Prewitt,  Rela- 
tionship between  the  unit-cell  volumes  and 
cation  radii  of  isostructural  compounds  and 
the  additivity  of  the  molecular  volumes  of 
carbonates,  Amer.  Mineral.,  76, 100-105, 1991. 


Phase  Transitions  in 
Framework  Minerals 

David  Palmer 

It  is  now  clear  that  certain  phase  transi- 
tions previously  dismissed  as  being  "subtle" 
phenomena,  may  in  fact  produce  dramatic 
anomalies  in  the  physical  properties  of 
minerals  and  significantly  modify  their  ther- 
modynamic behavior.  These  phase  transi- 
tions, which  typically  involve  displacive 
distortions  of  the  crystal  lattice  or  cation 
ordering  effects,  are  common  in  most  rock- 
forming  minerals  that  exist  in  the  Earth's 
crust.  They  are  also  expected  to  occur  in 
mantle  phases. 

Unlike  heterogeneous  reactions  be- 
tween crystallographically  unrelated  phases 
(which  are  more  usually  studied  by  earth 
scientists),  the  influence  of  a  displacive  or 
order/disorder  phase  transition  is  not  con- 


GEOPHYSICAL  LABORATORY 


121 


fined  to  the  phase  boundaries  of  an  equilib- 
rium system,  but  is  significant  at  pressures 
and  temperatures  far  below  the  transition 
point  itself.  As  a  consequence,  stability 
relations  between  mineral  assemblages  may 
be  perturbed  throughout  P-T  space. 


Quantitative  Analysis  of 
Mineral  Behavior 

Most  of  the  breakthroughs  in  the  study 
of  phase  transitions  have  come  from  solid- 
state  physics,  the  original  motivation  hav- 
ing been  to  relate  anomalies  in  physical  and 
thermodynamic  properties  to  changes  at  a 
crystal  structural  level.  From  a  substantial 
body  of  work  on  simple  crystal  structures, 
came  a  number  of  "mean  field"  theories,  all 
relating  to  the  macroscopic  behavior  of 
crystals.  These  concern  those  phase  transi- 
tions which  involve  the  lowering  of  crystal 
symmetry,  such  that  a  relationship  between 
the  symmetries  of  "high"  and  "low"  phases 
is  maintained  (i.e.,  a  supergroup-subgroup 
relation).  Most  displacive  and  order/disor- 
der phase  transitions  fit  into  this  category. 

Macroscopic  theories  based  on  ideas 
initially  propounded  by  Landau  (Landau 
and  Lifshitz,  1980)  have  been  used  exten- 
sively in  rationalizing  the  temperature  de- 
pendence of  crystal  behavior.  The  funda- 
mental starting  point  for  these  theories  is 
the  concept  of  a  macroscopic  order  param- 
eter, 2,  which  measures  the  progress  of  the 
phase  transition,  such  that  Q  =  0  in  the  high- 
symmetry  phase,  and  0<Q<1  in  the  low- 
symmetry  phase.  This  might  relate,  for 
example,  to  the  ordering  of  magnetic  spins, 
positions  of  certain  sets  of  cations  such  as 


Al  and  Si,  or  to  a  prevailing  lattice  distor- 
tion. Landau's  original  postulate  was  that 
the  energy  lowering  due  to  the  high  -  low 
transition,  the  excess  free  energy  of  the 
low-symmetry  phase,  could  be  represented 
as  a  power  series  in  Q.  From  this  potential, 
it  is  possible  to  derive  the  temperature 
dependence  of  the  order  parameter,  in  terms 
of  a  critical  exponent,  p.  It  also  becomes 
possible  to  relate  the  excess  thermody- 
namic properties — heat  capacity,  entropy, 
enthalpy  etc.  —  to  the  order  parameter  and 
hence  to  the  progress  of  the  phase  transi- 
tion. Further  developments  of  Landau 
Theory  allow  the  behavior  of  excess  physi- 
cal properties  to  be  related  to  the  order 
parameter  and  the  thermodynamic  proper- 
ties. 

The  Landau  approach,  by  relating  all 
excess  physical,  thermodynamic  and  struc- 
tural properties  to  the  macroscopic  order 
parameter  considerably  simplifies  the  de- 
scription of  phase  transition  behavior  and 
provides  for  a  detailed  understanding  of 
more  complex,  mineralogical  systems. 
Many  minerals  undergo  more  than  one 
phase  transition,  which  may  lead  to  seem- 
ingly very  complicated  behavior:  phase 
transitions  in  the  same  material  are  rarely 
independent  of  each  other.  In  Landau 
Theory,  these  effects  are  predicted  in  terms 
of  coupling  between  the  various  order 
parameters,  according  to  strict  symmetry 
rules. 

The  aim  of  this  particular  study  is  to 
increase  our  understanding  of  phase-tran- 
sition-related mineral  behavior,  focusing 
on  two  less  well  studied  areas,  (1)  high- 
temperature  thermodynamic  properties,  and 
(2)  high-pressure  structural  behavior. 


122 


CARNEGIE  INSTITUTION 


Framework  minerals  provide  the  "model 
systems"  for  this  work:  these  are  the  most 
abundant  materials  at  the  Earth's  surface; 
their  interconnected  topologies  ensure  long 
correlation  lengths  throughout  the  crystal 
structure,  and  account  for  the  many 
displacive  and  cation-ordering  phase  tran- 
sitions which  occur  with  increasing  tem- 
perature and  pressure.  In  this  context  there- 
fore, "mean  field"  theories  such  as  Landau 
Theory  are  particularly  relevant. 


Order  I  Disorder  Relations  in  Anorthite 

Feldspars  dominate  the  mineralogy  of 
the  Earth's  surface.  They  also  show  some 
of  the  most  complex  subsolidus  behavior 
of  all  minerals,  which  may  be  attributed  to 
the  interplay  between  the  effects  of  succes- 
sive displacive  and  cation-ordering  phase 
transitions.  Much  of  this  behavior  has  now 
been  quantified  and  rationalized  within  the 
scope  of  Landau  Theory  (Carpenter,  1988). 

The  ordering  of  Al  and  Si  at  low  tem- 
peratures is  common  in  many  minerals,  but 
it  has  always  been  assumed  that  this  pro- 
cess is  largely  configurational,  and  does 
not  change  the  heat  capacity.  However, 
many  studies  of  feldspars  have  shown  that 
Al/Si  ordering  induces  a  lattice  strain  in  the 
crystal,  thereby  coupling  to  any  prevalent 
displacive  distortion.  It  would  be  logical  to 
suppose  that  such  ordering  would 
renormalize  phonon  frequencies,  thereby 
altering  the  heat  capacity.  If  this  does  turn 
out  to  be  the  case,  then  current  attitudes  to 
mineral  energetics  will  have  to  be  reevalu- 
ated. 


One  is  not  predicting  substantial  ACP 
effects,  and  so  testing  this  hypothesis  re- 
quires sensitive  experimental  procedures. 
It  would  be  desirable  to  compare  the  ener- 
getics of  two  (or  more)  samples  from  the 
same  specimen,  which  differ  only  in  the 
extent  of  their  Al/Si  ordering  (measured  by 
the  order  parameter  Qod)-  Changes  in  the 
Al/Si  ordering  require  solid-state  diffusion, 
which  is  an  extremely  sluggish  process 
below  1500  K  or  so.  In  order  to  be  able  to 
prepare  samples  with  different  Qod  in  the 
laboratory,  one  needs  a  material  with  a  very 
high  equilibrium  order/disorder  phase  tran- 
sition temperature  (Tc).  Samples  can  then 
be  studied  at  lower  temperatures  without 
the  risk  of  continued  ordering. 

Anorthite,  CaAl2Si20s,  an  end-mem- 
ber plagioclase  feldspar,  is  a  good  candi- 
date for  this  study.  Not  only  has  the  mineral 
been  extensively  studied,  its  transition  be- 
havior is  extremely  well  characterized  (Car- 
penter 1988,  1991).  On  heating,  there  is  a 
displacive  phase  transition  from  P\  to  7T  at 
510  K.  Continued  heating  induces  pro- 
gressive disorder  of  Al  and  Si  over  the 
tetrahedral  sites,  and  the  symmetry  ap- 
proaches CT,  although  anorthite  melts  (at 
-1800  K)  before  reaching  the  hypothetical 
/T  -  C\  phase  transition  temperature.  The 
most  ordered  samples  available  have  Qod 
~  0.92,  and  this  can  be  reduced,  as  required, 
by  annealing  at  high  temperatures. 

For  this  study  on  the  effect  of  ordering 
on  the  molar  heat  capacity,  a  natural  anor- 
thite was  used  (Qod  =  0.92,  as  measured  by 
x-ray  diffraction).  A  portion  of  the  material 
was  annealed  at  1723  K  for  21  days  to 
induce  some  Al/Si  disorder,  thereby  reduc- 
ing Qod  to  0.82.    These  samples  were 


GEOPHYSICAL  LABORATORY 


123 


provided  by  Michael  Carpenter  (Univer- 
sity of  Cambridge).  Differential  scanning 
calorimetry  (DSC)  runs  performed  in  Cam- 
bridge revealed  no  differences  in  the  heat 
capacities  of  these  samples  at  low  tempera- 
tures (T<  900  K).  However,  this  technique 
cannot  be  used  at  higher  temperatures  and 
so  for  a  complete  investigation  it  was  de- 
cided to  use  transposed-temperature  drop 
calorimetry  at  Princeton  University,  in  col- 
laboration with  Alexandra  Navrotsky.  This 
technique  permits  determination  of  heat 
capacity,  albeit  indirectly,  by  measuring 
the  relative  enthalpies  at  different  tempera- 
tures. Precise  determination  of  Cp  necessi- 
tates many  measurements  at  closely  spaced 
temperatures,  but  as  a  preliminary  test  to 
see  whether  Al/Si  ordering  does  induce  a 
ACp  effect,  it  is  sufficient  to  measure  en- 
thalpies at  a  few  temperatures,  to  see  if  AH 
between  the  samples  varies  as  a  function  of 
temperature. 

Samples  of  anorthite  were  sealed  in  Pt 
foil  and  dropped  into  a  receptacle  within  a 
"Setaram"  calorimeter,  set  to  the  desired 
temperature.  The  enthalpy  change  of  the 
sample  from  room  temperature  to  the  calo- 
rimeter temperature  is  then  proportional  to 
the  energy  required  to  restore  the  tempera- 
ture of  the  assembly.  An  alumina-filled 
capsule  was  used  as  calibration  standard. 
Enthalpy  measurements  were  repeated  three 
to  seven  times  per  sample  at  each  tempera- 
ture, to  check  reproducibility.  The  results 
are  displayed  in  Table  20. 

It  must  be  stressed  that  these  are  rela- 
tive enthalpies,  that  is,  //r-//294K-  The  ac- 
tual enthalpy  of  ordering  between  the 
samples  is  ~  8  kJmoH  (Carpenter,  pers. 
comm.).  The  increase  in  AH  between  the 


samples  with  increasing  temperature  is  sig- 
nificant, and  may  well  indicate  a  ACp  ef- 
fect. Only  at  the  highest  temperature  is 
there  the  possibility  of  some  slight  disor- 
dering within  the  calorimeter,  though  this 
seems  unlikely  because  of  the  short  mea- 
surement time  (15  minutes,  typically). 
Because  the  samples  used  were  natural, 
albeit  very  pure,  one  cannot  entirely  dis- 
count other  effects  due  to  trace  impurities. 
A  second  set  of  experiments,  using  syn- 
thetic anorthites  is  now  underway  in  order 
to  clarify  this  point. 

These  measurements  have  shown  that 
it  is  possible  to  measure  enthalpy  differ- 
ences between  minerals  on  the  order  of  a 
few  joules  for  a  30-mg  sample  (3  kJmol-1 
for  anorthite)  at  temperatures  far  beyond 
the  reach  of  conventional  calorimetric  tech- 
niques. Although  this  method  is  not  as 
precise  as  low-temperature  DSC,  further 
enhancements,  such  as  more  sensitive  de- 
tector systems,  are  being  developed  to  al- 
low increased  resolution  at  higher  tem- 
peratures. 


Table  20.  Relative  enthalpy  of  anorthite  from 
973  K  to  1673  K.* 


T[K] 

HtH294K 

Qod=0M 

[kJmol1] 

Qoo=om 

973 

1273 
1473 
1673 

194(4) 
288(3) 
349(3) 
429(6) 

195(3) 
290(4) 
369(3) 
444(8) 

*  Errors  are  twice  the  standard  deviation  of 
the  mean. 


124 


CARNEGIE  INSTITUTION 


High-Pressure  Studies  of  Phase 
Transitions  in  Minerals 


A  New,  High  Pressure  Phase  Transition 
in  Leucite 


It  is  possible  to  extend  Landau  Theory 
to  the  description  of  high-pressure  phase 
transition  behavior,  using  suitably  chosen 
"model  systems".  The  aim  of  this  research 
is  to  concentrate  on  rock-forming  minerals 
that  exist  within  the  Earth's  crust,  where 
moderate  temperatures  and  pressures  pre- 
vail. The  study  of  the  pressure  dependence 
of  such  phases  provides  a  logical  step  from 
previous,  high- temperature  studies  of  phase 
transition  behavior. 

Feldspathoid  minerals  were  selected  for 
study.  They  are  relatively  abundant  within 
the  crust  and,  like  the  feldspars,  have  alu- 
minosilicate  frameworks  with  cations  in 
interstitial  sites.  Phase  transitions  involv- 
ing both  displacive  distortions  and  cation 
ordering  are  extremely  common. 
Feldspathoids  may  be  distinguished  from 
feldspars  by  the  presence  of  structural  chan- 
nels, which  provide  an  important  reposito- 
ries for  cations,  water,  organic  molecules 
etc.,  and  are  pathways  for  ionic  conduction 
(Palmer  and  Salje,  1990;  Alpena  ai,  1977). 
In  this  sense,  feldspathoids  may  be  com- 
pared to  zeolite  minerals.  The  presence  of 
structural  channels  makes  feldspathoids 
low-density  minerals;  thus  feldspathoids 
might  be  expected  to  be  very  unstable  at 
high  pressures.  However,  high-tempera- 
ture studies  of  feldspathoids  reveal  a  re- 
markable structural  adaptability,  and  this 
may  prolong  the  stability  (or  metastability) 
of  such  structures  at  much  higher  pressures 
than  previously  imagined. 


Leucite,  KAIS12O6,  is  a  feldspathoid 
mineral  associated  with  SiC>2-poor,  K-rich 
alkaline  volcanics,  which  has  been  well- 
studied  as  a  function  of  temperature.  On 
cooling,  leucite  undergoes  two  phase  tran- 
sitions, from  a  cubic  phase  Ia3d  to  an 
intermediate  tetragonal  phase  I4\lacd  at  Tc 
=  938  K,  described  by  order  parameter  Qi 
(Eg  symmetry);  then  to  a  low-7  tetragonal 
phase  14 1  la  at  7c  =  918  K,  described  by 
order  parameter  Qu  with  Tig  symmetry 
(Palmer  er  al,  1989).  There  is  a  continuous 
volume  decrease  in  the  low-rphase,  asso- 
ciated with  collapse  of  the  <1 1 1>  structural 
channels;  this  is  described  by  Qu .  The  first 
order  parameter,  Qjy  describes  a  ferroelastic 
distortion  of  the  unit  cell  (no  volume 
change).  Such  a  distortion  —  an  acoustic 
shear  mode  —  should  show  little  or  no 
pressure  dependence,  compared  to  the 
highly  pressure  dependent  volume  distor- 
tion. Increasing  pressure,  therefore,  is  ex- 
pected to  modify  the  transition  behavior  by 
enhancing  Qu  relative  to  Q/. 

The  pressure  dependence  of  leucite  has 
been  followed  up  to  60  kbar,  using  Raman 
spectroscopy.  A  single  crystal  was  placed 
within  a  Mao-Bell  diamond-anvil  cell,  us- 
ing an  organic  liquid,  "FC75"  (C8F16O)  as 
pressure  medium.  Pressure  determination 
was  by  the  ruby  fluorescence  method,  with 
ruby  spectra  measured  using  profile  refine- 
ment. 

Measurement  of  the  Raman  spectra  was 
possible  using  a  0.6W  He-Ne  laser,  with  a 


GEOPHYSICAL  LABORATORY 


125 


580 


570 


I  560 

|  550 

c 
® 

|  540 


530 


520 


■  increasing  pressure 
□  decreasing  pressure 


10         20         30         40 
Pressure,  kbar 


50 


60 


Fig.  67.  Pressure  dependence  of  the  Eg  Raman 
mode  forleucite.  The  large  increase  in  frequency 
at  Pc,  and  a  hysteresis  on  increasing  and  decreas- 
ing pressure,  indicate  the  existence  of  a  first-order 
phase  transition. 


"Triplemate"  detector  and  Princeton  In- 
struments control  system.  The  leucite 
Raman  spectrum  contains  two  intense 
peaks,  and  a  number  of  much  weaker  modes. 
High-temperature  Raman  spectroscopy 
(Palmer  et  aL,  1990)  showed  that  the  in- 
tense modes  relate  to  the  two  order  param- 
eters, and  may  be  assigned  to  T\g  and  Eg 
symmetries. 

For  this  study,  the  pressure  dependence 
of  the  Eg  mode  was  followed,  revealing  the 
presence  of  a  previously  unknown  phase 
transition  at  P  =  23  kb  on  increasing  pres- 
sure (Fig.  67).  The  similarity  of  the  Raman 
spectra  on  either  side  of  the  phase  transi- 
tion suggests  that  both  "high"  and  "low" 
phases  are  related.  The  fact  that  the  phase 
transition  is  totally  reversible,  together  with 
its  rapid  kinetics,  suggest  that  the  mecha- 
nism is  displacive.  The  existence  of  a 
hysteresis  implies  that  the  phase  transition 
is  first  order  in  character.  The  existence  of 


a  supergroup/subgroup  relation  to  this  phase 
transition  limits  the  choice  of  possible  high- 
pressure  phases.  We  hope  to  carry  out  x-ray 
work  to  refine  the  structure  of  the  high 
pressure  phase,  and  to  fully  characterize 
the  phase  transition  behavior. 


References 

Alpen,  U.U.,  H.  Schulz,  G.  H.  Tatat,  and  H. 
Boehm,  One-dimensional  cooperarive  Li-dif- 
fusion in  p-eucryptite.  Solid  State  Commun., 
25,911-914,  1977. 

Carpenter,  M.  A.,  Thermochemistry  of  alumi- 
nium/silicon ordering  in  feldspar  minerals,  in 
Physical  Properties  and  Thermodynamic 
Behaviour  of  Minerals,  E.K.H.  Salje,  ed., 
D.Reidel,  Dordrecht,  pp.  265-323,  1988. 

Carpenter,  M.  A.,  Thermodynamics  of  phase  tran- 
sitions in  minerals:  A  macroscopic  approach., 
in  Stability  of  Minerals,  G.D.  Price,  ed.,  Allen 
and  Unwin,  Boston  (in  press),  1991. 

Landau,  L.D.,  and  E.  M.  Lifshitz,  Statistical 
Physics,  Edition  3,  part  1,  Pergamon  Press, 
Oxford,  1980. 

Palmer,  D.C.,  and  E.  K.  H.  Salje,  Phase  transitions 
in  leucite:  dielectric  properties  and  transition 
mechanism.  Phys.  Chem.  Minerals,  17,  444- 
452,  1990. 

Palmer, D.C., E.  K.  H.  Salje,  and  W.  W.  Schmahl, 
Phase  Transitions  in  leucite:  X-ray  diffraction 
studies.  Phys.  Chem.  Minerals,  16,  714-719, 
1989. 

Palmer,  D.C.,  U.  Bismayer,  and  E.  K.  H.  Salje, 
Phase  transitions  in  leucite:  Order  parameter 
behaviour  and  the  Landau  Potential  deduced 
from  Raman  spectroscopy  and  birefringence 
studies.  Phys.  Chem.  Minerals,  17,  259-265, 
1990. 

Salje,  E.,  B.  Kuscholke,  B.  Wruck,  and  H.  Kroll, 
Thermodynamics  of  sodium  feldspar  II:  ex- 
perimental results  and  numerical  calculations. 
Phys.  Chem.  Minerals,  12,  99-1071,  1985. 


126 


CARNEGIE  INSTITUTION 


First-principles  Studies  of  Elasticity 

and  Post-Stishovite  Phase  Transitions 

in  S1O2* 

Ronald  E.  Cohen 

Stishovite  is  a  candidate  mineral  for  the 
Earth's  transition  zone  and  lower  mantle, 
and  is  also  the  prototypical  octahedrally 
coordinated  silicate.  It  is  an  open  ques- 
tion, however,  whether  stishovite  remains 
the  stable  form  of  Si02  throughout  the 
lowermantle,  or  if  a  new  structure  for  Si02 
forms  at  the  high  pressure  conditions  of  the 
lower  mantle.  This  is  an  important  ques- 
tion, because  the  presence  or  absence  of 
stishovite  depends  on  the  chemical  compo- 
sition of  the  lower  mantle,  particularly  the 
Fe-Mg  ratio  (e.g.  Fei  and  Hemley,  1991). 

Two  recent  studies  suggestpossible  phase 
transitions  under  mantle  conditions.  Park 
et  al.  (1988)  predicted  a  phase  transition 
from  stishovite  (rutile  structure)  to  the 
pyrite  structure  (Pa3)  at  60  GPa  using  self- 
consistent  Linearized  Augmented  Plane 
Wave  (LAPW)  calculations.  Tsuchidaand 
Yagi(1989)  looked  for  this  transition  using 
in  situ  x-ray  diffraction  in  the  diamond 
anvil  cell,  and  instead  reported  a  phase 
transition  in  Si02  from  stishovite  to  the 
CaCl2  structure  between  80-100  GPa.  A 
transition  from  stishovite  to  the  CaCb  struc- 
ture in  the  lower  mantle  would  be  very 
important  in  geophysical  modeling  of  the 
Earth.  This  phase  transition  involves  an 
elastic  instability  where  c\\-c\2  becomes 


*The  computations  were  performed  on  the  Cray  2 
at  the  National  Center  for  Supercomputing  Appli- 
cations under  the  auspices  of  the  National  Science 
Foundation. 


unstable  (Cohen,  1987;  Hemley,  1987). 
Such  a  transition  should  be  evident  in  seis- 
mological  data  (derived  acoustic  veloci- 
ties) if  stishovite  is  present  in  any  quantity 
in  the  deep  Earth.  No  such  features  are 
observed  in  the  lowermantle  except  for  the 
anomalous  D"  zone  at  the  base  of  the 
mantle.  One  must  conclude  either  that 
little  stishovite  is  present  in  the  deep  man- 
tle, or  that  the  phase  transition  in  Si02 
occurs  in  D"  and  seismic  anomalies  in  D" 
reflect  at  least  partially  that  transition. 

These  questions  are  addressed  here  us- 
ing the  Linearized  Augmented  Plane  Wave 
method  (Wei  and  Krakauer,  1985).  The 
present  calculations  represent  one  of  the 
most  extensive  studies  of  a  single  material 
using  this  technique — over  200  self-con- 
sistent calculations  were  performed  to  de- 
termine the  phase  relations,  elasticity,  and 
vibrational  properties  of  Si02  in  the 
stishovite,  CaCl2,  and  Pa3  structures 
(Cohen,  1 99 1  a,b).  These  calculations  make 
no  assumptions  about  ionicity,  bonding,  or 
form  of  the  electron  distribution.  The  only 
inputs  are  the  nuclear  charges  and  posi- 
tions, and  the  output  is  a  total  energy  for 
that  nuclear  configuration.  The  quantum 
mechanical  problem  is  solved  within  the 
local  density  approximation  (LDA)  (Hedin 
and  Lundqvist,  1 97 1 )  of  density  functional 
theory  (Hohenberg  and  Kohn,  1964).  The 
electrostatic  and  kinetic  energy  contribu- 
tions to  the  energy  and  potential  are  evalu- 
ated numerically,  and  can  be  converged  to 
any  necessary  accuracy.  In  the  LDA,  the 
quantum-chemical  contributions  are  mod- 
eled by  assuming  that  the  exchange  and 
correlation  contributions  to  the  potential 
and  energy  can  be  obtained  from  the  ex- 


GEOPHYSICAL  LABORATORY 


127 


> 

ID 


O   O* 


stishovite 


I       i    ■    i      'i 


30 


40 


50 


v(AJ) 


Fig.  68.  Equation-of- state  (energy  versus  volume 
at  0  K)  of  stishovite,  CaCl2,  and  Pa3.  The  transi- 
tion from  stishovite  to  Pa3  is  at  about  160  GPa. 
CaCh  becomes  stable  relative  to  stishovite  at  45 
GPa.  The  initial  energy  difference  between  CaCl2 
and  stishovite  is  very  small  relative  to  the  energy 
changes  with  volume. 


50  75 

P  (GPa) 


Fig.  69.  The  elastic  constant  c\  \-c\2  as  a  function 
of  pressure.  The  solid  line  is  a  spline  fit  to  the 
calculated  points  (circles).  The  plus  is  the  Brillouin 
scattering  data  of  Weidner  et  al.  (1982).  The 
dashed  curve  is  for  the  high  pressure  CaCl2  phase. 
The  phase  transition  is  predicted  to  occur  at  45 
GPa  at  0  K. 


change  and  correlation  functionals  for  the 
uniform  electron  gas.  During  the  last  ten 
years,  this  approximation  has  been  demon- 
strated to  be  quite  accurate  for  metals, 
semiconductors,  and  insulators,  with  some 
exceptions  primarily  in  magnetic  crystals. 

Calculations  were  performed  for  Si02  in 
the  stishovite  (rutile,  space  group  P4jJ 
mnm),  CaCl2  {Pnnm),  and  pyrite  (Pa3) 
structures  as  functions  of  volume.  The 
internal  structural  parameters  and  lattice 
parameters  were  optimized  at  each  vol- 
ume. Fitting  the  total  energies  to  an  equa- 
tion-of-state  and  to  polynomials  as  func- 
tions of  strain  and  phonon  amplitudes  gives 
the  Alg  and  Big  Raman  modes  as  well  as 
four  independent  elastic  constants  for 
stishovite  as  functions  of  pressure.  The  set 
of  calculations  also  gives  the  phase  transi- 
tion pressures  from  stishovite  to  the  Pa3 
and  CaCl2  structures.  All  of  the  present 
calculations  are  for  static  lattice  energies, 
classically  equivalent  to  a  temperature  of  0 
K. 

Figure  68  shows  the  energy  versus  vol- 
ume for  the  three  phases.  A  phase  transi- 
tion from  stishovite  to  Pa3  is  predicted  at 
156  GPa.  This  is  significantly  higher  in 
pressure  than  the  value  of  60  GPa  obtained 
by  Park  et  al.  (1988)  using  LAPW.  The 
reason  for  the  difference  is  not  clear,  but 
extensive  convergence  tests  of  the  present 
results  indicate  stability  of  about  1  GPa  in 
the  transition  pressure  with  respect  to  the 
convergence  parameters. 

Figure  69  shows  the  calculated  elastic 
constant  c\\-c\2  as  a  function  of  pressure. 
An  elastic  instability  (c\\-c\2  vanishes)  in 
stishovite  is  predicted  at  45  GPa,  at  which 


128 


CARNEGIE  INSTITUTION 


pressure  stishovite  would  transform  into 
the  CaCl2  structure  at  zero  temperature, 
Figure  68  shows  that  this  phase  transition 
has  only  a  small  effect  on  the  energetics  of 
Si02.  It  also  has  only  a  small  effect,  ini- 
tially, on  the  crystal  structure.  The  phase 
transition  is  continuous,  and  the  initial  dis- 
tortions from  the  rutile  structure  are  very 
small.  However,  the  phase  transition  has 
enormous  effects  on  the  elastic  properties 
of  high  pressure  Si02,  so  it  is  crucial  to 
consider  whether  this  transition  occurs  un- 
der mantle  conditions. 

The  phase  transition  is  likely  to  be  very 
sensitive  to  temperature  since  the  instability 
is  driven  by  a  Raman  mode.  Figure  70 
shows  the  Aig  and  Big  Raman  modes  as 
functions  of  pressure.  Agreement  with  the 
Raman  data  obtained  by  Hemley  (1987)  is 
excellent.  At  the  CaCl2  phase  transition 
the  Big  mode  becomes  an  Ag  mode  and 
begins  to  increase  in  frequency  with  pres- 
sure, whereas  as  lower  pressures  the  fre- 
quency decreases  with  increasing  pressure. 
This  is  probably  the  most  direct  way  of 
detecting  the  phase  transition,  since  the 
distortions  are  very  small  at  the  transition 
and  would  be  difficult  to  detect  with  x-rays. 
The  phase  transition  is  not  a  soft-mode 
transition  in  the  traditional  sense,  since  the 
Big  frequency  does  not  reach  zero  at  the 
transition.  The  Big  mode  does  drive  the 
transition,  since  if  the  atoms  were  not  al- 
lowed to  distort  along  the  Big  eigenvector 
during  a  c\i-c\2  strain,  c\\-c 12  would  not 
become  unstable.  It  is  the  coupling  be- 
tween the  strain  and  the  phonon  displace- 
ment that  makes  the  energy  decrease  as  a 
function  of  strain,  and  thus  causes  the  phase 
transition.  The  Big  Raman  mode  contrib- 


150 


Fig.  70.  Pressure  dependence  of  the  Aig  and  Big 
Raman  frequencies  for  stishovite,  and  an  Ag 
mode  in  CaCl2-  The  triangles  and  diamonds  are 
experimental  data  from  Hemley  (1987).  The 
dashed  line  is  a  prediction  for  the  CaCl2  phase. 
Note  that  the  "soft  mode"  does  not  go  to  zero,  or 
even  get  very  small  at  the  phase  transition. 


utes  a  factor  to  c\\-c\2  that  is  proportional 
to  -1/co2  (Miller  and  Axe,  1967),  so  that  as 
the  Big  frequency  decreases  the  elastic 
constant  c\\-c yi  is  destabilized.  Tempera- 
ture, however,  would  stabilize  the  higher 
symmetry  rutile  structure.  At  150  GPa,  the 
well  depth  (energy  gain  on  the  phase  tran- 
sition) expressed  as  temperature  is  only 
1 275  K,  so  at  mantle  temperatures  of  2000- 
3000  K  the  transition  may  not  occur  until 
much  higher  pressures  than  the  45  GPa 
calculated  for  zero  temperature.  The  exact 
transition  temperature  can  be  calculated 
using  molecular  dynamics  or  Monte  Carlo 
simulations  using  a  potential  model,  or  by 
fitting  a  simplified  Hamiltonian  to  the  total 
energy  results  and  using  statistical  thermo- 
dynamics to  evaluate  the  phase  diagram 
(Rabe  and  Joannopoulos,  1987).  Both  ap- 
proaches require  information  about  the 


GEOPHYSICAL  LABORATORY 


129 


coupling  between  cells,  as  well  as  the  total 
energies  for  zone  center  distortions  that  do 
not  increase  the  unit  cell  size.  A  potential 
model  is  now  being  developed  for  high- 
pressure  Si02  to  investigate  the  thermal 
properties  of  stishovite  and  the  tempera- 
ture dependence  of  the  phase  transition.  It 
is  quite  possible  that  temperature  will  sig- 
nificantly increase  the  depth  at  which  the 
transition  occurs,  and  thus  the  transition 
may  be  partly  responsible  for  the  anoma- 
lous seismic  properties  of  the  so-called  D" 
region. 


References 

Cohen,  R.  E.,  Calculation  of  elasticity  and  high 
pressure  instabilities  in  corundum  and 
stishovite  with  the  potential  induced  breathing 
model,  Geophys.  Res.  Lett.,  14,  37-40,  1987. 

Cohen,  R.  E.,  Bonding  and  elasticity  of  stishovite 
Si02  at  high  pressure:  Linearized  augmented 
plane  wave  calculations,  Amer.  Mineral.,  76, 
733-742,  1991a. 

Cohen,  R.  E.,  First-principles  predictions  of  elas- 
ticity and  phase  transitions  in  high  pressure 
Si02  and  geophysical  implications,  in  High 
Pressure  Research  in  Mineral  Physics:  Appli- 
cation to  Earth  and  Planetary  Science  (Pro- 
ceedings of  U.S. -Japan  Conference  on  High 
Pressure  Geophysics,  Ise,  Japan,  January, 
1991),  M.H.  Manghnani  and  Y.  Syono,  eds., 
in  press,  1991b. 

Fei,  Y.  and  R.  J.  Hemley,  Stability  of  (Fe,Mg)Si03- 
perovskite  in  the  lower  mantle,  Geophys.  Res. 
Lett.,  in  press,  1991. 

Hedin,  L.  and  B.  I.  Lundqvist,  Explicit  local 
exchange-correlation  potentials,  /.  Phys.,  C4, 
2064-2083,  1971. 

Hemley,  R.  J.,  Pressure  dependence  of  Raman 
spectra  of  Si02  polymorphs:  a-quartz,  coesite, 
and  stishovite.  In  High  -Pressure  Research  in 
Mineral  Physics,  M.H.  Manghnani  and  Y. 
Syono,  eds.,  pp.  347-359.  American  Geo- 
physical Union,  Washington,  D.C.,  1987 

Hohenberg,  P.,  and  W.  Kohn,  Inhomogeneous 
electron  gas,  P/ry.s./?ev.,7.?<5£,  864-871, 1964. 

Kohn,  W.  and  L.  J.  Sham,  Self-consistent  equa- 
tions including  exchange  and  correlation  ef- 


fects, Phys.  Rev.,  140  A,  1133-1140,  1965. 

Miller,  P.  B.  and  F.  D.  Axe,  Internal  strain  and 
Raman  active  vibrations  in  solids,  Phys.  Rev., 
163,  924-926,  1967. 

Park,  K.  T.,  K.  Terakura,  and  Y.  Matsui,  Theoreti- 
cal evidence  for  a  new  ultra-high-pressure 
phase  of  Si02,  Nature,  336,  670-672,  1988. 

Rabe,  K.  M.,  and  J.  D.  Joannopoulos,  Theory  of 
the  structural  phase  transition  of  GeTe,  Phys. 
Rev.  B,  36,  6631-6639,  1987. 

Tsuchida,  Y.  and  T.  Yagi,  A  new,  post-stishovite 
high-pressure  polymorph  of  silica,  Nature, 
340,  217-220,  1989. 

Tsuneyuki,  S . ,  M.  Tsukada,  H.  Aoki,  and  Y.  Matsui, 
First-principles  interatomic  potential  of  silica 
applied  to  molecular  dynamics,  Phys.  Rev. 
Lett.,  61,  869-872,  1988. 

Tsuneyuki,  S.,  Y.  Matsui,  H.  Aoki,  andM.  Tsukada, 
New  pressure-induced  structural  transforma- 
tions in  silica  obtained  by  computer  simula- 
tion, Nature,  339,  209-21 1,  1989. 

Wei,  S.  H.,  and  H.  Krakauer,  Local  density  func- 
tional calculation  of  the  pressure  induced  phase 
transition  and  metallization  of  BaSe  and  BaTe, 
Phys.  Rev.  Lett.,  55,  1200-1203,  1985. 

Weidner,  D.  J.,  J.  D.  Bass,  A.  E.  Ringwood,  and  W. 
Sinclair,  The  single-crystal  elastic  moduli  of 
stishovite,  /.  Geophys.  Res.,  87,  B4740-4746, 
1982. 


Molecular  Dynamics  Simulations  of 
Melting  of  MgO  at  High  Pressures* 

Zhaoxin  Gong,  Ronald  E.  Cohen,  and 
Larry  L.  Boyer** 

We  are  developing  a  general-purpose 
molecular  dynamics  (MD)  program  for 
studying  finite  clusters  of  atoms  using 
many-body  potentials  and  long-range 
forces.  This  program  will  be  used  to  study 
high-pressure  and  high-temperature  phase 


*  This  work  is  supported  by  the  Office  of  Naval 
Researchgrant#N00014-91-J-1227toREC.  Com- 
putations were  performed  with  the  support  of 
ONR  at  the  NRL  Connection  Machine  Facility 
and  at  the  Pittsburgh  Supercomputer  Center  under 
the  auspices  of  the  National  Science  Foundation. 
*Complex  Systems  Theory  Branch,  Naval  Re- 
search Laboratory,  Washington,  D.C.  20375 


130 


CARNEGIE  INSTITUTION 


transitions  as  well  as  ferroelectricity. 
Though  many  simulations  have  been  per- 
formed on  various  systems  of  interest,  rela- 
tively few  have  been  performed  on  ionic 
systems  using  first-principles,  non-empiri- 
cal potentials.  Such  potentials  are  expected 
to  be  more  reliable  outside  the  range  of 
experimental  data  than  empirical  poten- 
tials, and  furthermore  can  be  constrained 
for  arbitrary  ionic  configurations,  not 
merely  determined  from  bulk  properties 
with  atomic  positions  close  to  their  equilib- 
rium sites.  A  cluster  approach  is  important 
because  periodic  boundary  conditions  have 
been  found  to  greatly  inhibit  phase  transi- 
tions such  as  melting,  especially  in  systems 
with  long-range  forces,  because  the  peri- 
odic boundary  conditions  force  the  "liq- 
uid" state  onto  a  periodic  lattice.  Also,  in 
the  case  of  ferroelectricity,  electric  fields 
and  macroscopic  polarization  are  by  defi- 
nition "non-periodic,"  so  that  these  phe- 
nomena can  only  be  studied  approximately 
using  periodic  boundary  conditions.  Finite 
clusters  in  free  space  were  used  in  earlier 
MD  studies  of  melting  of  NaF  using  a 
Gordon-Kim  rigid  ion  potential  (Boyer  and 
Pawley,  1988;  Boyer  and  Edwardson, 
1988). 

In  order  to  use  a  cluster  method,  many 
atoms  must  be  be  included  in  order  to 
minimize  the  effects  of  surfaces;  however, 
with  long-range  forces  the  MD  problem  is 
an  N2  problem,  where  N  is  the  number  of 
atoms  in  the  cluster,  so  computational  effi- 
ciency is  very  important.  We  present  re- 
sults here  using  a  massively  parallel  com- 
puter, the  Connection  Machine  CM-2,  as 
well  as  results  calculated  on  an  IBM  RS/ 
6000  superworkstation  and  a  Cray  YMR 


Here  we  report  for  the  first  time  MD  results 
of  melting  of  MgO  at  high  pressures  using 
a  many-body  potential,  the  Potential  In- 
duced Breathing  (PIB)  model. 

The  PIB  model  is  an  ab  initio  model 
where  no  parameters  are  fitted  to  the  ex- 
periment. This  model  has  been  very  suc- 
cessful in  predicting  the  thermodynamic 
and  elastic  properties  of  alkaline  earth  ox- 
ides, including  the  Cauchy  violations  (c\\ 
is  not  equal  to  cyi  at  zero  pressure)  (Boyer 
era/.,  1985;  Mehle/tf/.,  1986;  Cohen  etaL, 
1987;  Isaak  et  al.,  1990).  Monte  Carlo 
(MC)  and  MD  simulations  for  the  PIB 
model  are  computationally  demanding  and 
complicated.  So  far  only  primitive  MC 
results  of  equation  of  state  at  zero  pressure 
using  the  PIB  model  have  been  reported  for 
a  sample  of  MgO  with  64  atoms  (Cowley  et 
al„  1990). 

Isaak  et  al.  (1990)  found  that  the  PIB 
model  predicts  that  the  zero  pressure  iso- 
thermal elastic  modulus  Cs(=c\i-C22)  of 
MgO  becomes  negative  at  a  temperature 
very  close  to  the  melting  temperature  of 
MgO.  The  Cy  instability  occurs  before  the 
bulk  modulus  instability  discussed  by  Boyer 
(1985)  with  increasing  temperature  if  the 
free  energy,  including  the  vibrational  con- 
tribution, is  calculated  as  a  function  of 
strain.  To  find  out  whether  this  instability 
is  related  to  the  melting,  we  investigate 
here  whether  the  same  PIB  potential  indi- 
cates a  melting  point  close  to  the 
quasiharmonic  elastic  instability.  Further- 
more, are  there  elastic  instabilities  that  cor- 
respond to  melting  at  high  pressures? 

An  important  question  of  geophysical 
importance  is  the  curvature  of  the  melting 
curve  at  high  pressures.  At  low  pressures 


GEOPHYSICAL  LABORATORY 


131 


the  melt  is  generally  (but  not  always)  less 
dense  than  the  crystalline  form  of  a  mate- 
rial, but  the  liquid  is  more  compressible. 
As  pressure  is  increased,  the  difference  in 
densities  of  liquid  and  solid  should  de- 
crease, and  the  slope  of  the  melting  curve, 
dT/dP,  should  decrease  with  pressure.  The 
melting  curve  may  become  horizontal  or 
may  reach  a  maximum  and  bend  over. 
MgO,  being  a  close-packed  solid,  is  an 
interesting  case  because  of  the  simple  crys- 
tal structure  and  the  probably  simple  struc- 
ture of  the  liquid.  MgO  is  also  an 
endmember  of  magnesiowustite  (Mg,Fe)0, 
which  is  considered  a  likely  mineral  in  the 
Earth's  lower  mantle. 

We  used  the  same  numerical  PIB  poten- 
tials as  in  Isaak  et  al.  (1990),  but  we  have 
fit  them  to  a  different  form  in  order  to 
increase  accuracy  and  computational  effi- 
ciency. One  problem  is  that  the  Madelung 
(electrostatic)  potentials  on  the  oxygen  ions 
on  the  surface  of  the  cluster  are  lower  in 
magnitude  than  in  the  bulk.  The  quantita- 
tive results  presented  here  must  be  consid- 
ered preliminary,  because  the  surface  atom 
potentials  often  fell  outside  the  range  of  the 
points  at  which  the  PEB  potential  was  fit. 
We  expect  the  changes  to  be  small  when  a 
better  potential  is  used,  because  the  form 
we  chose  appears  to  extrapolate  smoothly, 
and  the  great  majority  of  atoms  had  poten- 
tials that  fell  within  range  of  the  fit. 

We  tested  the  MD  code  by  comparing 
the  zero-pressure  density  of  a  free  cluster 
with  that  from  quasiharmonic  lattice  dy- 
namics (LD)  calculations  using  the  same 
model  (Isaak  etai,  1990).  The  density  was 
calculated  using  the  method  of  Boyer  and 
Pawley  (1988).  At  T=  1300  K,  the  average 


density  for  the  innermost  region  of  a  cluster 
with  216  atoms  differs  from  the  LD  result 
by  about  7%.  The  difference  decreases  to 
about  6  %  when  the  cluster  size  is  increased 
to  512  atoms.  We  also  found  that  the  total 
energy  is  conserved  to  better  than  six  sig- 
nificant figures  over  many  thousands  of 
time  steps,  giving  confidence  in  the  MD 
code. 

We  performed  a  series  of  runs  to  find 
out  the  melting  temperature  at  zero  pres- 
sure. The  starting  configuration  of  216  (= 
6x6x6)  atoms  in  their  perfect  lattice 
positions  was  initially  given  a  small  amount 
of  kinetic  energy,  equivalent  to  a  tem- 
perature T  =  300  K.  The  simulation  was 
then  allowed  to  proceed  at  constant  energy 
for  30  ps  (10,000  time  steps);  then  the  total 
energy  of  the  cluster  was  increased  by 
scaling  up  the  velocities  to  give  an  increase 
in  temperature  of  400  K.  This  procedure 
was  repeated  several  times.  The  results  are 
drawn  in  Figures  71  and  72  as  solid  lines. 


0.55 


060  0.65 

E  +  275  (Hartree) 


0.70 


Fig.  71.  Plot  of  equilibrium  temperature  T  versus 
the  total  energy  E  at  P  =  0  GPa.  The  solid  line 
corresponds  to  increasing  energy  while  the  dotted 
line  corresponds  to  decreasing  energy. 


132 


CARNEGIE  INSTITUTION 


055 


060 


065 


0.70 


E  +  275  (Hartree) 


Fig.  72.  Zero  pressure  plot  of  equilibrium  density 
versus  the  total  energy,  E,  in  a  cubic  region  cen- 
tered at  the  center  of  mass  with  the  cube  side  equal 
to  the  nearest  neighbor  distance  of  the  perfect 
lattice.  The  solid  line  corresponds  to  increasing 
energy  while  the  dotted  line  corresponds  to  de- 
creasing energy. 


At  low  temperatures,  the  temperature  is 
almost  a  linear  function  of  the  energy  E. 
Thus  MgO  is  quite  harmonic  up  to  about 
half  the  melting  temperature.  The  tempera- 
ture increases  as  the  energy  increases  until 
T=  3815  K,  where  increasing  energy  leads 
to  a  temperature  decrease  to  about  3045  K. 
(This  run  is  three  times  longer  than  the 
other  runs  in  this  figure.)  The  temperature 
of  the  cluster  then  increases  again  as  the 
energy  is  increased.  The  density  of  cluster, 
on  the  other  hand,  decreases  slowly  as  the 
energy  is  increased  until  the  temperature 
reaches  about  T=  38 1 5  K.  The  density  then 
decreases  by  a  much  larger  amount  as  the 
energy  is  increased  by  another  increment 
and  the  temperature  drops  to  3045  K.  This 
simultaneous  larger  decrease  in  tempera- 
ture and  density  indicates  that  the  cluster 
melted  and  kinetic  energy  is  transferred 
into  potential  energy  (Boyer  and  Pawley, 


1988;  Boyer  and  Edwardson,  1988).  The 
dotted  lines  in  these  two  figures  correspond 
to  the  process  of  cooling.  With  decreasing 
energy,  the  cluster  freezes  to  a  crystal  with 
defects  due  to  the  high  cooling  rate,  which 
leads  to  hysteresis.  The  middle  of  the 
hysteresis  loop  is  about  3100  K,  which 
compares  remarkably  well  with  the  experi- 
mental melting  temperature  of  3098  ±  20  K 
(Stull  and  Prophet,  1971)  and  the  c\\ -c\ 2 
elastic  instability  (Isaak  et  al.,  1990)  using 
the  same  potential. 

In  a  second  series  of  calculations,  the 
cluster  is  confined  in  a  box  so  that  its 
pressure  can  be  varied.    It  is  natural  to 
repeat  procedures  for  a  free  cluster  to  find 
kinks  in  the  T  versus  E  curve,  while  the 
volume  is  kept  fixed.    At  large  volumes 
(low  pressures)  this  procedure  is  satisfac- 
tory, but  it  becomes  increasingly  difficult 
to  locate  the  kinks  as  they  becomes  less  and 
less  prominent  as  volume  becomes  smaller 
and  smaller.  At  a  kink,  either  a  large  differ- 
ence in  temperature  Tor  a  large  difference 
in  pressure  P  is  observed.  When  the  cluster 
melts,  its  temperature  decreases  while  its 
pressure  increases.  For  example,  when  the 
box  size  is  set  equal  to  23.735  bohr,  the 
drop  in  temperature  is  about  260  K,  from 
about  5880  K  to  about  5620  K,  and  the 
increment  in  pressure  is  about  4.5  GPa, 
from  about  7.7  GPa  to  about  12.2  GPa. 
However,  for  a  box  of  cube  side  of  23.23 
bohr,  we  were  not  able  to  identify  any  kink 
with  the  drop  in  temperature  larger  than  70 
K  and  the  increment  in  pressure  larger  than 
1.7  GPa. 

To  determine  the  melting  temperatures 
at  high  pressures,  we  first  find  the  tempera- 
ture and  energy  at  given  pressure  and  given 


GEOPHYSICAL  LABORATORY 


133 


*  8- 


0.55  0.60  0.65  0.70 

E  +  275  (Hartree) 


0.75 


Fig.  73.  Plot  of  equilibrium  temperature  T  versus 
the  total  energy  E  at  P  =  10  GPa.  The  solid  line 
corresponds  to  increasing  volume  while  the  dotted 
line  corresponds  to  decreasing  volume. 


P  (GPa) 

Fig.  74.  Plot  of  the  melting  temperature  Tm  versus 
the  pressure  P.  The  bars  indicate  the  maxima  and 
minima  of  the  heating  and  cooling  curves,  respec- 
tively, around  the  melting  points. 


volume,  then  we  plot  the  temperature  ver- 
sus energy  at  a  given  pressure  (Fig.  73). 
The  solid  line  represents  the  process  of 
increasing  volume,  while  the  dotted  line 
represents  decreasing  volume.  The  cube 
side  increment  in  this  figure  is  0.505  bohr, 
and  the  starting  box  side  is  20.2  bohr.  We 
identify  the  kinks  in  the  plot  as  associated 
with  the  melting. 

In  Figure  74  we  plot  the  melting  tem- 
perature versus  pressure.  The  melting  tem- 
perature, because  of  the  van  der  Waals  loop 
typically  present  in  our  results,  is  taken  to 
be  the  average  temperature  of  two  middle 
points  in  the  heating  and  cooling  curves. 
Although  only  three  points  have  been  cal- 
culated, it  is  obvious  that  there  is  signifi- 
cant curvature  in  the  melting  curve.  Melt- 
ing temperatures  are  presented  only  to  20 
GPa  because  of  an  instability  that  occurred 
at  high  temperatures,  due  to  the  fitting 
range  of  the  Watson  sphere  potential.  At 
high  temperatures  the  Watson  potentials  of 


many  atoms  lie  outside  the  fitting  range, 
which  can  be  corrected  by  the  development 
of  better  inter-atomic  potentials.  Further 
results  will  be  on  very  high  pressures  and 
larger  samples. 

Although  the  PIB  model  is  more  com- 
plicated than  the  rigid  ion  model,  many 
calculations  can  utilize  parallel  processing. 
For  example,  checking  for  atoms  that  cross 
the  box  boundary  can  be  done  in  parallel  in 
all  three  directions  for  all  the  atoms  in  the 
cluster.  The  most  time  consuming  part,  the 
force  calculation,  can  also  be  done  in  paral- 
lel. Many  simulations  were  done  on  a  CM- 
2  from  Thinking  Machine  Inc.  at  the  Naval 
Research  Laboratory  (NRL).  On  the  CM- 
2,  the  row  column  difference  (RCD)  method 
by  Boyer  and  Pawley  (1988)  is  used  when 
calculating  pair  interactions.  The  block 
sizes  used  are  the  largest  possible,  i.e,  block 
sizes  of  N  x  N,  which  has  been  shown  to  be 
the  most  efficient  way  of  calculating  pair 
interactions  for  a  cluster  of  512  atoms 


134 


CARNEGIE  INSTITUTION 


Table  21.  Timing  results  (sec  per  time  step)  for  MD  simulations  on  the  CM  (one  8k 
sequencer),  the  CRAY-YMP  (one  processor),  and  the  IBM  RS/6000-320  power  worksta- 
tion. 


Rigid  Ion  Model* 


PIB  Model 


CM 

CM 

CRAY 

IBM 

CPU  time 
Elapsed  Time 

216  atoms 
512  atoms 
1000  atoms 

216  atoms 
512  atoms 
1000  atoms 

0.23 
0.76 

0.32 
1.44 

0.13 
0.31 
1.02 

0.23 
0.43 
1.22 

0.055 
0.29 
1.01 

0.082 
0.46 
2.16 

0.31 

2.45 
12.3 

0.32 
2.50 
12.5 

The  timing  results  for  rigid  ion  model  are  from  Boyer  and  Edwardson  (1988). 


(Boyer  and  Edwardson,  1988).  For  a  clus- 
ter of  1000  atoms,  Boyer  and  Edwardson 
(1988)  were  limited  to  a  block  size  of  256 
x  256  by  memory  constraints.  In  the 
present  hardware  at  NRL,  we  encountered 
no  memory  constraints  for  clusters  of  up  to 
1000  atoms.  Timings  are  given  in  Table  21 . 
The  elapsed  times  are  for  an  otherwise 
empty  machine  except  for  the  Cray  tim- 
ings, which  were  taken  under  a  typical 
load. 

In  summary,  we  have  presented  first 
results  on  melting  of  a  oxide  using  a  non- 
empirical  many-body  model.  MD  simula- 
tions were  performed  on  finite  clusters 
containing  up  to  1000  atoms  on  the  mas- 
sively parallel  CM-2.  Significant  curva- 
ture in  the  melting  curve  is  predicted.  This 
method  will  be  applicable  to  study  a  large 
variety  of  phase  transitions  as  functions  of 
temperature  and  pressure. 


References 

Boyer,  L.  L.,  Theory  of  melting  based  on  lattice 
instability,  Phase  Transitions,  5,  1-48,  1985. 

Boyer,  L.  L.,  M.  J.  Mehl,  J.  L.  Feldman,  J.  R. 
Hardy,  J.  W.  Flocken,,  C.  Y.andFong,  Beyond 
the  rigid  ion  approximation  with  spherically 


symmetric  ions,  Phys.  Rev.  Lett.,  54,  1940- 
1943,  1985. 

Boyer,  L.  L.,  and  P.  J.  Edwardson,  Application  of 
massively  parallel  machines  to  molecular  dy- 
namics simulation  of  free  clusters,  Proceed- 
ings of  the  2nd  symposium  on  the  frontiers  of 
.  massively  parallel  computation,  Fairfax,  Vir- 
ginia, USA,  October  10-12,  1988. 

Boyer,  L.  L.,  andG.  S.  Pawley,  Molecular  dynam- 
ics of  clusters  of  particles  interacting  with 
pairwise  forces  using  a  massively  parallel  com- 
puter, /.  Comput.  Phys.,  78,  405-423,  1988. 

Cohen,  R.  E.,  L.  L.  Boyer,  and  M.  J.  Mehl,  Lattice 
dynamics  of  the  potential-induced  breathing 
model:  Phonon  dispersion  in  the  alkaline -earth 
oxides,  Phys.  Rev.  B,  35,  5749-5760,  1987. 

Cowley,  E.  R.,  S.  H.  Liu,  and  G.  K.  Horton,  Monte 
Carlo  calculations  of  the  equations  of  state  of 
alkaline  earth  oxides,  Ferroelectrics,  111,  33- 
42,  1990. 

Isaak,  D.  G.,  R.  E.  Cohen,  and  M.  J.  Mehl,  Calcu- 
lated elastic  and  thermal  properties  of  MgO  at 
high  pressures  and  temperatures,  J.  Geophys. 
Res.,  95,  7055-7067,  1990. 

Mehl,  M.  J.,  R.  J.  Hemley,  and  L.  L.,  Boyer, 
Potential-induced  breathing  model  for  the  elas- 
tic moduli  and  high  pressure  behavior  of  the 
cubic  alkaline-earth  oxides,  Phys.  Rev.  B,  33, 
8685-8696,  1986. 

Stull,  D.  R.  and  H.  Prophet  (Eds.),  JANAF  Ther- 
mochemical  Tables,  2nd  ed.,  Office  of  Stan- 
dardReference  Data,  NIST,  Washington, D.C., 
1971. 


GEOPHYSICAL  LABORATORY 


135 


Glass  Diffraction  Measurements  with 
Polychromatic  Synchrotron  Radiation 

Charles  Meade  and  Russell  J.  Hemley 

The  structure  of  liquids  and  glasses  at 
high  pressures  is  an  issue  of  great  impor- 
tance in  the  Earth  and  Material  Sciences.  At 
present,  however,  little  is  known  on  this 
subject  because  of  the  difficulty  of  obtain- 
ing direct  information  about  the  structure 
of  noncrystalline  materials  at  high  pres- 
sures. Spectroscopic  studies  (both  optical 
and  x-ray)  have  provided  constraints  on  the 
vibrational  properties  and  short-range  or- 
der in  glasses  under  pressure,  but  they 
typically  provide  only  an  indirect  probe  of 
structure.  Perhaps  the  most  direct  means  of 
determining  glass  and  liquid  structures  at 
very  high  pressures  (P  >  10  GPa)  is  to 
obtain  x-ray  diffraction  measurements  from 
these  materials  in  the  diamond  cell.  Here, 
we  investigate  the  use  of  high-energy  poly- 
chromatic synchrotron  radiation  for  this 
purpose. 

To  obtain  precise  measurements  of  glass 
diffraction  under  these  conditions,  two  prob- 
lems must  be  addressed.  First,  one  has  to 
know  the  intensity  distribution  of  the  x-ray 
source  to  interpret  the  x-ray  patterns.  Sec- 
ond, in  the  analysis  of  the  measured  dif- 
fraction spectra,  one  must  be  able  to  sub- 
tract the  background  diffracted  intensity 
produced  by  the  diamond  anvils  (both  Bragg 
and  Compton  scattering)  from  the  signal 
due  to  the  amorphous  material. 

In  this  report,  we  will  address  the  first 
question,  constraining  the  x-ray  source  spec- 
tra in  glass  diffraction  experiments.  To  il- 
lustrate a  new  analysis  that  we  have  devel- 


oped for  this  problem,  and  to  demonstrate 
the  advantages  of  using  high  energy  syn- 
chrotron radiation  for  these  experiments, 
we  have  measured  the  x-ray  diffraction 
from  a  small  platelet  of  Si02  glass  under 
ambient  pressures  and  temperatures  at  the 
X-17C  beamline  of  the  National  Synchro- 
tron Light  Source,  Brookhaven  National 
Laboratory  (B  adding  et  a/.,  1990;  Mao  et 
al.  1990).  Because  these  measurements 
were  made  outside  of  the  diamond  cell  on 
a  known  material  (Mozzi  and  Warren,  1 969; 
Konnert  and  Karle,  1973),  they  provide  an 
initial  test  of  our  ability  to  constrain  the 
intensity  distribution  of  the  x-ray  source  in 
our  experiments. 

Examples  of  the  measured  and  known 
diffraction  patterns  for  Si02  glass  are  shown 
in  Figure  76.  From  this  comparison,  it  is 
immediately  evident  that  that  the  shape  of 
the  x-ray  source  distribution  strongly  influ- 
ences the  observed  x-ray  measurements. 
To  normalize  these  diffraction  patterns, 
one  can  measure  the  intensity  of  the  x-ray 
source;  however,  these  measurements  are 
difficult  given  the  the  detector  geometry  at 
X-17C.  One  could  calculate  the  source 
distribution  (e.g.  Krinsky  et  al.,  1983), 
though,  this  requires  precise  knowledge  of 
the  upstream  filters  and  the  critical  energy 
of  the  superconducting  wiggle r.  Moreover, 
calculated  spectra  cannot  account  for  par- 
tial absorption  of  the  beam  by  upstream 
slits. 

Thus,  we  have  adopted  an  alternative 
approach  where  we  take  advantage  of  the 
redundancy  in  our  diffraction  spectra  ob- 
tained at  several  different  scattering  angles. 
Consider  that  the  observed  intensity  spec- 
tra at  particular  diffraction  angle  20/  (cor- 


136 


CARNEGIE  INSTITUTION 


J  tMJM   TH"MJTint   ""  Jlttip   I   I  I   [I   I   H   |Mtl|l 


B 


s.A-1 


Fig.  76.  (A)  Diffraction  pattern  for  Si02  glass  at  ambient  pressures  and  temperatures  measured  from 
monochromatic  x-ray  source  by  Konnert  and  Karle  (1973).  The  horizontal  axis  is  the  wavenumber 
S=47tsm0/X.  (B)  Observed  intensities  (corrected  for  sample  and  air  absorption)  from  Si02  glass 
measured  at  two  different  scattering  angles  (26)  with  a  polychromatic  x-ray  source  at  beamline  X-17C 
of  the  National  Synchrotron  Light  Source.  The  differences  between  these  data  and  with  the  previously 
measured  values  are  due  to  the  distribution  of  x-ray  intensities  in  the  polychromatic  synchrotron  source. 

rected  for  air  and  sample  x-ray  absorption  common  scale  we  define  the  wavenumber 
and  polarization  of  the  diffracted  beam)  s  =  AKsin6lX.  We  can  then  write  equation 
can  be  represented  as  (1 )  as 


/$s(£)  =  [B(28i,  E)  +  C(2Q,  E)]f0(E)  ( i )      /obsW  =  [B(s)  +  C(j)]/o(£,20) 


(2) 


where  B  and  C  are  functions  that  respec- 
tively describe  the  coherent  (Bragg)  and 
incoherent  (Compton)  scattering  from  the 
sample  and  Io  is  the  source 
spectrum.  Whereas  5  and  Care  functions  of 
2  0  and  the  energy  of  the  incident  beam  (E), 
Io  is  a  function  of  E  only.  In  practice,  we 
make  measurements  at  several  diffraction 
angles.  Thus,  to  express  all  of  the  data  on  a 


Here,  B,  C,  and  Iobs  are  functions  of  s  only. 
In  this  representation,  the  source  function 
depends  on  s  and  20.  With  this  transforma- 
tion, the  measurements  at  different  diffrac- 
tion angles  can  be  collapsed  down  to  the 
single  function  Iobs(s).  Even  though/?  and 
C  are  not  known  a  priori,  one  can  use  the 
knowledge  that  they  are  functions  of  s  only 
to  constrain  the  source  function  Iq.  Specif  i- 


Fig.  77.  (A)  Composite  diffraction  pattern  for  Si02  from  measurements  at  seven  angles  ranging  from 
2  0  =  6°  to  45°.  Below  s  =  1 ,  the  data  are  extrapolated  to  zero  intensity.  (B)  The  same  diffraction  spectra 
as  shown  in  Figure  76B.  When  the  spectra  are  normalized  with  the  correct  source  function,  they  are 
equivalent  when  plotted  in  terms  of  the  wavenumber  s.  For  clarity,  the  spectra  are  offset  in  the  vertical 
direction.  (C)  X-ray  source  function  for  X-17C  that  is  consistent  with  these  measurements.  The 
intensities  drop  off  below  12  keV  because  of  absorption  in  upstream  beryllium  and  carbon  filters. 


GEOPHYSICAL  LABORATORY 


137 


1"" 

1 1  ii 

wh 

nrpiii|iiii|iin 

1 1  ■  1 1 1 1 1  ( 1 

pm 

iTrrpTTT 

jm 

Mill 

c 

"\ 

- 

_ 

CO 

c 

CD 

•4— • 

- 

u„J 

■  ml 

Mill 

jj,Liiii lui liin  ' 

.ml.... 

I....I 

111 1I111 1 

7177 

TTTTT 

cally,  we  determine  Io  by  requiring  that 
I0DS{s)  must  be  the  same  for  all  diffraction 
angles. 

An  example  of  a  "composite"  diffrac- 
tion pattern  obtained  in  this  way  and  the 
corresponding  source  function  is  shown  in 
Figure  77.  When  compared  to  the  previ- 
ously measured  pattern,  the  advantages  of 
using  this  approach  is  apparent.  Because 
one  can  measure  diffraction  to  high  ener- 
gies (-60  keV)  at  high  angles,  diffraction 
spectra  can  be  obtained  to  much  higher 
values  of  s  than  in  conventional  laboratory 
measurements.  In  this  study,  we  obtained 
data  to  s  =  26  A"1.  Measurements  to  40  A" 
1  are  possible.  We  expect  that  such  mea- 
surements over  a  wide  range  of  s  values 
will  allow  strong  constraints  on  glass  struc- 
ture at  high  pressures. 


References 

Badding,  J.V.,  H.  K.  Mao,  J.  Z.  Hu,  R.  J.  Hemley, 
C.  Meade,  and  J.  F.  Shu,  High  pressure  energy 
dispersive  x-ray  diffraction  at  X-17C,  EOS, 
71,  1620,  1990. 

Konnert,  J.  H.,  and  J.  Karle,  The  computation  of 
radial  distribution  functions  for  glassy  materi- 
als, Acta.  Cryst.,  A29,  702-710,  1973. 

Krinsky,  S.,  M.  L.  Perlman,  and  R.  E.  Watson, 
Characteristics  of  synchrotron  radiation  and 
of  its  sources,  in  Handbook  on  Synchrotron 
Radiation,  E.  E.  Koch,  ed.,  North  Holland, 
New  York,  pp.  65-171,  1983. 

Mao,  H.  K.,  J.  Z.  Shu,  J.  F.  Shu,  and  R.  J.  Hemley, 
Ultrahigh-pressure  experimentation  above  300 
GPa,  Bull  Ame.r  Phys.  Soc,  36,  529,  1990. 

Mozzi,  R.  L.,  and  B.  E.  Warren,  The  structure  of 
vitreous  silica,  /.  Appl.  Cryst.,  2,  162-172, 
1969. 


10       20       30       40        50 
Energy,  keV 


60       70 


138 


CARNEGIE  INSTITUTION 


X-ray  Diffraction  of  Solid  Nitrogen- 
Helium  Mixtures* 

Willem  L.  Vos,  Larry  W.  Finger, 
Russell  J.  Hemley,  Ho-Kwang  Mao, 

Jing  Zhu  Hu,  Jin  Fu  Shu, 

Richard LeSar* Andre  de  Kuijper, * 

and  Jan  A.  Schouten** 

The  study  of  simple  molecular  mixtures 
in  their  solid  phases  at  high  pressures  has 
recently  become  of  increasing  interest.  This 
work  is  important  for  statistical  and  con- 
densed matter  physics,  specifically  to  in- 
vestigate the  influence  of  size  ratio,  pack- 
ing, and  van  der  Waals-like  interactions  on 
the  stability  of  mixed  structures.  Such  stud- 
ies also  provide  a  starting  point  for  for 
modeling  the  interiors  of  the  outer  planets 
of  our  solar  system. 

One  of  the  most  studied  mixtures  in  this 
respect  is  N2-He;  helium  is  still  a  fluid 
under  conditions  where  nitrogen  has  so- 
lidified (see  Fig.  78).  At  room  temperature, 
their  freezing  pressures  differ  by  a  factor  of 
5:  i.e.,  24  vs.  120  kbar).  By  a  combination 
of  visual  observations  and  quasi-isochoric 
p-T  scans,  Vos  and  Schouten  (1990)  found 
that  around  10  mol%  He  is  soluble  in  the 
orientationally  ordered  rhombohedral  e  - 
phase  of  nitrogen  (Mills  et  ai,  1986);  in 
contrast,  the  solubility  of  helium  in  the 
disordered  (3  or  8  -  phases  (Cromer  et  ai, 

*This  collaboration  was  supported  by  a  NATO 
Collaborative  Research  Grant. 
*  T-ll,  Los  Alamos  National  Laboratory,  Los 
Alamos  New  Mexico  87545 

van  der  Waals-Zeeman  Laboratorium, 
Universiteit  van  Amsterdam,  1018XE  Amsterdam, 
The  Netherlands. 


1981)  is  negligible.  Due  to  this  solubility, 
the  stability  of  the  e-phase  was  found  to 
increase  drastically  with  respect  to  the  8  - 
phase  (see  Fig.  78).  The  shift  of  the  8  -  e 


100 


1 

J* 
<xi 

3 


CD 

k— 
CL 


250  350  450  550 

Temperature,  K 

Fig.  78.  PTphase  diagrams  of  N2  (solid  lines),  He 
(short  dashed  line),  and  projected  phase  diagram 
of  N2-He  (dashed  lines).  The  solid  lines  are  in 
order  of  increasing  pressure  the  P-fluid,  5-p  and 
e-8  lines,  the  short  dashed  line  is  the  melting  line 
of  helium  and  the  dashed  lines  are  the  three  -  phase 
lines  e-8  -F  of  the  mixture. 


transition  was  subsequently  confirmed  by 
vibrational  Raman  spectroscopy 
(Scheerboom  and  Schouten,  1991).  Here, 
we  report  preliminary  investigations  of  the 
structure  of  the  mixed  phases  (indicated  by 
asterisks)  by  x-ray  diffraction  and  a  com- 
parison of  the  results  with  computer  simu- 
lations. 

X-ray  diffraction  studies  were  performed 
at  beamline  X-17C  of  the  National  Syn- 
chrotron Light  Source,  at  Brookhaven  Na- 
tional Laboratory,  using  a  single-crystal 
type  diamond  anvil  cell  (Mao  and  Bell, 
1980).  The  samples  were  prepared  from 
high  purity  (99.999%)  gases  and  loaded  in 


GEOPHYSICAL  LABORATORY 


139 


a  pressure  vessel  at  a  pressure  of  about  2 
kbar.  All  experiments  were  performed  at 
room  temperature,  and  pressures  were  ob- 
tained from  the  linear  ruby  scale.  Three 
experiments  were  performed:  one  on  pure 
N2,  to  check  the  structure  of  the  e-phase, 
since  this  had  been  obtained  from  powder 
diffraction  (Mills  et  al.,  1986,  Olijnyk, 
1990),  and  two  on  the  8*  -  phase  at  compo- 
sitions of  5.0(2)  and  10.0(2)  mol%  helium. 
The  diffraction  experiment  on  e-N2  was 
performed  at  195  kbar  since  the  transition 
from  8  to  e  takes  place  at  1 65  kbar  (Olijnyk, 
1990).  The  sample  showed  clear  colors 
under  crossed  polarizers,  which  distin- 
guishes it  from  the  non-birefringent  8  - 
phase.  The  sample  consisted  of  at  least  one 
large  crystal  and  many  small  grains,  since 
strong  reflections  were  observed  at  distinct 
(X,oS)  angles,  while  a  powder-like  pattern 
with  preferred  orientation  was  observed 
irrespective  of  the  ix,(o)  angles.  This  con- 
figuration was  often  encountered  and  was 
named  "single-powder."  The  d-spacings  of 
the  powder  pattern  agree  with  the  ones 
reported  previously  (Mills  et  aL,  1986; 
Olijnyk,  1990)  and  yield  a  unit  cell  of 
dimensions  a=  7.63(4)  A,  c=  10.37(10)  A, 
c/a=  1.36  with  24  molecules  and  space 

group  R3c  (Mills  et  ai,  1986).  Ten  reflec- 
tions were  observed  from  the  single  crys- 
tal, that  were  also  consistent  with  the  R3c 
structure.  Special  attention  was  paid  to 
overtones  and  to  possible  lower-order  re- 
flections, since  some  computer  simulation 
results  yielded  superstructures  with  unit 
cells  containing  up  to  64  molecules  (Nose 
and  Klein,  1986,  Belak  era/.,  1990).  How- 
ever, no  longer  spacings  were  found  and 
only  a  few  weak  overtones  were  observed, 


also  consistent  with  the  aforementioned 
structure. 

With  5  mol  %  He,  experiments  were 
performed  at  126  and  144  kbar.  The  pres- 
ence of  the  e*  -  phase  was  checked  by  the 
observation  of  birefringence.  The  sample 
consisted  again  of  a  "single-powder."  The 
spectrum  of  d-spacings  consisted  of  lines 
from  both  8*  and  £*  -  phases.  This  was 
confirmed  by  Raman  scattering,  which 
showed  that  the  vi  mode  was  clearly  split, 
which  is  not  the  case  in  the  pure  phases.  The 
spacings  of  the  8*  -  phase  are  the  same  as 
those  of  the  8  -  phase  at  the  same  pressures 
(Olijnyk,  1990),  which  means  that  there  is 
negligible  solubility  of  helium  in  this  phase. 
The  other  d-spacings  could  all  be  fitted  to  a 
hexagonal  phase  with  a  -  8.050(5)  A,  c  - 
9.469(12)  A,  da  =  1.176  at  126  kbar  and 
7.959(5)  A,  9.353(17)  A,  1.175  respec- 
tively at  144  kbar.  Since  the  intensities 
from  the  8*  -  phase  are  large  both  in  the  x- 
ray  diffraction  and  the  Raman  scattering 
experiments,  this  phase  must  be  a  signifi- 
cant portion  of  the  volume.  From  mass- 
balance  considerations,  the  composition  of 
the  e*  -  phase  is  clearly  larger  than  5  mol  % 
He. 

The  experiment  on  e*  with  1 0  %  He  was 
performed  at  92  kbar.  Before  the  experi- 
ment was  done,  the  sequence  of  transitions 
leading  to  this  phase  was  verified,  as  well 
as  the  vibrational  Raman  spectrum.  The 
sample  consisted  of  several  crystals,  but  no 
additional  powder,  probably  due  to  the 
presence  of  a  small  amount  of  the  helium- 
rich  fluid  phase.  Strong  reflections  includ- 
ing series  of  overtones  were  obtained  (see 
Fig.  79).  From  the  measured  d-spacings,  a 
hexagonal  unit  cell  with  a  -  8.258(2)  A,  c 


140 


CARNEGIE  INSTITUTION 


100000 


'  t   t  t  I 


■ 


30 


50 


70 


Energy  (keV) 


Fig.  79.  Diffraction  pattern  of  e*  with  10  %  helium 
taken  at  26=9°.  The  fundamental  reflection  of  the 
(101)  class  at  13.7  keV  and  4  overtones  at  27.3, 
41.1,  54.7  and  68.4  keV  are  indicated  by  the 
arrows.  The  peaks  near  18  keV  are  escape  peaks 
from  the  first  overtone  and  the  other  peaks  origi- 
nate from  the  gasket. 


=  9.747(5)  A  and  c/a=1.180  was  obtained, 
similar  to  the  results  at  5  mol%  He.  Four 
reflections  could  be  attributed  to  one  crys- 
tal and  five  reflections  to  a  second  one. 

The  reflections  of  the  e*  -  phase  do  not 
fulfill  the  condition  -h+k+l=3n  of  rhombo- 
hedral  structures.  Therefore,  we  can  rule 

out  the  R3c  space  group  for  this  phase. 
However,  the  contents  of  the  unit  cell  can 
be  estimated  on  the  basis  of  the  available 
information.  We  assume  the  phase  to  be 
stoichiometric  and  calculate  the  free  en- 
thalpy difference  between  this  phase  and 
the  coexisting  phases,  using  known  p(V) 
isotherms  of  N2  (Olijnyk,  1 990)  and  He  (Le 
Toullec  et  ai,  1990)  and  making  assump- 
tions for  the  volumes  of  the  mixture  that  are 
justified  by  theory  and  simulations.  It  then 
turns  out  that  the  e*  -  phase  becomes  stable 
when  it  contains  at  least  22  N2  molecules 
and  2  He  atoms  per  unit  cell.  This  also 
yields  a  volume  difference  with  the  coex- 


isting phases  that  is  reasonable  in  view  of 
the  pressure  jumps  measured  previously  at 
transitions  involving  this  phase  (Vos  and 
Schouten,  1990;  Vos,  1991).  Furthermore, 
this  is  also  consistent  with  the  fact  that  the 
Raman  shifts  of  this  phase  are  very  similar 
to  those  of  the  8  and  e  -  phases  at  the  same 
pressure  (Scheerboom  and  Schouten,  1 99 1 ). 
Computer  simulations  were  performed 
using  a  variable  shape  simulation  cell 
(Rahman-Parinello  method)  to  allow  for 
crystal  transformations.  The  best  available 
potentials  for  N2  and  He  were  used  (for 
details  see  de  Kuijper,  1991).  At  300  K  and 
pressures  of  200  and  300  kbar,  it  was  found 

that  pure  N2  remains  in  the  R3c  space 
group.  This  result  was  obtained  with  both 
hexagonal  and  rhombohedral  simulation 
cells,  which  would  have  permitted  the  ap- 
pearance of  the  alternative  structures  that 
were  observed  earlier  (Nose  and  Klein, 
1986;  Belak  et  al,  1990)  and  therefore 
agrees  with  the  experiment.  For  compari- 
son with  pure  N2,  simulations  were  per- 
formed on  N2-He,  starting  from  the  R3c 
space  group.  It  turns  out  that  this  structure 
is  maintained,  indicating  that  it  is  at  least 
metastable  for  the  mixture.  An  interesting 
result  is  that  the  da  drops  considerably, 
from  1.30  in  pure  N2  to  1.17  at  10  mol  % 
He. 

From  the  present  x-ray  experiments,  it 
can  be  concluded  that  at  room  temperature 
there  is  a  negligible  solubility  of  He  in  the 
8  -  phase  of  N2.  Furthermore,  the  solubility 
in  the  e*  -  phase  is  clearly  larger  than  5  mol 
%,  in  agreement  with  previous  experiments. 
However,  it  turns  out  that  He  stabilizes  a 
phase  with  a  different  structure  than  pure  e- 


GEOPHYSICAL  LABORATORY 


141 


N2.  This  phase  has  a  smaller  c/a  ratio  than 
pure  N2  at  the  same  pressure,  which  is 
consistent  with  computer  simulations. 

If  the  e*  -  phase  turns  out  to  be  stoichio- 
metric, the  behavior  of  N2-He  documented 
here  bears  some  similarity  with  that  found 
for  colloidal  suspensions.  There,  it  was 
recently  found  that  mixtures  with  a  size 
ratio  of  0.61  (cf.,  roughly  0.6  for  helium 
and  nitrogen)  form  a  stoichiometry  with  a 
structure  that  is  not  encountered  in  the  pure 
components  (Bartlett  et  al.,  1990). 


Scheerboom,  M.  I.  M.,  and  J.  A.  Schouten,  Detec- 
tion of  the  e-5  phase  transition  in  N2  and  the 
N2-He  mixture  by  Raman  spectroscopy:  new 
evidence  for  the  solubility  of  fluid  He  in  solid 
N2,/.  Phys.:  Condens.  Matter,  in  press,  1991. 

Vos,  W.  L.,  Phase  equilibria  in  simple  systems  at 
high  pressure,  Ph.  D.  dissertation,  Universiteit 
van  Amsterdam,  1991. 

Vos,  W.  L.,  and  J.  A.  Schouten,  Solubility  of  fluid 
helium  in  solid  nitrogen  at  high  pressure,  Phys. 
Rev.  Lett.,  64,  898-901,  1990. 


Evidence  for  Orientational  Ordering  of 
Solid  Deuterium  at  High  Pressures* 

Russell  J.  Hemley  and  Ho-Kwang  Mao 


References 

Bartlett,  P.,  R.  H.  Ottewill,  and  P.  N.  Pusey, 
Freezing  of  binary  mixtures  of  colloidal  hard 
spheres,/.  Chem.  Phys. ,93, 1299-1312, 1990. 

Belak,  J.,  R.  LeSar,  and  R.  D.  Etters,  Calculated 
thermodynamic  properties  and  phase  transi- 
tions of  solid  N2  at  temperatures  0<T<300  K 
and  pressures  0<p<100  GPa,  /.  Chem.  Phys., 
92,5430-5441,1990. 

Cromer,  D.  T.,  R.  L.  Mills,  D.  Schiferl,  and  L.  A. 
Schwalbe,  The  structure  of  N2  at  49  kbar  and 
299  K,  Acta  Cryst.  B37,  8-11,  1981. 

de  Kuijper,  Jhr.  A.,  Computer  simulations  of  phase 
equilibria  in  molecular  systems,  Ph.  D.  disser- 
tation, Universiteit  van  Amsterdam  1991. 

Le  Toullec,  R.,  P.  Loubeyre,  and  J.  -  P.Pinceaux, 
Refractive-index  measurements  of  dense  he- 
lium up  to  16  GPa  at  T=298K:  Analysis  of  its 
thermodynamic  and  electronic  properties, 
Phys.  Rev.  B40,  2368-2378,  1990. 

Mao,  H.  K.,  and  P.  M.  Bell,  Design  and  operation 
of  a  diamond-window,  high-pressure  cell  for 
the  study  of  single-crystal  samples  loaded 
cryogenically,  Carnegie  Instn.  Washington 
Year  Book,  79,  409-411,  1980. 

Mills,  R.  L.,  D.  T.  Cromer,  B.  dinger,  Structures 
and  phase  diagrams  of  N2  and  CO  to  1 3  GPa  by 
x-ray  diffraction,  /.  Chem.  Phys.,  84,  2837- 
2845,  1986. 

Nose,  S.,  and  M.  L.  Klein,  Constant-temperature  - 
constant-pressure  molecular-dynamics  calcu- 
lations for  molecular  solids:  Application  to 
solid  nitrogen  at  high  pressure,  Phys.  Rev. 
B33,  339-342,  1986. 

Olijnyk,  H.,  High-pressure  x-ray  diffraction  stud- 
ies on  solid  N2  up  to  43.9  GPa,  /.  Chem.  Phys., 
93,  8968-8972,  1990. 


At  low  pressures  hydrogen  forms  an 
insulating  molecular  solid  with  the  mol- 
ecules in  states  of  complete  rotational  dis- 
order over  a  wide  range  of  temperature. 
With  increasing  pressure,  the  rotational 
motion  of  the  molecules  is  expected  to 
become  more  restricted,  ultimately  leading 
to  orientational  ordering.  Detailing  the  na- 
ture of  possible  ordering  transitions  is  im- 
portant for  understanding  the  mechanism 
of  pressure-induced  metallization,  a  transi- 
tion with  important  implications  for  both 
condensed-matter  and  planetary  physics. 
Raman  measurements  of  the  high-fre- 
quency intramolecular  vibrational  mode 
(vibron)  indicate  that  the  molecular  solid 
remains  stable  to  at  least  -250  GPa  but 
undergoes  a  phase  transition  at  150  GPa  (at 
77  K;  Hemley  and  Mao,  1988).  The  low- 
frequency  vibrational  spectrum  provides 
information  on  the  state  of  ordering  in  the 
solid  and  constraints  on  the  crystal  struc- 

*  This  work  was  supported  by  NSF  (DMR-89 
12226  and  EAR-8904080)  and  NASA  (NAGW- 
1722). 


142 


CARNEGIE  INSTITUTION 


ture  at  these  pressures,  which  are  beyond 
the  range  of  current  x-ray  diffraction  tech- 
niques (Mao  etal.,  1988).  Previously,  we 
reported  measurements  of  the  evolution  of 
the  low-frequency  rotational  bands  and  lat- 
tice phonon  of  hydrogen  to  162  GPa  at  77- 
295  K  (Hemley  et  al.y  1990a).  Over  this 
pressure  interval  the  rotational  bands  per- 
sist but  broaden  and  the  lattice  phonon, 
which  correlates  with  the  Z?2g  optical  pho- 
non of  the  hexagonal-close  packed  struc- 
ture, shifts  continuously.  The  continuity  of 
the  low-frequency  bands  as  a  function  of 
pressure  indicates  that  an  underlying  hex- 
agonal structure  persists  into  the  high-pres- 
sure phase  above  1 50  GPa  . 

Examination  of  the  low-frequency  spec- 
trum of  deuterium  is  important  for  under- 
standing isotope  effects  on  a  variety  of 
properties  of  hydrogen  at  high  densities. 
Orientational  ordering  is  energetically  fa- 
vored in  the  heavier  isotope  as  a  result  of  its 
smaller  rotational  constant  (#D2  =  29.9 
cm-1  versus  #H2  =  59.3  cm-1  in  the  gas 
phase),  which  results  in  a  stronger  mixing 
of  free  molecule  rotational  states  in  con- 
densed phase.  At  low  densities,  changes  in 
temperature  result  in  variations  in  the  rela- 
tive population  of  ortho  and  para  species 
(even  and  odd  /  rotational  states).  How- 
ever, at  high  densities  the  single  molecule 
ortho-para  distinction  breaks  down  as  a 
result  of  mixing  of  rotational  states,  and 
this  is  expected  to  occur  at  lower  densities 
in  D2.  Evidence  for  this  is  found  in  the 
differences  in  the  pressure  at  which  sym- 
metry breaking  occurs  in  the  7=0  solids  at 
very  low  temperatures  (Silvera  and 
Wijngaarden,  1981).  Isotope  effects  are 
also  observed  in  the  pressure  dependence 


Hydrogen 
93.8  GPa 


200    400    600    800    1000 

Wavenumber,  cm"1 


1200 


Fig.  80.  Examples  of  low-frequency  Raman  spec- 
tra of  hydrogen  and  deuterium  at  77  K.  The  two 
broadened  low-frequency  bands  observed  in  H2 
correlate  with  the  So(0)  and  S\(0)  rotational  tran- 
sitions. The  intense  band  observed  in  D2  at  250 
cm-1  is  identified  as  a  libra tional  mode  in  an 
orientationally  ordered  structure.  The  optical  pho- 
nons  observed  in  both  isotopes  are  indicated. 

of  the  molecular  vibron;  the  shift  is  signifi- 
cantly stronger  in  hydrogen  relative  to  deu- 
terium (see  Hemley  et  al.,  1991).  It  is 
therefore  of  interest  to  determine  whether 
or  not  this  difference  is  associated  with 
structural  differences  between  the  two  iso- 
topes. Finally,  a  distinct  isotope  effect  is 
observed  in  the  pressure  of  the  low-tem- 
perature 150-GPa  phase  transition,  which 
is  >10  GPa  higher  in  deuterium.  Under- 
standing the  origin  of  this  effect  is  of  inter- 
est because  the  transition  appears  to  be 
associated  with  changes  in  electronic  prop- 


GEOPHYSICAL  LABORATORY 


143 


erties,  such  as  metallization  (Hemley  etal., 
1990a;  Hemley  and  Mao,  1990). 

Samples  were  loaded  at  room  tempera- 
ture in  a  modified  Mao-Bell  diamond-cell 
with  composite  rhenium/T301 -stainless 
steel  gaskets.  Here  we  report  the  results  of 
four  separate  experiments  on  deuterium 
carried  out  at  pressures  from  20  GPa  to 
above  100  GPa  and  temperatures  between 
77  K  and  295  K.  Raman  spectra  were 
measured  using  optical  techniques  de- 
scribed previously  (Hemley  and  Mao,  1988; 
Hemley  et  aL,  1990a).    Low-frequency 
Raman  spectra  show  a  strong,  weakly  pres- 
sure dependent,  band  at  240  cm-1,  together 
with  a  weaker  peak  at  higher  frequency 
which  exhibits  a  large  pressure  dependence. 
A  representative  low-frequency  Raman 
spectrum  of  deuterium  at  99  GPa  is  com- 
pared with  that  measured  for  hydrogen  at 
similar  pressures  in  Fig.  80.  On  the  basis  of 
x-ray  diffraction  measurements  (Mao  et 
aL,  1988;  Hemley  et  aL,  1990b)  and  the 
continuity  of  the  spectra  with  increasing 
pressure,  the  higher  frequency  band  is  iden- 
tified as  the  E2g  Raman-active  phonon  in 
the  hexagonal-close  packed  structure, 
analogous  to  that  found  for  hydrogen. 

The  low-frequency  spectrum  of  hydro- 
gen at  77  K  below  100  GPa  is  characterized 
by  two  broadened  rotational  bands  [So(0) 
and  Si(0),  corresponding  to  the  AJ  =  2,  /  = 
0-2  and  AJ = 2,  /=  1  -3  excitations  in  the  free 
molecule].  The  persistence  of  these  bands 
to  100  GPa  was  interpreted  as  an  indication 
of  free  (or  nearly  free)  rotation  of  the  mol- 
ecules. In  contrast,  the  low-frequency  band 
observed  in  D2  does  not  fit  a  rotational 
transition:  i.e.,  the  frequency  of  the  band  is 
240  cm~l,  whereas  the  frequencies  of  the 


So(0)  and  Si(0)  occur  at  6£  =  180  cm"1  and 
10#  =  300  cm-1.  We  interpret  this  band  as 
indicative  of  a  new  high-pressure  molecu- 
lar phase  of  deuterium.  In  view  of  the 
presence  of  the  optical  phonon,  we  suggest 
that  the  phase  is  an  ordered  (or  partially 
ordered)  form  with  a  hexagonal  close- 
packed  structure  and  that  the  low-frequency 
band  is  associated  with  librational  motion. 
The  latter  is  consistent  with  its  weak  pres- 
sure dependence. 

Further  evidence  for  ordering  is  found 
in  the  frequency  shift  of  the  optical  phonon. 
The  volume  dependence  of  the  optical  pho- 
non frequency  for  the  two  isotopes  is  shown 
in  Fig.  81.  The  D2  curve  is  parallel  to  that 
obtained  by  Wijngaarden  et  aL  (1983)  at 


E 
o 

i_r 

CD 

.Q 

E 

c 

CD 

> 

cd 


PHONON 
77  K 


200  _ 


Volume  (cm3/mol) 

Fig.  81.  Volume  dependence  of  the  optical  pho- 
non for  H2  and  D2:  squares,  present  work;  circles, 
Wijngaarden  etal.  (1983).  The  volume  was  calcu- 
lated from  the  pressure  using  the  experimental 
equation  of  state  determined  to  26.5  GPa  at  room 
temperature  (Mao  et  aL,  1988;  Hemley  et  aL, 
1990).  The  dotted  line  shows  the  frequencies 
expected  for  D2  on  the  basis  of  the  measurements 
forH2.  The  data  of  Wijngaarden  etal.  (1983)  have 
been  corrected  using  the  new  equation  of  state. 


144 


CARNEGIE  INSTITUTION 


much  lower  pressures,  as  noted  previously 
for  H2  (Hemley  et  al,  1990a).    At  low 
compressions,  the  frequencies  of  the  modes 
differ  by  a  factor  of  V2  as  a  result  of  the 
differences  in  masses  [i.e.,  (mnz/^D2)^], 
as  expected  if  the  modes  (assumed  to  be 
harmonic),  crystal  structure,  and  volume 
are  identical  for  the  two  isotopes.    It  is 
evident,  however,  that  this  relationship  does 
not  hold  at  higher  compressions:  the  D2 
curve  is  higher  than  that  expected  on  the 
basis  of  the  measurements  for  H2.    This 
offset  may  arise  from  a  modification  of  the 
crystal  structure  by  ordering,  which  could 
affect  the  frequency  of  the  optical  phonon 
owing  to  changes  in  intermolecular  inter- 
actions in  the  ordered  state  (even  at  con- 
stant volume).  Alternatively,  we  note  that 
orientational  ordering  should  result  in  a 
more  efficient  packing  of  the  molecules 
relative  to  the  rotationally  disordered  state, 
and  that  the  frequency  of  the  phonon  is  a 
strong  function  of  volume.  Thus,  a  second 
possibility  is  that  the  offset  indicates  that 
the  molar  volume  of  deuterium  (which  is 
ordered)  is  lower  than  that  of  hydrogen 
(which  appears  not  to  be  fully  ordered)  at 
the  same  pressures.  This  needs  to  be  exam- 
ined by  low-temperature  x-ray  diffraction. 
The  available  diffraction  data  provide  some 
evidence  for  a  lower  volume  in  D2  even  at 
room  temperature  (-2%  at  30  GPa),  so  it  is 
possible  that  effects  of  rotational  ordering 
are  present  at  room  temperature  (Fig.  81). 
It  should  be  pointed  out  that  if  the  volume 
difference  at  77  K  persists  to  higher  pres- 
sure (>150  GPa),  it  may  contribute  to  the 
isotope  effect  on  the  pressure  of  the  high- 
pressure  phase  transition  (i.e.,  the  low- 
pressure  phase  is  stabilized  to  higher  pres- 


sures in  the  heavier  isotope). 

The  results  may  be  compared  with  the 
measurements  of  Silvera  and  Wijngaarden 
(1981),  who  studied  ortho-D2  at  5  K  to  54 
GPa.  They  reported  the  observation  of  a 
broadened  low-frequency  band  at  220-240 
cm-  *  above  28  GPa,  which  they  interpreted 
as  arising  from  an  ordering-type  transition 
(broken  symmetry  transition  in  the  J  =  0 
molecules).  Since  Silvera  and  Wijngaarden 
(1981)  were  unable  to  measure  the  optical 
phonon  above  the  transition,  they  proposed 
that  the  high-pressure  phase  has  the  cubic 
Pa3  structure.  The  present  observations  of 
the  optical  phonon  indicate  that  the  struc- 
ture of  the  solid  at  77  K  (and  above)  is  not 
Pa3  because  the  Raman-active  excitations 
in  this  structure  comprise  only  librational 
modes  (phonon  is  inactive).  The  close  simi- 
larity in  the  librational  modes  measured  in 
the  two  studies  strongly  suggests  that  the 
transition  observed  at  5  K  also  takes  place 
within  the  hep-type  structure.  This  assign- 
ment is  consistent  with  the  results  of  recent 
theoretical  calculations  which  indicate  that 
structures  based  on  hep  are  stable  relative 
to  cubic  (e.g.,  Pa3)  at  high  densities  (Raynor, 
1987;  Barbee  etai,  1989;  Ashcroft,  1991; 
Kaxirase/a/.,  1991). 

Kaxiras  et  al.  ( 1 99 1 )  have  performed  an 
extensive  series  of  calculations  of  different 
molecular  ordering  schemes  within  hep.  A 
new  class  of  oriented  hexagonal  structures 
based  on  a  herring-bone  type  configuration 
has  been  found  to  be  energetically  favored 
and  to  have  larger  band  gaps  than  that  of  the 
structure  assumed  in  previous  work.  We 
propose  that  the  structure  of  the  phase  of  D2 
observed  here  is  closely  related  to  these 
structures.  This  assignment  is  consistent 


GEOPHYSICAL  LABORATORY 


145 


Deuterium 

130GPa 

77  K 


138  GPa 
295  K 


200  400  600  800         1000 

Wavenumber,  cm-1 


1200 


Fig.  82.  Raman  spectrum  of  deuterium  at  130-138 
GPa  at  77  K  and  295  K. 


with  the  experimental  evidence  for  a  band 
gap  (insulating  state)  to   high   pressures 
(i.e.,  to  at  least  -150  GPa),  which  is  one  of 
the  key  problems  with  previously  proposed 
ordered  hexagonal  structures  [see,  Ashcroft 
(1991)  and  Kaxiras  etal.  (1991)].  Confir- 
mation of  this  structure  should  be  possible 
by  low-temperature  single-crystal  x-ray 
diffraction.  Further  work  is  also  required  to 
determine  the  P-T  stability  field  of  the 
phase  as  function  of  pressure,  temperature, 
and  ortho-para  state  (at  lower  pressures). 
The  librational  band  weakens  gradually 
with  increasing  pressure  above  100  GPa, 
and  diamond  fluorescence  tends  to  increase 
at  these  pressures.  As  a  result,  the  phonon 
could  not  be  measured  above  100  GPa, 
although  the  stronger  low-frequency  band 
is  readily  apparent  (Fig.  82).  With  increas- 
ing pressures  above  -140  GPa  at  77  K  the 


band  appears  to  weaken  somewhat  but  was 
observed  through  the  high-pressure  phase 
transition  at  -165  GPa,  despite  the  marked 
discontinuity  in  the  vibron  frequency 
(Hemley  etal.,  1991).  Hence,  D2  is  appar- 
ently ordered  over  the  entire  pressure  inter- 
val of  this  study  at  77  K,  although  the 
broadening  of  the  band  at  room  tempera- 
ture (Fig.  82)  may  indicate  that  the  mol- 
ecules are  disordered  at  higher  tempera- 
tures. At  still  higher  pressures,  an  increase 
in  scattering  intensity  is  observed  in  the 
vicinity  of  the  low-frequency  band,  a  de- 
tailed study  of  which  will  be  presented 
elsewhere. 


References 

Ashcroft,  N.  W.,  Optical  response  near  a  band 
overlap:  Application  to  dense  hydrogen,  in 
Molecular  Systems  under  High  Pressure,  R. 
Pucci  and  G.  Piccitto,  eds.,  pp.  201-222, 
Elsevier,  Amsterdam,  1991. 

Barbee,  T.  W.,  A.  Garcia,  M.  L.  Cohen,  and  J.  L. 
Martins,  Theory  of  high-pressure  phases  of 
hydrogen,  Phys.  Rev.  Lett.  62,  1150-1153, 
1990. 

Hemley,  R.  J.,  and  H.  K.  Mao,  Phase  transition  in 
solid  molecular  hydrogen  at  ultrahigh  pres- 
sures, Phys.  Rev.  Lett.,  61,  857-860,  1988. 

Hemley ,  R.  J.,  andH.  K.  Mao,  Critical  behavior  in 
the  hydrogen  insulator-metal  transition,  Sci- 
ence, 249,  391-393,  1990. 

Hemley,  R.  J.,  H.  K.  Mao,  and  J.  F.  Shu,  Low- 
frequency  vibrational  dynamics  and  structure 
of  hydrogen  at  megabar  pressures,  Phys.  Rev. 
Lett.  65,  2670-2673,  1990a. 

Hemley,  R.  J.,  H.  K.  Mao,  L.  W.  Finger,  A.  P. 
Jephcoat,  R.  M.  Hazen,  and  C.  S.  Zha,  Equa- 
tions of  state  of  solid  hydrogen  and  deuterium 
from  single-crystal  x-ray  diffraction  to  26.5 
GPa,  Phys.  Rev.  B  42,  6458-6470,  1990b. 

Hemley,  R.  J.,  H.  K.  Mao,  and  M.  Hanfland, 
Spectroscopic  investigations  of  the  insulator- 
metal  transition  in  solid  hydrogen,  in  Molecu- 
lar Systems  under  High  Pressure,  R.  Pucci 
and  G.  Piccitto,  eds.,  pp.  223-243,  Elsevier, 
Amsterdam,  1991. 

Kaxiras,  E.,  J.  Broughton,  and  R.  J.  Hemley,  Onset 
of  metallization  and  related  transitions  in  solid 


146 


CARNEGIE  INSTITUTION 


hydrogen,  Phys.  Rev.  Lett.,  67,  1138-1141, 
1991. 

Mao,  H.  K.,  A.  P.  Jephcoat,  R.  J.  Hemley,  L.  W. 
Finger,  C.  S.  Zha,  R.  M.  Hazen,  andD.  E.  Cox, 
Synchrotron  x-ray  diffraction  measurements 
of  single  crystal  hydrogen  to  26.5  GPa,  Sci- 
ence 239,  1131-1134,1988. 

Raynor,  S.,  Novel  ab  initio  self-consistent-field 
approach  to  molecular  solids  under  pressure. 
II.  Solid  H2  under  high  pressure,  /.  Chem. 
Phys.,  87,  2795-2799,  1987. 


Silvera,  I.  F.,  and  R.  J.  Wijngaarden,  New  low- 
temperature  phase  of  molecular  deuterium  at 
ultrahigh  pressure,  Phys.  Rev.  Lett.,  47, 39-42, 
1981. 

Wijngaarden,  R.  J.,  V.  V.  Goldman,  and  I.  F. 
Silvera,  Pressure  dependence  of  the  optical 
phonon  in  solid  hydrogen  and  deuterium  up  to 
230  kbar,  Phys.  Rev.,  B  27,  5084-5087,  1983. 


GEOPHYSICAL  LABORATORY 


147 


BlOGEOCHEMISTRY 


Nitrogen  Isotope  Tracers  of 
Atmospheric  Deposition  in  Coastal 
Shelf  Waters  off  North  Carolina. 

Marilyn  L.  Fogel  and  Hans  W.  Paerl  * 

Nitrogen  plays  a  key  role  in  regulating 
marine  primary  and  secondary  production 
both  on  regional  and  global  scales  (Ryther 
and  Dunstan,  1971;  Nixon  et  al,  1986). 
Nitrogen  sources  in  aquatic  ecosystems 
may  be  external  ("new")  or  internal  ("re- 
cycled"). The  relative  utilization  of  "new" 
and  "recycled"  N  inputs  by  phytoplankton 
is  important  in  determining  levels  of  pri- 
mary and  secondary  production  in  coastal 
ecosystems.  The  need  for  a  more  detailed 
understanding  of  the  dynamics  of  new  vs. 
recycled  production  in  coastal  waters  is 
pressing,  because  previously  pristine  seg- 
ments of  the  coastal  oceans  are  now  exhib- 
iting both  incipient  and  advanced  stages  of 
eutrophication  (Cosper  etaL,  1987;  Paerl, 
1988).  Terrigenous  point  and  nonpoint 
inputs  have  traditionally  been  identified  as 
the  most  likely  nutrient  sources  supporting 
new  production  in  heavily  impacted,  eutro- 
phic  estuaries,  including  the  Chesapeake, 
San  Francisco,  Delaware,  and  Narraganset 
Bays  (Boynton  et  al.,  1982).  The  connec- 
tion between  man's  watershed  activities 
and  coastal  eutrophication,  however,  ap- 


Institute  of  Marine  Sciences,  University  of  North 
Carolina 


pears  more  cryptic  in  many  other  places 
(e.g.,  South  Atlantic  Bight).  Recently,  ni- 
trogen deposition  from  rainfall  has  been 
shown  to  stimulate  primary  production  in 
coastal  waters  adjacent  to  North  Carolina 
(Paerl,  1985;  Paerl  et  a/.,  1990). 

Atmospheric  deposition,  as  wet  and 
dryfall,  is  an  increasingly  important  source 
of  biologically-usable  nitrogen  in  estua- 
rine  and  coastal  regions  (Paerl,  1985; 
Legendre  and  Gosselin,  1989).  Large  East 
Coast  estuaries  and  certain  European  seas 
currently  receive  about  20-50%  of  their 
combined  nitrogen  loading  from  atmo- 
spheric sources  (Fisher  etaL,  1988;  Prado- 
Fiedler,  1990;  Loye-Pilot  et  al,  1990). 
Nitrogen  from  atmospheric  deposition  (AD) 
may  be  a  unique  source  in  coastal  waters,  as 
direct  surface  water  deposition  may  occur 
downstream  of  estuarine  zones  where  much 
of  the  terrigenous  nitrogen  has  been  as- 
similated. Proper  identification  of  differ- 
ent N  sources  and  their  fluxes  are  of  prime 
concern  in  understanding  and  ultimately 
controlling  recent  eutrophication  problems 
in  coastal  ecosystems. 

We  have  tested  the  possibility  that  AD 
represents  both  a  unique  and 
biogeochemically  significant  source  of  ni- 
trogen supporting  new  production  in  North 
Carolina's  coastal  Atlantic  waters  by  trac- 
ing the  fate  of  AD  products  into  natural 
phytoplankton  assemblages  through  the  use 
of  stable  N  isotope  signatures.  Nitrogen 
isotopes  at  the  natural  abundance  level 
have  been  used  extensively  in  tracing  ei- 
ther biochemical  processes  or  sources  of 


148 


CARNEGIE  INSTITUTION 


food  in  complex  ecosystems  (Owens,  1 987). 
A  study  by  Showers  et  al.  (1990)  in  the 
Neuse  River  of  North  Carolina  demon- 
strated distinct  nitrogen  sources  on  the  ba- 
sis of  isotopic  composition.  The  isotope 
ratio  of  the  nitrate  from  sewage  treatment 
or  point  sources  was  isotopically  heavy 
((515N  =~12%o).  Showers  et  al  (1990) 
were  able  to  distinguish  a  source  of  nitrate, 
enriched  in  15N  due  to  intense  agricultural 
activity,  coming  from  nonpoint  soil  runoff 
(<515N  =7%o). 

The  study  site  for  this  investigation 
was  Bogue  Sound  and  coastal  waters  di- 
rectly off  Morehead  City  and  Beaufort, 
North  Carolina.  In  this  region  N  sources 
from  sewage  treatment  are  almost  nonex- 
istent. Therefore,  remineralization  and  ag- 
ricultural runoff  provide  the  primary  sources 
of  nitrogen  to  the  phytoplankton  in  this 
ecosystem.  We  also  expected  that  the  <515N 
particulate  nitrogen  would  be  influenced 
by  recycled  nitrogen,  as  15N-enriched  am- 
monium (Cifuentes  et  al. ,  1 989),  and  a 1 5N- 
enriched  nitrate  from  agricultural  runoff, 
providing  these  were  the  only  two  sources 
of  nitrogen  for  primary  production. 

Nitrogen  in  wet  or  dry  deposition  is 
considerably  more  depleted  in  15N  relative 
to  recycled  or  agricultural  inputs  (Heaton, 
1986).  Variability  in  the  isotopic  composi- 
tion of  the  nitrogen  pools  in  acid  deposition 
may  be  indicative  of  certain  atmospheric  N 
sources.  For  example,  in  industrial  zones  of 

Europe,  the  <515N  of  dissolved  ammonium 
has  a  mean  value  of  -12  %c,  whereas  an 
average  value  for  dissolved  nitrate  is  -3  %c 
(Freyer,  1979).  Atmospheric  nitrate  as 
NOx  from  industrial  and  automobile  pollu- 
tion should  have  a  615N  near  that  of  air  (0 


%c).  If,  however,  the  source  of  NOx  is  from 
nitrification  in  soils,  then  owing  to  large 
isotope  fractionation  (Mariotti  et  al.  ,1981), 

the  8  N  of  dissolved  nitrate  may  be  more 
negative. 

The  concentration  and  isotopic  compo- 
sition of  N  in  certain  rainfall  events  was 
determined.  Collections  of  atmospheric 
wet  and  dry  deposition  were  made  on  the 
roof  of  the  Institute  of  Marine  Sciences,  a 
location  free  of  potential  sources  of  con- 
tamination (e.g.,  trees,  powerlines,  other 
buildings).  Large  polypropylene  pans  hav- 
ing splashproof  walls  were  carefully  acid- 
washed  (0.01  N  HC1),  then  rinsed  three 
times  with  deionized  water  in  order  to 
remove  any  traces  of  nutrients.  Pans  were 
placed  in  elevated  stands  on  the  IMS  roof- 
top, prior  to  precipitation  events,  to  collect 
rainwater.  Collectors  were  deployed  just 
prior  to,  and  removed  immediately  after, 
precipitation  events  in  order  to  minimize 
contamination. 

Concentrations  of  dissolved  inorganic 
nitrogen  species  were  first  analyzed  at  the 
University  of  North  Carolina  on  a  subsample 
using  the  methods  of  Strickland  and  Par- 
sons (1972).  Rainwater  samples  were 
shipped  frozen  to  the  Geophysical  Labora- 
tory and  thawed  just  before  isotope  analy- 
sis. Initial  trials  with  rainwater  indicate 
that  molecular  sieve  Zeolite  W-85  adsorbed 
the  ammonium  from  the  rainwater  directly 
without  distillation.  The  <515N  of  solutions 
containing  a  known  NH  +  was  within  the 
standard  error  of  the  measurement  (±  0.5 
%o)  (Velinsky  et  al,  1989).  After  NH4+ 
removal,  some  aliquots  were  freeze-dried. 
The  residual  material,  containing  the  dis- 
solved nitrate  and  any  dissolved  organic 


GEOPHYSICAL  LABORATORY 


149 


Atlantic  Ocean 


Cape 

ookogt 


7 


65  km 


Offshore  Sampling  Site 


Fig.  83.  Map  of  coastal  North  Carolina  showing  three  sampling  locations:  Bogue  Sound,  Nearshore,  and 
Offshore  sites. 


150 


CARNEGIE  INSTITUTION 


matter,  was  combusted  for  isotopic  analy- 
sis. 

In  bioassays  and  natural  waters,  the  N 
isotopic  ratio  of  the  whole  phytoplankton 
sample  was  measured.  Samples  were  col- 
lected from  three  locations:  a  coastal 
nearshore  site,  Bogue  Sound  (near  Beau- 
fort Inlet),  and  90  km  offshore  (Fig.  83). 
These  sites  represent  N-depleted,  full-sa- 
linity Atlantic  coastal  and  mesohaline  es- 
tuarine  waters.  Hydrologically,  Bogue 
Sound  is  a  portion  of  the  meso-  to  euhaline 
component  of  the  Newport  River  Estuary. 
During  incoming  tides,  however,  Bogue 
Sound  is  a  conduit  for  nearshore-coastal 
Atlantic  Ocean  water.  Chronic  N  limita- 
tion characterizes  both  estuarine  and  coastal 
waters  transitting  Bogue  Sound  (Thayer, 
1978;  Paerl,  1988).  Water  samples  were 
filtered  first  through  Nitex  screening  to 
remove  zooplankton.  Following  initial 
screening,  remaining  phytoplankton  were 
collected  on  Whatman  GF/F  filters  (0.7  m 


Q 

Z 
in 


b- 

j_ 

NH4 

NOx  +  DON 

■ 

A 

o- 

V   A 

-5- 

10- 

15  i 
( 

i 
)               60 

i 
120 

■ 

180 

■       i       •       i 

240            30( 

■         i 
)            360 

Julian  Days 

Fig.  84.  Nitrogen  isotopic  composition  of  NH4+ 
and  NOx  plus  dissolved  organic  N  (DON)  in 
continental  rain  events  occurring  on  the  North 
Carolina  coast  over  a  year's  time  from  1  January 
to  31  December  (Julian  Days).  Samples  were 
collected  from  April-May  1988  (n=2)  and  August 
1990- April  1991  (n  =  6).  Variations  in  <515N  most 
likely  reflect  differences  in  the  sources  of  com- 
bined N  to  atmosphere  (e.g.,  agricultural  input  vs. 
industrial  input). 


nominal  pore  size).  Particulate  organic 
matter  on  glass  fiber  filters  was  ground 
with  CuO  and  placed  in  preheated  quartz 
tubes  with  copper  metal.  The  quartz  tubes 
were  evacuated  and  sealed  off,  combusted 
batchwise  at  910° C  for  2  h,  and  cooled  at  a 
controlled  rate.  Replicate  analysis  of  fil- 
ters plus  organic  material  gave  values  with 
standard  deviations  of  ±  0.5  %o. 

The  <515N  of  the  NH.+ from  continental 

4 

rainfalls  varied  throughout  the  year  (Fig. 
84),  but  had  values  considerably  more 
negative  than  either  recycled  or  agricul- 
tural inputs.  When  nitrogen  is  incorporated 
into  phytoplankton  during  primary  pro- 
duction, the  isotopic  compositions  of  algae 
will  depend  on  (1)  any  biochemical  frac- 
tionations that  may  occur  and  (2)  the  iso- 
tope ratios  of  the  available  nitrogen  sources 
(Cifuentes  et  al,  1989).  Accordingly,  the 
<515N  of  primary  producers  should  shift 
with  the  addition  of  nitrogen  from  AD.  In 
fact,  when  significant  rainfall  events  oc- 
curred, the  <515N  of  particulate  nitrogen 
decreased  within  a  few  days  after  the  event 
(Fig.  85). 


20 


•z. 


15- 


10- 


5- 


A     U 


:i 


t  * 


r 


D    Offshore 

/k    Bogue  Sound 

■    Nearshore 


0        30       60       90       120     150      180     210     240     270 
Fall  Winter  Spring 

Time  (Days) 

Fig.  85.  Nitrogen  isotopic  composition  of  particu- 
late material  sampled  from  three  coastal  and  estua- 
rine sites  off  North  Carolina  as  a  function  of  time. 
Day  1  =  1  August  1990.  Arrows  indicate  the 
timing  of  a  continental  rainfall  event. 


GEOPHYSICAL  LABORATORY 


151 


^j" 

bU  - 
50- 
40- 
30- 
20- 
10- 
0- 

c 

A 

• 

Mixing  depth  =  1 .5  m 

z 

Q 

|2 

• 

• 

• 

\ 

[PN]  ( ji  gN/L) 

V  •  Am       •*  • 

)          30 

i 
60 

90 

120 

150       180      210      240 

Time  (Days) 


50 


40 


30- 


Q      20 

a 

o 

•"      10 


B 


Mixing  Depth=  2  m 


[PN]  (W  N/L) 


-i ■ r 


V  I  tSafL 


30  60  90  120         150         180        210        240 

Time  (Days) 


Mixing  Depths 
■  0.5m 
•  5m 


a 

o 


90        120      150       180      210      240 
Time  (Days) 

mm  Continental     V~Z\    Mixed     K/sA   Oceanic 

Fig.  86.  Total  dissolved  inorganic  nitrogen  (DIN) 
from  atmospheric  deposition  that  is  mixed  into 
surface  layers  of  the  North  Carolina  Coast.  All 
rainfall  events  occurred  from  1  August  1990  to 
April  1991  (Table  22).  (A)  Bogue  Sound  site.  (B) 
Nearshore  site.  (C)  Offshore  site.  Mixing  depths 
are  approximations  estimated  from  the  depth  of 
the  water  column  and  the  photic  zone.  The  line 
indicates  average  ambient  particulate  N  concen- 
trations measured  periodically  throughout  the  time 
period.  Levels  of  DIN  were  generally  <  1  |imole 
N/L. 


Table  22.   Dissolved  inorganic  N  (DIN)  content  of  selected  significant  (more  than  0.5  cm) 
rainfall  events  at  the  UNC  Institute  of  Marine  Sciences,  Morehead  City,  North  Carolina.* 


Date 

pH 

Origin                    [NOx] 

[NH/] 

Amount(cm) 

9  Aug  90 

3.58 

Continental 

659 

1115.5 

4.23 

10  Sep  90 

3.41 

Continental 

2213 

1416.4 

0.51 

24-26  Oct  90 

4.60 

Mixed 

295 

103.4 

3.98 

6  Nov  90 

5.74 

Mixed 

594.5 

128 

0.53 

30  Nov  90 

4.75 

Mostly  Oceanic 

216.5 

48.7 

1.85 

4  Dec  90 

5.40 

Oceanic 

40.2 

77.9 

1.43 

9  Dec  90 

4.65 

Mostly  Oceanic 

130.6 

44 

2.49 

21  Dec  90 

4.45 

Mixed 

280.4 

74.5 

1.19 

22  Dec  90 

4.81 

Oceanic 

60.1 

9.13 

0.77 

3-4  Jan  91 

4.17 

Mostly  Continental 

434.4 

84.8 

0.44 

9  Jan  91 

4.79 

Mostly  Oceanic 

75.2 

46.8 

4.66 

12  Jan  91 

4.79 

Mixed 

224 

122 

1.94 

16  Jan  91 

5.10 

Oceanic 

83.5 

27 

2.84 

20  Jan  91 

5.16 

Oceanic 

30.2 

43.9 

3.78 

25  Jan  91 

4.74 

Mixed 

118.8 

28.9 

2.27 

8  Feb  91 

4.89 

Oceanic 

44.6 

ND 

1.72 

5  Mar  91 

5.13 

Mostly  Oceanic 

87.2 

76.2 

5.30 

14  Mar  91 

4.36 

Mostly  Continental 

376.7 

275.6 

1.56 

30  Mar  91 

5.07 

Mostly  Oceanic 

159.9 

187.3 

2.24 

*  Weather  fronts  were  documented  from  satellite  imagery  and  dominant  wind  direction. 
DIN  concentrations  are  in  Lig  N/  L.  ND  indicates  non-detectable  concentrations. 


152 


CARNEGIE  INSTITUTION 


Table  23.  Bioassay  experiments  to  assay  incorporation  of  atmospheric  deposition  into  particulate 
nitrogen  from  coastal  samples. 

Type  of  Date         Final  Rain    Initial    Final  5^N    Rainwater        Origin  of         Date  of 

Bioassay  dilution    S^N  PN       PN         S^N  NH44"       last  rain         last  rain 


Cubitainer 
Cubitainer 
Mesocosm 


13-Sep-90 
17-Oct-90 
27-Apr-91 


1%  +4.2 

5%         +10.8 
1.80%    +16.3 


+4.4  ND*  Continental  10-Sep-90 

+3.5  -9.5  Mixed  22-Sep-90 

+12.2  -1.5  Mixed  22-Apr-91 

+15.9  ND  Oceanic  27-Apr-91 


Initial  particulate  nitrogen  samples  were  collected  from  Bogue  Sound,  not  determined. 


Rainfall  events  in  the  Beaufort,  North 
Carolina,  area  originate  from  either  conti- 
nental or  oceanic  sources,  or  are  frequently 
a  mixture  of  the  two  (Table  22). 

The  amount  of  dissolved  inorganic  ni- 
trogen (DIN)  and  pH  can  usually  be  related 
to  the  origin  of  the  rainfall  event.  Continen- 
tal storms  had  an  average  DIN  concentra- 
tion of  1643  jig  N/L  (range  =  519-3629 
jigN/L),  whereas  oceanic  events  contained 
as  little  as  150  jug  N/L  (range  =  69-347 
jigN/L).  To  assess  the  relative  importance 
of  AD  nitrogen  to  the  ecosystem,  the  flux  of 
N  falling  on  the  three  coastal  sites  has  been 
calculated  for  each  rainfall  event.  Depend- 
ing on  the  mixing  depth  at  each  location 
and  the  amount  of  rainfall,  rain  from  all 
three  origins  contributed  significant 
amounts  of  nitrogen,  especially  at  the  off- 
shore site  (Fig.  86). 

To  determine  direct  uptake  of  nitrogen 
from  AD,  short  and  long-term  bioassays 
were  designed  to  examine  impacts  of  AD 
on  isotopic  composition  and  primary  pro- 
duction under  natural  irradiance  and  tem- 
perature conditions.  Two  independent  bio- 
assays were  employed.  The  first  used  rela- 
tively small  volume  Cubitainers  (4L)  incu- 
bated in  situ  for  delineating  short-term  (1 
day-1  week)  impacts  (Paerl  et  a/.,  1990). 
The  second  used  previously  constructed 


mesocosms  (670  L),  designed  to  evaluate 
longer-term  chronic  loading  effects  and  to 
approximate  the  estuarine  environments 
surrounding  the  Beaufort-Morehead  City 
area.  When  a  rainfall  event  was  antici- 
pated, a  control  set  of  mesocosms  was 
covered  with  transparent  polyethylene, 
thereby  excluding  rain  water.  A  second  set 
of  mesocosms  remained  uncovered,  thereby 
allowing  rain  input.  To  a  third  set  of 
mesocosms,  rainfall  from  a  previous  con- 
tinental event  was  added.  Added  rainfall 
simulated  dilutions  commonly  experienced 
in  coastal  waters  (Table  23). 

An  aliquot  of  14C-NaHC03  was 
added  to  each  vessel  in  order  to  monitor 
photosynthetic  14C02  assimilation  as  a 
measure  of  microalgal  production.  A  par- 
allel 14C-free  set  of  vessels  were  deployed 
for  stable  isotope  analyses.  Initial  samples 
for  particulate  <515N  analyses  were  taken  at 
this  time.  After  4  days,  samples  were  col- 
lected and  filtered  for  14C02  assimilation 
and  particulate  <515N. 

The  first  Cubitainer  experiment  fol- 
lowed directly  after  a  continental  rainfall 
that  contained  a  significant  amount  of  DIN 
(Table  23).  Although  there  was  some  stimu- 
lation of  primary  production  as  measured 

with  14C  uptake,  no  isotopic  changes  re- 
sulted. In  the  second  experiment,  the  added 
rainfall  was  from  9  Aug  1990  water,  which 


GEOPHYSICAL  LABORATORY 


153 


had  a  515N-NH,+  of  -9.5  %o.  Prior  to  this 

4 

experiment,  no  significant  rainfall  had  oc- 
curred in  the  area  for  25  days.  The  addition 
of  5  %  rainwater  to  the  Cubitainer  caused  a 
7  %o  decrease  in  the  <515N  of  the  particulate 
N,  indicative  of  the  uptake  of  the  isotopi- 
cally-light  N  from  the  acid  rain. 

The  mesocosm  experiment  was  con- 
ducted following  a  period  when  rain  fall 
events  had  been  primarily  of  oceanic  or 
mixed  origin  (see  Table  22).  Rain  from  a 
mixed  event  (22  Apr  91)  was  added  to  a  1 .8 
%  dilution  to  certain  mesocosms,  while 
others  received  oceanic  rainfall  that  oc- 
curred a  day  later  (Table  23).  Primary 
production  was  stimulated  with  the  addi- 
tion of  the  22  April  rainwater,  and  5  N  of 
particulate  N  decreased  by  4  %o.  In  con- 
trast, no  stimulation  of  primary  production 
as  determined  by  14C  uptake  was  mea- 
sured, relative  to  controls,  with  the  oceanic 
event,  and  no  change  in  the  5  N  of  particu- 
late N  was  detected. 

Previous  results  suggest  that  atmo- 
spheric N  inputs,  which  may  supply  as 
much  as  20  to  50  %  of  coastal  ocean  "new" 
nitrogen  inputs,  represent  a  significant 
source  of  N,  that  contribute  to  current  rates 
of  eutrophication.  By  demonstrating  the 
actual  incorporation  of  nitrogen  from  AD 
into  primary  producers  with  N  isotope  trac- 
ers, we  conclude  that  the  biological  mea- 
surements of  enhanced  primary  productiv- 
ity are  due  to  the  presence  of  this  alterna- 
tive source  of  a  limiting  nutrient.  Atpresent 
we  do  not  know  how  primary  productivity 
alterations  in  response  to  acid  rain  might 
shape  short-  or  long-term  production  trends 
of  estuarine  and  coastal  ocean  food  chains. 
This  problem  is  by  nature  a  global  one.  The 


oxides  of  nitrogen  responsible  are  largely 
generated  in  major  urban  and  industrial 
centers  located  to  the  north  and  west  of 
North  Carolina,  although  their  exact  origin 
can  only  be  speculated  upon  on  the  basis  of 
concentration  and  location.  Fertilizers  and 
animal  wastes  associated  with  farming  may 
also  be  important  sources  of  NH  +  in  this 
coastal  area.  Isotope  tracers  of  atmospheric 
deposition  may  not  only  be  used  in  tracing 
N  into  the  food  chain,  but  also  may  have 
some  relevance  to  the  origin  of  the  atmo- 
spheric nitrogen. 


References 

Boynton,  W.  R.,  W.  M.  Kemp,  and  C.  W.  Keefe, 
A  comparative  analysis  of  nutrients  and  other 
factors  influencing  estuarine  phytoplankton  pro- 
duction, in  Estuarine  Comparisons,  V.  S. 
Kennedy,  ed.,  pp.  69-90,  Academic  Press,  NY, 
1982. 

Cifuentes,  L.  A.,  M.  L.  Fogel,  J.  R.  Pennock,  and 
J.  H.  Sharp,  Biogeochemical  factors  that  influ- 
ence the  stable  nitrogen  isotopic  ratio  of  dis- 
solved ammonium  in  the  Delaware  estuary, 
Geochim.  Cosmochim.  Acta,  53,  2713-2721, 
1989. 

Cosper,  E.  M,  W.  C.  Dennison,  E.  J.  Carpenter,  V. 
M.  Bncelj,  J.  G.  Mitchell,  S.  H.  Kuenstner,  D. 
Colflesh,  and  M.  Devey,  Recurrent  and  persis- 
tent brown  tide  blooms  perturb  coastal  marine 
ecosystem,  Estuaries,  10,  284-290,.  1987. 

Fisher,  D.,  J.  Ceraso,  T.  Mathew,  and  M. 
Oppenheimer,  Polluted  Coastal  Waters:  The 
Role  of  Acid  Rain.  Environmental  Defense 
Fund,  New  York,  1988. 

Freyer,  H.  D,  Seasonal  trends  of  NH4+  and  N03' 
nitrogen  isotope  composition  in  rain  collected 
at  Julich,  Germany,  Tellus,  30,  83-92,  1979. 

Heaton,  T.H.E.,  Isotopic  studies  of  nitrogen  pollu- 
tion in  the  hydrosphere  and  atmosphere:  A 
review,  Chem.  Geol,  59,  87-102,  1986. 

Legendre,  L.  O.  and  Gosselin,  M.  New  production 
and  export  of  organic  matter  to  the  deep  ocean: 
consequences  of  some  recent  discoveries, 
Limnol.  Oceanogr.,  34,  1374-1380,  1989. 

Loye-PiJot,  M.  D.,  J.M.  Martin,  and  J.  Morelli, 
Atmospheric  input  of  inorganic  nitrogen  to  the 
western  Mediterranean,  Biogeochem.,  9,  1 17- 
134,  1990. 


154 


CARNEGIE  INSTITUTION 


Mariotti,  A.,  J.  C.  Germon,  P.  Hubert,  P.  Kaiser,  R. 
Letolle,  A.  Tardieux,  and  P.  Tardieux,  Experi- 
mental determination  of  nitrogen  kinetic  iso- 
tope fractionation,  some  principles;  illustration 
for  the  denitrification  and  nitrification  prin- 
ciples, Plant  and  Soil,  62,  413-430,  1981. 

Nixon,  S.  W.,  C.  A.  Oviatt,  J.  Frithsen,  and  B. 
Sullivan,  Nutrients  and  the  productivity  of  es- 
tuarine  and  coastal  marine  ecosystems.  /. 
Limnol.  Soc.  So.  Afr.,  12,  43-71,  1986. 

Owens,  N.  J.  P.,  Natural  variations  in  15N  in  the 
marine  environment,  Adv.  in  Mar.  Biol.,  24, 
411-451,1987. 

Paerl,  H.  W.,  Enhancement  of  marine  primary 
production  by  nitrogen-enriched  acid  rain,  Na- 
ture, 316,  747-749,  1985. 

Paerl,  H.  W.,  Nuisance  phytoplankton  blooms  in 
coastal,  estuarine,  and  inland  water,  Limnol. 
Oceanogr.,  33,  823-847,  1988. 

Paerl,  H.  W.,  J.  Rudek,  and  M.  A.  Mallin,  Stimula- 
tion of  phytoplankton  production  in  coastal 
waters  by  natural  rainfall  inputs:  Nutritional 
and  trophic  implications,  Mar.  Biol,.  107, 247- 
254,  1990. 

Prado-Fiedler,  R.,  Atmospheric  input  of  inorganic 
nitrogen  species  to  the  Kiel  Bight,  Helgolander 
Meersuut,  44,  21-30,  1990. 

Ryther,  J.  H.,  and  W.  M.  Dunstan,  Nitrogen, 
phosphorus,  and  eutrophication  in  the  coastal 
marine  environment,  Science,  171, 1008-1013, 
1971. 

Showers,  W.  J.,  D.  M.  Eisenstein,  H.  W.  Paerl,  and 
J.  Rudek,  Stable  isotope  tracers  of  nitrogen 
sources  to  the  Neuse  River,  North  Carolina, 
Water  Resources  Institute  Report  No.253, 1990. 

Strickland,  J.  D.  H.  and  T.  R.  Parsons,  A  Practical 
Handbook  of  Seawater  Analysis.  Fish.  Res. 
Board  Can.  Bull.  167,  1972. 

Thayer,  G.  W.,  Identity  and  regulation  of  nutrients 
limiting  phytoplankton  production  in  the  shal- 
low estuaries  near  Beaufort,  N.C.,  Oecologia, 
14,  75-92,  1978. 

Velinsky,  D.  J.,  L.  A.  Cifuentes,  J.  R.  Pennock,  J. 
H.  Sharp,  and  M.  L.  Fogel,  Determination  of 
the  isotopic  composition  of  ammonium-nitro- 
gen at  the  natural  abundance  level  from  estua- 
rine waters,  Marine  Chemistry,  26,  351-361, 
1989. 


Nitrogen  Diagenesis  in  Anoxic  Marine 
Sediments:  Isotope  Effects 

David  J.  Velinsky,  David  J.  Burdige*  and 
Marilyn  L.  Fogel 

Due  to  the  complexity  of  particulate 
nitrogen  (PN)  transformations,  the  stable 
isotopic  composition  of  sedimentary  nitro- 
gen has  received  little  attention  as  an  indi- 
cator of  changes  in  the  nitrogen  biogeo- 
chemistry  of  the  oceans.  During  diagenesis 
nitrogen  can  change  oxidation  state  (-3  to 
+5)  through  a  series  of  bacterially  -  medi- 
ated reactions  (Klump  and  Martens,  1983). 
These  transformations  have  a  range  of  iso- 
tope effects,  related  mainly  to  kinetic  dif- 
ferences in  the  reactivity  between  the  light 
(14N)  and  heavy  (15N)  isotopes  (Wada  et  al. , 
1 975).  As  a  result,  diagenetic  processes  can 
affect  the  overall  isotopic  composition  of 
dissolved  and  particulate  nitrogen  within 
the  sediments  and  overlying  waters.  For 
example,  if  the  extent  of  denitrification 
(N03"  — >  N2)  in  the  water  column  and 
continental  shelf  sediments  changed  over 
time,  the  #5N  of  oceanic  nitrogen  could 
shift  significantly  due  to  the  large  isotope 
effect  associated  with  denitrification  (e  = 
30-40  %c>;  Wada  et  a/.,  1975;  Cline  and 
Kaplan,  1975)  and  the  extent  of  this  process 
as  a  sink  for  combined  nitrogen  in  the 
oceans  (Christensen,  1987).  The  source 
and  isotopic  composition  of  nitrogen  to  the 
sediments  can  change  depending  on  con- 
ditions in  surface  waters.  Rau  et  al.  (1987) 


*  Department  of  Oceanography,  Old  Dominon 
University,  Norfolk,  Virginia 


GEOPHYSICAL  LABORATORY 


155 


showed  that  the  #5N  of  PN  in  organic 
carbon-rich  Cretaceous  marine  sediments 
were  isotopically  light  (-4.0  to  +1.0  %6) 
compared  to  typical  marine  sediments  (gen- 
erally >4  %6).  They  speculated  that  N2  fixa- 
tion was  the  dominant  source  of  organic 
nitrogen  to  these  Cretaceous  sediments. 

The  type  and  isotopic  composition  of 
particulate  nitrogen  that  is  incorporated 
into  the  sediments  depends  on  source  varia- 
tions and  isotope  effects  during  diagenesis. 
For  source  variations  to  be  evaluated,  it  is 
important  to  determine  if  any  post-deposi- 
tional  isotope  effect  occurs  during  early 
diagenesis.  The  purpose  of  this  paper  is  to 
investigate  possible  isotope  effects  during 
diagenesis  of  PN  in  anoxic  marine  sedi- 
ments. Once  diagenetic  effects  are  deter- 
mined, a  more  accurate  interpretation  of 
nitrogen  isotopes  in  ancient  and  present^ 
day  marine  sediments  can  be  obtained. 


Framvaren  Fjord 


Great  Marsh 


▲       A 

Leaves,  Framvaren  Fjord 


-30 


-25 


-20 


-15 


513C 


Fig.  87.  The  <513C  of  organic  carbon  and  <515N  of 
particulate  nitrogen  from  sediments  taken  from 
the  Chesapeake  Bay,  Framvaren  Fjord  and  Great 
Marsh.  Also  included  are  isotope  values  from 
various  plants  around  the  Framvaren  Fjord. 


Methods  and  Study  Areas 

Concentrations  and  <515N  of  pore  water 
NH4+  and  sedimentary  PN  were  deter- 
mined from  cores  taken  in  three  contrasting 
coastal  marine  environments:  Framvaren 
Fjord,  Norway  (FF;  June  1989),  Great 
Marsh,  Delaware  (GM;  June,  1988),  and 
Chesapeake  Bay  (CB;  June,  1988). 

The  Framvaren  Fjord  is  a  permanently 
anoxic  fjord  with  concentrations  of  dis- 
solved NH4+  and  hydrogen  sulfide  (H2S) 
in  the  bottom  waters  (maximum  depth  180 
m)  of  2  mM  NH4+  and  6  mM  H2S,  respec- 
tively. FF  sediments  are  permanently 
anoxic.  The  sources  of  organic  matter  to  the 
sediments  are  primarily  from  bacterial  and 
phytoplankton  production  in  the  water  col- 
umn along  with  some  terrestrial  inputs  (e.g., 
leaves). 

The  Great  Marsh  is  a  coastal  salt  marsh 
dominated  by  the  short  form  of  Spartina 
alterniflora.  Marsh  sediments  in  the  upper 
12  cm  (i.e.,  the  active  root  zone)  cycle 
seasonally  betwenn  oxidized  and  reducing 
conditions  (Velinsky  and  Cutter,  1991). 
Sediments  below  -12  cm  are  permanently 
anoxic.  Sources  of  organic  matter  to  these 
sediments  include  Spartina  production 
and  upland  runoff. 

Two  sediment  cores  were  obtained  from 
the  Chesapeake  Bay,  near  38°56'  N, 
76°23'W,  just  south  of  the  Chesapeake  Bay 
Bridge  near  Annapolis,  Maryland.  The 
water  column  in  this  section  of  the  bay 
(maximum  depth  30  m)  is  seasonally  anoxic 
or  sub-oxic  whereas  the  sediments  are  gen- 
erally permanently  anoxic  (San  Diego- 
McGlone,  1991).  Sources  of  organic  mat- 
ter to  the  sediments  of  this  portion  of  the 


156 


CARNEGIE  INSTITUTION 


0 
0 


10    - 


20    - 


E 


30    - 


40    - 


50 


%PN  and  NH+  (mM) 
1.0  2.0 


515N 


3.0 


0 


8 


10 


1 1 1 1 1 1 1— 

—1 • 1 1 

C/N  , 

- 

I        \ 

\  nh;  ■ 

PN    A 

\ 

r 

i             1 

_i i 

10  15 

C/N  (atomic) 


20 


Fig.  88.  The  depth  distributions  of  particulate  nitrogen  (PN),  dissolved  ammonium  (NH4+)  and  the 
organic  carbon  to  particulate  nitrogen  ratio  (C/N)  along  with  the  isotopic  composition  (515N)  of 
dissolved  NH4+  and  PN  in  Framvaren  Fjord  sediments. 


bay  are  primarily  derived  from  phy toplank- 
ton  production  and  to  a  lesser  extent  from 
land  runoff. 

Box  (CB)  or  gravity  (FF  and  GM)  cores 
were  sectioned  at  specific  intervals.  The 
CB  core  was  sectioned  every  2  cm,  and  the 
FF  core  was  sectioned  at  5-cm  intervals. 
The  GM  core  was  sectioned  every  2.5  cm, 
and  due  to  the  small  volume  of  pore  fluids 
obtained  and  the  low  concentrations  of 
dissolved  NH4+ in  the  upper  15  cm,  the  2.5- 
7.5  cm  and  the  10-15  cm  sections  were 


combined  for  isotope  analysis.  Pore  fluids 
were  separated  from  sediments  by  either 
centrif  ugation  (FF)  or  with  Reeburgh  ( 1 967 ) 
sediment  squeezers  (CB  and  GM),  then 
filtered  through  Nuclepore  0.4  jiim  filters. 
Both  sediment  and  pore  waters  were  stored 
frozen  until  sample  preparation  and  analy- 
sis. 

The  methods  for  the  preparation  and 
determination  of  the  <515N-NH4+and  <515N- 
PN  are  described  in  Velinsky  et  al.  (1989) 
and  Cifuentes  et  al.  (1988).  The  data  are 


GEOPHYSICAL  LABORATORY 

Table  24.  Isotope  discrimination  between  PN  and  NH4+. 


157 


Location 

S15  PN* 

Reference 

Frarnvaren  Fjord 

3.3±0.9 

-0.210.8 

This  Study 

Chesapeake  Bay 

Core  41 
Core  42 

9.4±0.5 
9.8±1.1 

-  2.8  to  -  0.7 
2.3  to  -  0.4 

This  Study 
This  Study 

GreatMarsh,DE. 

5.2±0.8 

-  0.810.6 

This  Study 

Santa  Barabara  Basin 

SOG  005 
GAS  24 
GAS25 
GAS  29 

7.111.2 
7.010.9 
7.310.3 
6.110.9 

+  2.611.5 
+  3.211.0 
+  4.110.6 
+  5.011.7 

Sweeney  and  Kaplan  (1980) 

*  Average  ^l^PN 

**Average  A  =  #15NH4+  -  515PN 


%PN  and  NH  +  (mM) 
0  1.0  2.0  3.0 


u 

* 

>          p       [ 

— I ' 1 r" 

C/N    , 

10 

>v                         I        / 

- 

E 

20 

\   NH4     " 

CD 

Q 

30 

1 

PN    6 

►                > 

40 

)    ' 

[ 

3                 • 

50 

i 

« 

10  15  20 

C/N  (atomic) 

Fig.  89.  Similar  to  Fig.  88,  except  for  the  Great  Marsh. 


158 


CARNEGIE  INSTITUTION 


reported  in  the  standard  S  notation  {i.e., 
<515N  =  [(#sample/#standard)  -  1]  x  103;  where 
R  =  15n/14N}  and  the  ratios  are  reported 
against  air  (<515N  =  0).  Precision  of  repli- 
cate samples  for  ammonium  and  PN  isoto- 
pic  analysis  is  approximately  ±  0.5  %o  and 
±  0.2  %o,  respectively.  Ammonium  concen- 
trations were  determined  with  the  proce- 
dures described  in  Sharp  et  al.  (1982)  and 
solid  phase  PN  concentrations  were  deter- 
mined with  a  Carlo-Erba  ANA  1500  car- 
bon and  nitrogen  analyzer. 


Results  and  Discussion 

To  illustrate  the  differences  between 
sedimentary  environments,  the  isotopic 
compositions  of  organic  carbon  (<513C)  and 
particulate  nitrogen  (<515N)  are  plotted  (Fig. 
87).  Because  of  the  predominance  of 
Spartina  aiterniflora  (a  C4  plant)  in  the 
GM,  the  <513C  of  the  sediments  are  greater 
than  those  in  either  the  FF  and  CB  sedi- 
ments. The  carbon  and  nitrogen  composi- 
tion of  FF  sediments  reflect  terrestrial  and 


%PN 
0.3  0.4 


0.5 


15.0 


NH4  (mM) 


Fig.  90a.  Similar  to  Fig.  88,  except  for  Chesapeake  Bay,  Core  41. 


GEOPHYSICAL  LABORATORY 


159 


E 
o 


Q. 
CD 

Q 


5.0 


7.5 


815N 

10.0 


12.5 


15.0 


NH^(mM) 


Fig.  90b.  Similar  to  Fig.  90a,  except  for  Chesapeake  Bay,  Core  42. 


water  column-derived  inputs,  whereas  the 
CB  sediments  are  isotopically  enriched  in 
15N  and  13C,  indicating  typical  estuarine 
phytoplankton. 

In  the  FF  sediments,  the  <515N  of  both 
pore  water  NH4+  and  sedimentary  PN  did 
not  change  significantly  with  depth  (Fig. 
88).  As  such,  the  isotopic  difference  (A  = 
5!5N-NH4+  -  <515N-PN)  between  the  NH4+ 
and  PN  is  small  (~  -0.2%o)  with  the  NH4+ 
being  15n  depleted  (Table  24).  These  data 
indicate  no  expression  of  significant  iso- 
tope effect  during  ammonification  of  PN  in 
the  FF  sediments. 


The  concentrations  of  PN  decreased 
significantly  (60%)  in  the  GM  sediments 
(Fig.  89)  from  the  surface  to  approximately 
27  cm.  Concurrently,  the  <515N-PN  in- 
creased very  slightly  in  this  interval  with  an 
overall  average  of  5.66  ±  0.46  %o  (n  =  8). 
Whereas  pore  water  NH4+  concentrations 
increased  with  depth,  the  <515N-NH4+  did 
not  change  and  averaged  4.30  ±  0.41  %o 
(n=l).  Similar  to  FF  sediments,  the  <515N- 
NH4+  was  lighter  than  the  <515N-PN,  indi- 
cating a  small  negative  discrimination  with 
depth  (i.e.,  14N-NH4+  is  preferentially  re- 
leased compared  to  15N-NH4+;  Table  24). 


160 


CARNEGIE  INSTITUTION 


This  discrimination  is  a  result  of  a  number 
of  processes  that  could  fractionate  NH4+, 
including  ion-exchange,  deamination  of 
amino  acids  (Macko  and  Estep,  1984),  up- 
take by  bacteria  within  the  sediments  and 
diffusion  out  of  the  sediments.  The  ob- 
served distribution  of  <515N-NH4+  with 
depth  most  likely  results  from  a  combina- 
tion of  these  processes. 

The  Chesapeake  Bay  cores  exhibited  a 
slightly  different  trend  in  <515N-NH4+  with 
depth  (Fig.  90a).  In  Core  41 ,  the  concentra- 
tion of  PN  decreased  with  depth  in  the 


upper  10  cm  and  remains  constant  below 
10  cm,  while  the  <515N-PN  did  not  change 
in  the  upper  10  cm.  Conversely,  as  the 
concentration  of  NH44"  increased  with 
depth,  the  <515N-NH4+  also  increased.  The 
maximum  increase  in  <515N-NH4+  was  in 
the  same  depth  intervals  as  the  decrease  in 
PN  and  increase  in  dissolved  NH4+.  The 
differences  between  Core  41  and  42  (Fig. 
90b)  reflect  the  spatial  heterogeneity  of 
sediments  in  the  bay.  A  sharp  maximum  in 
pore  water  NH4+  and  <515N  occurs  in  the  4- 
6  cm  interval  and  this  was  related  to  an 


PN 


PN 


R 


Nitrogen  Cycle  in  Sediments 


NH 


NO, 


Water 


uuj 


2UUUUU 


i«p 


PN, 


PN 


R 


Burial 


K-2.B 


Sediment 


4 

no3 

K 

N 
2 

k, 

ki 

flux  of  PN  to  the  sediments 

k4 

denitrification 

k2 

am  monification 

k5 

nitrate  diffusion 

k-2 

microbial  uptake 

k6 

ammonium  diffusion 

k3 

nitrification 

k7 

burial 

Fig.  90.  A  conceptual  model  for  the  pathways  of  PN  mineralization  in  marine  sediments.  Note  that  the 
PN  is  broken  into  two  fractions:  a  labile  phase  (PNl)  and  a  refractory  phase  (PNr).  With  depth  and  burial, 
the  PNl  fraction  would  decrease  as  would  the  rate  of  remineralization  (k2;  Burdige,  1991). 


GEOPHYSICAL  LABORATORY 


161 


extensive  shell  layer  found  at  this  depth. 
This  shell  layer  contained  additional  or- 
ganic material  that  enhanced  sulfate  reduc- 
tion causing  the  pore  water  sulfate  concen- 
trations to  decrease  to  undetectable  levels 
at  4-6  cm  versus  12-14  cm  for  Core  41.  As 
a  result,  pore  water  concentrations  of  dis- 
solved NH4+  increased  dramatically  from 
the  surface  (1.3  mM)  to  4-6  cm  (6  mM). 
Below  the  NH4+  maximum,  concentra- 
tions decreased  to  near  constant  levels, 
whereas  the  concentration  of  PN  decreased 
to  a  minimum  at  approximately  1 8  cm  (Fig. 
90b).  The  S^N  of  both  NH4+  and  PN 
reflected  the  change  in  source  of  PN  with 
distinctly  different  isotopic  ratios  of  nitro- 
gen in  the  4-6  cm  interval. 

The  absence  of  any  significant  isotope 
effect  during  PN  remineralization  may  re- 
flect the  variety  of  processes  that  are  affect- 
ing PN  in  sediments.  Each  process  could 
have  an  intrinsic  isotope  effect  that  is  fully 
expressed  in  the  overall  solid  phase  PN 
distribution.  Alternately,  the  lack  of  any 
observable  shift  in  the  <515N-PN  with  depth 
may  be  due  to  a  "mass"  effect,  in  that  only 
a  small  fraction  of  the  PN  is  remineralized 
compared  to  the  total  PN.  As  such  it  would 
be  difficult  to  see  any  discrimination  unless 
it  is  very  large.  Pore  water  NH4+  should  be 
a  better  indicator  of  mineralized  material 
and,  as  such,  should  be  a  more  accurate 
reflection  of  any  diagenetic  isotope  effect 
on  the  PN. 

The  distributions  of  A  are  similar  for  the 
Chesapeake  Bay  cores.  In  both  cores,  A  in 
the  surface  section  was  approximately  -2.5 
%o(i.e.,  the  NH4+  is  isotopically  lighter  than 
the  PN)  and  increased  with  depth.  This 
trend  indicates  a  possible  selective 


remineralization  in  the  upper  sediments  of 
a  more  labile  fraction  of  the  PN  with  a 
lighter  isotopic  composition  (PNl;  Fig. 
91).  This  fraction  of  PN  could  represent 
"fresh"  phytoplankton  material  that  has 
reached  the  bottom  sediments  relatively 
unaltered.  Montoya  et  al.  (1990)  showed 
that  particulate  material  from  the  mainstem 
of  Chesapeake  Bay  during  the  spring  had 
nitrogen  isotopic  compositions  of  between 
6.2  and  10.5  %o.  If  this  scenario  is  the  case, 
the  preferential  degradation  of  this  more 
labile  organic  matter  (Westrich  and  Berner, 
1984;  Burdige,  1991)  would  release  dis- 
solved NH4+  into  the  pore  waters  with 
<515N  similar  to  that  of  the  source  material 
(Fig.  91).  With  depth  less  of  this  lighter 
material  remains  and  eventually  the  <5l5N- 
NH4+  of  the  pore  waters  would  reflect 
remineralization  of  the  "background",  more 
refractory,  particulate  nitrogen  (PNr).  Such 
results  were  shown  by  Sweeney  and  Kaplan 
(1980)  for  sediments  taken  from  the  Santa 
Barbara  Basin  (Table  24).  They  demon- 
strated that  the  <515N  of  the  pore  water 
NH4+  reflected  the  degradation  of  a  marine 
source  of  PN  with  a  <5i5N  of  approximately 
10  %o.  The  range  of  <5i5N  for  total  nitrogen 
in  these  sediments  was  thought  to  be  de- 
rived from  a  mixture  of  a  marine  (i.e., 
phytoplankton  at  10  %o)  and  terrestrial 
sources  (i.e.,  sewage-derived  at  2  %c).  Dis- 
solved NH4+  in  the  pore  waters  was  there- 
fore postulated  to  be  derived  mainly  from 
the  preferential  degradation  of  the  plank- 
tonic  organic  nitrogen.  Although  the 
data  from  Chesapeake  Bay  reflects  the  pos- 
sible degradation  of  a  labile  fraction  of  PN, 
the  data  from  the  Framvaren  Fjord  and 
Great  Marsh  did  not  reflect  this  distribu- 


162 


CARNEGIE  INSTITUTION 


tion  (Figs.  88  and  89;  Table  24).  It  is  pos- 
sible that  in  the  Framvaren  Fjord  the  major- 
ity of  the  degradation  of  PN  occurred  in  the 
water  column  above  the  sediments.  By  the 
time  this  material  reached  the  sediments 
any  labile  phase  was  already  degraded.  In 
the  GM  sediments,  labile  N  could  be  re- 
leased in  the  upper  10-12  cm  and  either 
consumed  by  the  Spartina  or  possibly 
released  via  diffusion  to  the  adjacent  creek 
waters 

In  conclusion,  only  a  small  isotopic 
discrimination  is  expressed  during  diagen- 
esis  (Table  24).  Whereas,  the  bulk  Sl5N- 
PN  did  not  significantly  change  with  depth, 
it  appears  that  selective  remineralization  of 
a  labile  fraction  of  N  may  occur  in  certain 
environments.  This  observation  indicates 
the  <515N  of  specific  fractions  of  the  PN 
could  be  used  as  a  tracer  of  recently  formed 
organic  material.  Downcore  variations  in 
<515N  of  solid-phase  nitrogen  probably  re- 
flect the  material  deposited  to  the  sediment 
surface.  It  is  not  possible  to  tell  however,  if 
the  isotopic  composition  of  the  PN  formed 
in  the  water  column  is  reaching  the  sedi- 
ment intact,  only  that  the  isotopic  integrity 
of  the  bulk  PN  appears  to  remain  unaf- 
fected. 


References 

Burdige,  D.  J.,  The  kinetics  of  organic  matter 
mineralization  in  anoxic  marine  sediments, 
Jour.  Mar.  Res.,  in  press,  1991. 

Christensen,  J.  P.,  J.  W.  Murray,  A.  H.  Devol,  and 
L.  A.  Codispoti,  Denitrification  in  continental 
shelf  sediments  has  major  impact  on  the  oce- 
anic nitrogen  budget,  Global  Biogeochem. 
Cycles,  1(2),  97-116,  1987. 

Cifuentes,  L.A.,  J.  H.  Sharp,  and  M.  L.  Fogel, 
Stable  carbon  and  nitrogen  isotope  biogeo- 


chemistry  in  the  Delaware  estuary,  Limnol. 
Oceangr.,33,  1102-1115,  1988. 

Cline,  J.  D.  and  I.  R.  Kaplan,  Isotopic  fractionation 
of  dissolved  nitrate  during  denitrification  in 
the  eastern  Tropical  North  Pacific,  Mar.  Chem., 
5,271-299,  1975. 

Klump,  J.  V.,  and  C.  S.  Martens,  Benthic  nitrogen 
regeneration,  in  Nitrogen  in  the  Marine  Envi- 
ronment, Chapter  12,  EJ.  Carpenter  and  D.G. 
Capone,  eds.,  Academic  Press,  New  York,  pp. 
411-458,1983. 

Macko,  S.  A.,  and  M.  L.  F.  Estep,  Microbial 
alteration  of  stable  nitrogen  and  carbon  isoto- 
pic compositions  of  organic  matter,  Org. 
Geochem.,  6,  787-790,  1984. 

Montoya,  J.  P.,  S.  G.  Horrigan,  and  J.  J.  McCarthy, 
Natural  abundance  of  15N  in  particulate  nitro- 
gen and  zooplankton  in  the  Chesapeake  Bay, 
Mar.  Ecol.  Prog.  Series,  65,  35-61,  1990. 

Rau,  G.  H.,  M.  A.  Arthur,  and  W.  E.  Dean,  15N/14N 
variations  in  Cretaceous  Atlantic  sedimentary 
sequences:  implication  for  past  changes  in 
marine  nitrogen  biogeochemistry,  Earth 
Planet.  Sci.  Let.,  82,  269-279,  1987. 

Reeburgh,  W.  S.,  An  improved  interstitial  water 
sampler,  Limnol.  Oceanogr.,  12,  223-234, 
1967. 

San  Diego-McGlone,  M.  L.,  Processes  affecting 
the  behavior  of  redox  sensitive  elements  in 
Chesapeake  Bay,  Ph.D  dissertation,  Old  Do- 
minion University,   1991. 

Sharp,  J.  H.,  C.  H.  Culberson,  and  T.  M.  Church, 
The  chemistry  of  the  Delaware  estuary.  Gen- 
eral Considerations,  Limnol.  Oceanogr.,  27, 
1015-1028,  1982. 

Sweeney,  R.  E.  and  I.  R.  Kaplan,  Natural  abun- 
dances of  15N  as  a  source  indicator  for  near- 
shore  marine  sedimentary  and  dissolved  nitro- 
gen, Mar.  Chem.,  9,  81-94,  1980. 

Velinsky,  D.  J.  and  G.  A.  Cutter,  Geochemistry  of 
selenium  in  a  coastal  salt  marsh,  Geochim. 
Cosmochim.Acta,55,  179-191,  1991. 

Velinsky,  D.  J.,  J.  R.  Pennock,  J.  H.  Sharp,  L.  A. 
Cifuentes  and  M.  L.  Fogel,  Determination  of 
the  isotopic  composition  of  ammonium- nitro- 
gen at  the  natural  abundance  level  from  estua- 
rine  waters,  Mar.  Chem.,  26,  351-361,  1989. 

Wada,  E.,  T.  Kadonaga  and  S.  Matsuo,  15N  abun- 
dance in  nitrogen  of  naturally  occurring  sub- 
stances and  global  assessment  of  denitrifica- 
tion from  isotopic  viewpoint,  Geochem.  Jour., 
9,  139-145,  1975. 

Westrich,  J.  T.,  and  R.  A.  Berner,  The  role  of 
sedimentary  organic  matter  in  bacterial  sulfate 
reduction:  The  G  model  tested,  Limnol. 
Oceangr.,  29(2),  236-249,  1984. 


GEOPHYSICAL  LABORATORY 


163 


The  Isotopic  Ecology  Of  Plants  And 

Animals  In  Amboseli  National  Park, 

Kenya 

Paul  L.  Koch,  Anna  K.  Behrensmeyer* 
and  Marilyn  L.  Fogel 

Variations  in  the  stable  isotope  ratios  of 
carbon  ("C/1^),  nitrogen  ("N/^N),  and 
oxygen  ("O/^O)  provide  information  about 
the  ecology,  physiology,  and  habitats  of 
living  and  extinct  animals.  For  example, 
the  ^3C  values  of  an  animal's  tissues  are 
controlled  by  the  isotopic  composition  of 
its  diet,  which,  for  herbivores,  is  related  to 
the  photo  synthetic  pathway  of  food  plants 
(DeNiro  and  Epstein,  1978;  Vogel,  1978). 
Although  affected  by  dietary  #5N  values, 
N  isotopes  in  animals  vary  with  rainfall 
amounts  among  ecosystems  and  among 
trophic  levels  in  an  ecosystem  (DeNiro  and 
Epstein,  1981;  Heaton  et  ai,  1986). 

We  are  investigating  the  isotopic 
ecology  of  plants  and  animals  in  Amboseli 
National  Park,  Kenya,  for  several  reasons. 
First,  investigations  of  floral  and  faunal 
isotopic  composition  in  terrestrial  ecosys- 
tems are  uncommon  (e.g.,  Ambrose  and 
DeNiro,  1986;  Sealy  era/.,  1987),andnone 
evaluates  C,  N,  and  O  isotopes  simulta- 
neously. Ecosystem  studies  test  the  gen- 
erality of  relationships  determined  either  in 
the  laboratory  or  in  comparisons  of  indi- 
viduals from  different  regions.  Second, 
stable  and  radiogenic  isotopes  have  been 
employed  to  identify  sources  of  elephant 

*Department  of  Paleobiology,  National  Museum 
of  Natural  History,  Smithonian  Institution,  Wash- 
ington, DC 


ivory  and  rhinoceros  horn,  in  order  to  con- 
trol sales  of  poached  versus  legally  hunted 
animals  (van  der  Merwe  etal.,  1990;  Vogel 
et  ai,  1990).  Ivory  from  various  African 
parks  can  be  distinguished  by  its  N,  C,  and 
either  Sr  or  Pb  isotopic  composition.  How- 
ever, if  the  isotopic  composition  of  el- 
ephants varies  with  time,  because  of  habi- 
tat, diet,  or  climate  change,  isotopic  identi- 
fication of  source  region  may  be  unreli- 
able. Either  the  isotopic  composition  of  a 
species  must  be  constant  through  time 
within  an  ecosystem,  or  the  secular  trends 
must  be  minor  when  compared  to  differ- 
ences between  populations.  Finally,  isoto- 
pic patterns  in  modern  ecosystems  can  serve 
as  analogs  for  interpretation  of  the  fossil 
record.  African  faunas,  with  their  diversity 
of  large  mammals,  are  excellent  analogs  of 
typical  faunas  before  the  Pleistocene  ex- 
tinction. 


Study  Area,  Materials,  and  Methods 

Amboseli  Park  is  located  in  southern 
Kenya  (20°40'S,37°15'E;  mean  elevation, 
1140  m).  Annually,  temperature  averages 
23°C  and  ranges  from  15°  to  31°C.  Rain 
falls  in  two  seasons  and  averages  300  cm/ 
year.  However,  the  park  is  continuously 
supplied  with  spring  water  fed  by  melting 
snow  on  Mt.  Kilimanjoro.  Habitats  in  the 
park  include  grasslands,  bushlands, 
swamps,  seasonal  lakes,  and  woodlands. 
Woodlands  have  retreated  since  the  early 
1970s,  perhaps  due  to  overbrowsing  by 
elephants  or  increased  soil  salinity.  Tree 
loss  has  altered  the  abundances  of  herbi- 


164 


CARNEGIE  INSTITUTION 


vores;  there  are  more  grazers  (animals  that 
eat  grass)  and  fewer  browsers  (animals 
eating  herbaceous  and  woody  plants)  and 
mixed  feeders  (D.  Western,  pers.  comm.). 

Plant  samples  (mixtures  of  leaves  and 
stems)  were  collected  in  September  1 990  at 
eight  localities  in  the  woodland,  swamp, 
swamp  edge,  plains,  and  bush  habitats. 
Faunal  samples  (tooth  dentin  or  bone)  were 
collected  from  carcasses  throughout  the 
park.  Samples  were  collected  from  1975 
through  1990,  and  were  in  different  states 
of  weathering.  The  minimum  number  of 
years  since  death  can  be  estimated  from 
weathering  stage  (Behrensmeyer,  1978). 
For  carcasses  in  advanced  weathering 
stages,  however,  determining  actual  time 
since  death  is  difficult. 

Plants  were  air  dried  in  the  field,  freeze- 
dried  in  the  laboratory,  and  then  lightly 
crushed.  Bones  and  teeth  were  demineral- 
ized  with  EDTA  or  0.1  N  HC1  to  isolate 
collagen  (Tuross  et  ai,  1988),  and  then 
treated  with  chloroform/methanol  solution 
to  remove  lipids.  Plant  and  collagen  samples 
were  placed  in  preheated  quartz  tubes  with 
CuO  and  Cu  metal.  Tubes  were  evacuated, 
sealed,  combusted  at  910°C  for  2  h,  then 
cooled  at  a  controlled  rate.  Standard  devia- 
tions for  analysis  of  standards  were  ±0.2%c 
for  #3C  and  S^N. 


I  so  topic  Variation  in  Amboseli  Plants 

The  plants  of  Amboseli  segregate  into 
two  populations  isotopically  (Table  25,  Fig. 

92A).  The  5  C  values  of  grasses,  which 
use  C4  photosynthesis,  have  a  mean  value 


iu  ■ 

A 

E 

aquatic 

E 

succulent 

8  - 

ID 

shrub  &  tree 

w 

□ 

herb 

3 

■ 

grass 

.y  e- 

- 

- 

"O 

c 

o   4- 

-         r 

o 

„ 

n 

n 

2  - 

i 

_ 

. 

-      X 

n  - 

WW 

n 

"n 

613C 


10 


8 
tn 
aJ 

|  6 

C 

o   4 

d 

z 


2- 


1 


^    :, 


□  elephant 

□  mixed  feeder 

□  browser 
■  grazer 

Q  carnivore 


fiii 


^       00       O        OJ       ^t       CD 


613C 


Fig.  92.  (A)  Histogram  of  carbon  isotope  compo- 
sitions for  Amboseli  plants  subdivided  according 
to  physiogamy.  (B)  Histogram  of  carbon  isotope 
compositions  for  Amboseli  mammals  subdivided 
according  to  feeding  type. 


±  one  standard  deviation  of  -13.2  ±  0.9  %o. 
The  herbaceous  plants  and  woody  plants 

employ  CL  photosynthesis  and  have  mean 
values  of -27.1  ±  1.9  %o and  -27.7  ±  2.4  V 
respectively.     This  isotopic  segregation 

between  CL  shrubs,  trees  and  herbs  and  C4 
grasses  is  expected  in  a  hot,  dry  region 
(Tieszen  and  Boutton,  1988).  Succulent 
herbaceous  and  woody  plants  exhibit  a 


GEOPHYSICAL  LABORATORY 


165 


Table  25.  Isotopic  data  for  Amboseli  plants  collected  in  1990 


Species 

Family 

Habitat 

8^N 

5l3C 

Grasses  and  Sedges 

Cynodon  dactylon 

Graminae  (m) 

swamp 

10.6 

-13.0 

Sporobolus  consimilis 

Graminae  (m) 

swamp  edge 

8.9 

-13.4 

Sporobolus  spicatus 

Graminae  (m) 

swamp  edge 

8.6 

-13.2 

Sporobolus  kentrophyllum 

Graminae  (m) 

swamp  edge 

10.1 

-13.4 

Cynodon  plectostachys 

Graminae  (m) 

woodland 

9.4 

-14.2 

Sporobolus  helvolus 

Graminae  (m) 

bush 

8.8 

-15.1 

Sporobolus  ioclades 

Graminae  (m) 

bush 

11.9 

-13.0 

Chloris  roxburghiana 

Graminae  (m) 

bush 

8.7 

-13.4 

Chloris  virgata 

Graminae  (m) 

bush 

10.6 

-13.2 

Enneapogon  cenchroides 

Gramiriae  (m) 

bush 

9.4 

-13.3 

Cyperus  immensus 

Cyperaceae  (m) 

swamp 

4.4 

-11.2 

Cyperus  laevigatus 

Cyperaceae  (m) 
Submerged  Aquatic  Plants 

swamp 

7.8 

-12.2 

Ceratophyllum  sp.  1 

Ceratophyllaceae 

swamp 

6.4 

-23.0 

Ceratophyllum  sp.  2 

Ceratophyllaceae 
Herbs 

swamp 

9.3 

-19.7 

Pistia  stratiotes 

Araceae  (m) 

swamp 

11.5 

-29.0 

Solanum  incanum 

Solanaceae 

swamp  edge 

10.4 

-25.0 

Justicia  odora 

Acanthaceae 

woodland 

13.3 

-27.6 

Diplictera  albicauda 

n.d. 

woodland 

8.0 

-27.8 

Abutilon  mauritanium 

Malvaceae 

plains 

11.0 

-29.7 

Pluchea  ovalis 

Asteraceae 

plains 

8.1 

-30.5 

Cissampelos  mucronata 

Menispermaceae 

plains 

7.5 

-27.7 

Commicarpus  sp. 

Nyctaginaceae 

plains 

8.6 

-28.9 

Achyranthes  aspera 

Amaranthaceae 

plains 

11.4 

-28.4 

Withania  somnifera 

Solanaceae 

plains 

9.4 

-25.2 

Indigofera  sp. 

Leguminosae 

bush 

10.4 

-25.0 

Duosperma  eremophiloum 

Acanthaceae 

bush 

8.2 

-27.1 

Barleria  spinisepala 

Acanthaceae 

bush 

11.2 

,24.3 

Shrubs.  Trees,  and  Succulent  Plants 

Trianthema  ceratosepala 

Aizoaceae 

bush 

12.8 

-21.8 

Sansevieria  sp. 

Agavaceae  (m) 

bush 

12.6 

-14.5 

Euphorbia  sp.  1 

Euphorbiaceae 

bush 

13.3 

-14.9 

Euphorbia  sp.  2 

Euphorbiaceae 

bush 

13.9 

-14.0 

Sueda  monoica 

Chenopodiaceae 

swamp  edge 

13.6 

-13.3 

Continued  on  next  page 


166 


CARNEGIE  INSTITUTION 


Table  25.  Continued 


Species 


Family 


Habitat  515N       <513C 


Salvador  a  persica 
Maerua  triphyllum 
Commiphora  sp. 
Boscia  angustifolia 
Acacia  sp. 
Azima  tetracantha 
Balanites  glabra 
Phoenix  reclinata 
Acacia  tortilis 


Shrubs  and  Trees 

Salvadoraceae 

swamp  edge 

5.3 

-25.4 

Capparidaceae 

bush 

8.8 

-29.3 

Burseraceae 

bush 

15.4 

-29.3 

Capparidaceae 

bush 

11.2 

-24.6 

Leguminosae 

bush 

10.2 

-23.7 

Salvadoraceae 

woodland 

7.6 

-26.8 

Balanitaceae 

woodland 

7.7 

-29.9 

Palmae  (m) 

woodland 

7.1 

-29.4 

Leguminosae 

woodland 

8.9 

-29.7 

(m)  indicates  monocotyledons,  all  other  plants  are  dicotyledons 


range  of  8   C  values  and  may  use  either  C. 

or  Crassulacean  acid  metabolism.  Finally, 

i  ^ 
submerged  aquatic  plants  have  5  C  values 

that  can  range  between  -12  and  -33  %o, 
depending  on  the  pathway  of  carbon  up- 
take. Unlike  terrestrial  plants,  which  di- 
rectly incorporate  atmospheric  CCL,  sub- 
merged plants  may  accumulate  either  dis- 
solved C02  or  HC03"  (Raven,  1987). 
Amboseli  aquatic  plants  have  #3C  values 
intermediate  between  C3  and  C.  plants. 

The  <515N  values  of  Amboseli  plants 
form  a  unimodal  distribution  with  a  mean 
of  9.8  ±  2.4  %o  (Table  25,  Fig.  93 A).  This 
mean  value  is  slightly  higher  than  that 
reported  by  Sealy  et  al.  (1987)  for  plants 
from  a  region  receiving  300  mm  of  rain. 

Plant  <515N  values  are  not  dependent  on 
either  location  or  habitat  type,  although 
plants  from  the  bush  habitat  may  be  slightly 

15N-enriched.  Plant  <515N  values  are  not 
influenced  by  physiogamy  or  photosyn- 
thetic  pathway,  with  one  exception.  All 

Amboseli  succulents  are  15N-enriched  (1 3.2 


±  0.5  %c).  These  species  are  only  distantly 
related  to  each  other,  and  the  cause  of  15N 
enrichment  is  unclear. 


Isotopic  Variation  in  Amboseli  Mammals 

The  C  isotope  difference  between  C3 
and  C4  plants  provides  a  tool  for  tracing  the 
diets  of  Amboseli  mammals.  Previous  field 
studies  demonstrate  a  consistent  difference 
in  <513C  values  between  diet  and  collagen 
of  ~+5  %o  (Vogel  1978;  van  der  Merwe, 
1989).  Amboseli  mammals  with  pure  graz- 
ing diets  should  have  <513C  values  of  -8  to 
-9  %o.  All  Amboseli  grazers  (buffalo,  spring 
hare,  warthog,  wildebeest,  zebra)  have  col- 
lagen <513C  values  in  this  range  (Table  26, 
Fig.  92B). 

In  contrast,  Amboseli  animals  with  pure 
browsing  diets  should  have  collagen  <513C 
values  of  -22  to  -23  %o.  None  of  the 
browsers  (rninoceros,  giraffe)  have  values 
this  low,  indicating  a  small  but  persistent 
fraction  of  grasses  or  succulents  in  their 


GEOPHYSICAL  LABORATORY 


167 


Table  26.  Isotopic  data  for  Amboseli  mammals,  excluding  elephants. 


Secies 

Common  Name 

Year  of 
death 

<515N 

<5l3C 

Carnivores.  Insectivores.  and  Omn 

ivores 

Crocuta  crocuta 

Spotted  hyena 

*1984 

16.7 

-9.2 

Panthera  leo 

Lion 

n.d. 

17.1 

-10.1 

Panthera  leo 

Lion 

n.d. 

15.6 

-6.4 

A  cinonyx  jubatus 

Cheetah 

*1984 

17.2 

-15.7 

Canis  adustus 

Jackal 

*1989 

15.4 

-11.2 

Canis  adustus 

Jackal 

*1988 

13.8 

-9.9 

Canis  adustus 

Jackal 

*1988 

14.9 

-14.8 

Otocyon  megalotis 

Bat-eared  fox 

*1971 

12.2 

-10.6 

Otocyon  megalotis 

Bat-eared  fox 

n.d. 

14.6 

-12.0 

Otocyon  megalotis 

Bat-eared  fox 

*1974 

14.2 

-15.7 

Ichneumia  albicauda 

White-tailed  mongoose 

n.d. 

17.4 

-8.3 

Orycteropus  afer 

Aardvark 

n.d. 

9.7 

-12.1 

Papio  cynocephalus 

Yellow  baboon 

Grazers 

1989 

10.8 

-14.3 

Pedetes  capensis 

Spring  hare 

*1975 

9.8 

-9.0 

Equus  burchelli 

Burchell's  zebra 

1974 

9.8 

-8.6 

Equus  burchelli 

Burchell's  zebra 

*1984 

9.1 

-8.6 

Equus  burchelli 

Burchell's  zebra 

1990 

10.0 

-8.5 

Connochaetes  taurinus 

White-bearded  wildebeest 

1975 

12.4 

-8.0 

Connochaetes  taurinus 

White-bearded  wildebeest 

1974 

11.0 

-8.7 

Connochaetes  taurinus 

White-bearded  wildebeest 

*1988 

13.8 

-7.8 

Connochaetes  taurinus 

White-bearded  wildebeest 

*1989 

11.8 

-9.0 

Syncerus  coffer 

Buffalo 

1968 

10.1 

-7.9 

Syncerus  caffer 

Buffalo 

*1984 

10.8 

-8.0 

Phacochoerus  aethiopicus 

Warthog 

1990 

10.8 

-8.8 

Phacochoerus  aethiopicus 

Warthog 

Browsers 

*1989 

11.0 

-9.2 

Dicer os  bicornis 

Black  rhinoceros 

1961 

6.4 

-19.8 

Dicer os  bicornis 

Black  rhinoceros 

1974 

8.2 

-18.7 

Diceros  bicornis 

Black  rhinoceros 

*1984 

7.9 

-19.0 

Giraffa  camelopardalis 

Giraffe 

*1989 

11.5 

-20.3 

Giraffa  camelopardalis 

Giraffe 

Mixed  Feeders 

*1986 

11.6 

-19.7 

Hystrix  cristata 

Porcupine 

*1973 

9.9 

-15.6 

Hippopotamus  amphibius 

Hippopotamus 

*1973 

8.9 

-10.0 

Hippopotamus  amphibius 

Hippopotamus 

1990 

13.1 

-8.6 

Hippopotamus  amphibius 

Hippopotamus 

*1989 

8.9 

-10.9 

Aepyceros  melampus 

Impala 

1985 

12.4 

-14.0 

Aepyceros  melampus 

Impala 

*1986 

11.7 

-13.6 

Aepyceros  melampus 

Impala 

*1988 

13.1 

-15.8 

Gazella  granti 

Grant's  gazelle 

*1988 

10.5 

-16.0 

Gazella  granti 

Grant's  gazelle 

*1989 

10.2 

-15.6 

Gazella  thomsoni 

Thomson's  gazelle 

1990 

14.2 

-10.7 

Gazella  thomsoni 

Thomson's  gazelle 

*1986 

10.8 

-17.6 

*  determined  by  weathering  stage 


168 


20 


w    15 

CO 

■g 

"> 

■o 

c 


10   11    12   13  14   15  16   17  18 


815N 


10  - 


5- 


□  elephant 

□  mixed  feeder 

□  browser 
■  grazer 
H  carnivore 


J=L 


3     4     5     6     7 


10  11   12  13  14  15  16  17  18 


CARNEGIE  INSTITUTION 

Table  27.  Isotopic  data  for  Amboseli  elephants. 


Secimen 


Year  of 
death 


<515N         «513C 


African  Elephant:  Loxodonta  africana 


C75-3 

1974 

12.0 

-18.3 

C75-6 

*1973 

9.4 

-17.9 

E-l                 ) 

late  70s 

10.4 

-13.3 

E-3                 ] 

late  70s 

10.3 

-13.9 

E-8                 ] 

late  70s 

9.7 

-17.3 

E-ll               ] 

late  70s 

10.7 

-13.2 

E-13               ] 

late  70s 

11.4 

-13.9 

E-15               ] 

late  70s 

11.9 

-13.7 

E-17               ] 

late  70s 

10.3 

-15.3 

E-19               ] 

late  70s 

10.1 

-13.4 

E-21               ] 

late  70s 

10.4 

-14.9 

E-23               ] 

late  70s 

10.4 

-11.9 

E-25               ] 

late  70s 

10.2 

-15.5 

E-27               ] 

late  70s 

10.4 

-15.5 

E-29               ] 

late  70s 

10.3 

-13.7 

E-30               ] 

late  70s 

10.7 

-11.9 

E-34               ] 

late  70s 

9.8 

-14.2 

All  specimens  with  Year  of  death  of  late  70s  were 
collected  by  Cynthia  Moss  and  have  known  dates 
of  death  that  we  have  not  yet  received.  For  Fig. 
94 A,  these  animals  are  plotted  as  deaths  in  1978. 


Fig.  93.  (A).  Histogram  of  nitrogen  isotope  com- 
positions for  Amboseli  plants  subdivided  accord- 
ing to  physiogamy.  (B)  Histogram  of  nitrogen 
isotope  compositions  for  Amboseli  mammals  sub- 
divided according  to  feeding  type.  Note  that  the 
difference  between  the  mean  £15N  values  of  plants 
and  animal  collagen  is  only  ~+l  %o. 


diets.  Animals  known  to  eat  a  mixture  of 
plants  (elephant,  Grant's  and  Thomson's 
gazelle,  hippopotamus,  impala,  and  porcu- 
pine) have  <513C  values  intermediate  be- 
tween browsers  and  grazers  (Tables  26  and 
27). 

The  link  between  the  <513C  of  diet  and 
collagen  is  more  difficult  to  unravel  in 
carnivores  and  omnivores,  because  these 
animals  obtain  carbon  both  from  different 
tissues  within  a  body  (fat,  muscle,  skin)  and 


from  plants.  All  these  sources  may  have 
different  <513C  values.  Generally,  herbi- 
vore meat  and  carnivore  collagen  differ  by 
~  +5%o,  but  the  difference  between  herbi- 
vore collagen  and  carnivore  collagen  is 
+2%c  (van  der  Merwe,  1989).  Observa- 
tions of  hunting  carnivores  suggest  that 
hyena  and  lion  consume  chiefly  C4-feed- 
ing  herbivores  (wildebeest  and  zebra), 
whereas  cheetah  eat  mixed  feeders 
(Thomson's  and  Grant's  gazelle  and  im- 
pala). These  observations  are  supported  by 
#3C  values  (Table  26,  Fig.  92B).  The 
smaller  carnivores  (fox,  jackal,  mongoose) 
eat  smaller  animals  from  across  the  dietary 
spectrum  and  exhibit  a  spread  of  <513C 
values. 


GEOPHYSICAL  LABORATORY 


169 


IB  - 
16- 

A 

a  Carnivores 

•  Browser  &  Mixed 

14- 

A# 

10 
to 

12- 
10- 

Carnivore           A    • 

4 

• 

•              *" 

8- 

Browser  &  Mixed 

• 
• 

6- 

1 1 1 r 

1955    1960    1965    1970    1975   1980    1985    1990    1995 

Year  of  Death 


16 
14 
12 
10  H 

z 

£        8 
to 

6- 

4- 
2- 


-f- 


4 

hT~'teHrH 


hh 


* 


-30 


-25 


I         I       '»        I        |         ill         l 


-20 
813C 


-15 


-10 


A    Rhinoceros 
•    Elephant 
*±*  Mean  &  std.  dev.  of 
African  park  elephants 


Fig.  94.  (A)  Secular  variation  in  the  nitrogen 
isotope  composition  of  mammals.  Grazers:  Y= 
-68.25  +.0.04X  r=0.24,  slope  is  not  significantly 
different  from  0.  Data  and  regression  line  are  not 
plotted.  Browsers:  Y=  -279.33  +  0.15X  r=0.62, 
slope  is  significantly  different  than  0.  Carnivores: 
Y=  -253.66  +  0. 14X  r=0.58,  slope  is  significantly 
different  than  0.  The  elephants  listed  as  late  70s 
deaths  are  included  on  the  figure,  and  given  1978 
as  the  year  of  death,  but  they  were  not  used  in  the 
regression  calculation.  (B)  Carbon  and  nitrogen 
isotopic  composition  of  Amboseli  elephants  and 
rhinoceros.  Also  plotted  are  the  means  and  stan- 
dard deviations  for  elephant  ivory  from  16  other 
African  parks  and  preserves.  Firm  determination 
of  temporal  isotopic  trends  within  species  must 
await  analysis  of  specimens  with  a  greater  range  of 
known  ages  of  death.  However,  the  spread  in  the 
isotopic  data  from  both  species  may  result  from 
coupled  increases  in  <513C  and  <515N  values  with 
time.  Although  values  from  Amboseli  elephants 
do  not  overlap  with  values  from  many  other  parks, 
they  vary  by  amounts  as  great  as  those  used  to 
discriminate  between  other  park  populations. 


A  fractionation  of  ~  +3%c  between  the 
<515N  value  of  plant  food  and  herbivore 
collagen  has  been  reported  in  previous 
studies  (DeNiro  and  Epstein,  1981 ;  Hare  et 
ai,  1991).  Given  this  fractionation,  the 
average  Amboseli  herbivore  should  have  a 
collagen  Sl 5N  value  of  1 2- 1 3  %o.  Although 
some  animals  have  values  in  this  range, 
most  are  more  negative  (Tables  26  and  27, 
Fig.  93B)  Indeed,  using  the  mean  ^5N  of 
animals  and  a  fractionation  of  +3  %o,  we 
would  expect  dietary  plants  with  values  of 
l%o  or  less.  Few  plants  within  the  park 
have  isotopic  values  this  low. 

There  are  several  plausible  explana- 
tions for  this  discrepancy.   First,  we  ana- 


lyzed only  plants  collected  in  a  dry  season. 
In  the  wet  season,  plants  may  have  lower 
<515N  values.  Second,  if  the  515N  value  of 
plants  from  Amboseli  has  increased  re- 
cently, current  vegetation  may  not  be  repre- 
sentative of  the  foods  eaten  by  the  sampled 
animals .  Finally,  the  fractionation  between 
diet  and  herbivore  collagen  may  not  be 
+3%o.  Variability  in  this  fractionation  has 
been  detected  previously  (Ambrose  and 
DeNiro,  1986;  Heatonef  a/.,  1986;  Sealy  et 
al.,  1987),  and  attributed  to  differences  in 
N  metabolism  between  different  herbivores. 
A  fractionation  of  ~+\%o  is  observed  be- 
tween current  Amboseli  plants  and  the 
sampled  herbivores. 


170 


CARNEGIE  INSTITUTION 


Differences  in  collagen  <515N  between 
herbivores  and  carnivores  are  well  studied 
and  range  from  +3  to  +6  %o  (Schoeninger 
and  DeNiro,  1984;  Ambrose  and  DeNiro, 
1986;  Sealy  et  al.,  1987).  Amboseli  graz- 
ers, mixed  feeders  and  browsers  averaged 
10.9  ±  1.3  %o,  9.1  ±  2.3  %o9  and  10.9  ±  1.3 
%o,  respectively,  whereas  true  Amboseli 
carnivores  averaged  15.4  ±  \.6%c.  Conse- 
quently, there  is  a  trophic  level  fraction- 
ation of  ~+5  %o.  The  omnivorous  yellow 
baboon  has  a  lower  <515N  value,  suggesting 
a  preponderance  of  plant  foods  in  the  diet. 
Finally,  the  aardvark,  which  consumes  ants 
and  termites,  has  a  <515N  value  within  the 
herbivore  range.  However,  insect  chitin  is 
known  to  be  15N-depleted  relative  to  di- 
etary plants  (Schimmelmann,pers.  comm.), 
which  would  lead  to  lower  values  in  the 
collagen  of  insectivores  relative  to  carni- 
vores. 


Secular  Trends  in  the  Isotopic  Composi- 
tion of  Amboseli  Mammals 

The  Amboseli  ecosystem  has  changed 
dramatically  since  1960  because  of  a  loss 
of  trees  and  the  expansion  of  grassland.  To 
document  isotopic  trends  in  park  mammals, 
multiple  samples  of  individual  species  from 
different  time  periods  must  be  examined. 
Our  data  are  currently  insufficient  for  such 
a  treatment,  but  trends  within  broad  feed- 
ing categories  may  be  examined.  Collagen 
<515N  and  <513C  values,  and  thus  the  diet  of 
grazers,  have  remained  constant  from  1968 
to  1990  (Table  26).  The  &$N  of  browsers 
and  mixed  feeders  has  increased  by  a  sta- 
tistically significant  amount  (Fig.  94A). 


This  trend,  however,  is  strongly  influenced 
by  a  low  value  for  a  single  old  specimen. 
Although  most  elephant  deaths  are  only 
roughly  dated  to  the  late  1 970s,  the  popula- 
tion seems  to  be  trending  towards  higher 
#5n  and  8&C  values  (Fig.  94B).  The 
magnitude  of  this  variation  is  significant 
when  compared  to  the  differences  between 
populations  from  different  parks.  There  is 
a  suggestion  of  similar  coupled  increases 
for  rhinoceros,  but  the  sample  is  quite  small. 
Finally,  carnivores  also  increase  in  <515N 
with  time  by  amount  similar  to  browsers 
(Fig.  94A). 

We  hypothesize  that  as  the  park  was 
stripped  of  trees,  browsers  and  mixed  feed- 
ers have  been  forced  to  consume  more 
grass.  Increased  grass  consumption  is  par- 
ticularly evident  for  elephants.  However, 
grass  is  relatively  nutrient-poor  compared 
to  browse.  Eventually,  the  browsers  and 
mixed  feeders  suffered  nutritional  stress. 
Nutritional  stress  may  cause  an  animal  to 
remetabolize  previously  deposited  proteins, 
and  can  potentially  produce  an  increase  in 
collagen  <515N  in  bones  equivalent  to  that 
generated  by  feeding  at  a  higher  trophic 
level  (Tuross,  pers.  com.).  The  grazers 
thrived  as  the  grasslands  expanded  and 
exhibit  no  isotopic  changes.  Carnivores 
eat  both  types  of  herbivores,  and  conse- 
quently exhibit  intermediate  isotopic  trends. 


Conclusions 

The  carbon  and  nitrogen  isotopes  in 
most  plants  from  Amboseli  National  Park 
varied  as  expected,  with  a  strong  differen- 
tiation in  5J  3C  between  C3,  C4,  and  aquatic 


GEOPHYSICAL  LABORATORY 


171 


plants,  and  a  high  mean  d^N  value.  The 
15N  enrichment  of  succulent  plants  was 
unexpected,  and  is  currently  unexplained. 
Differences  in  the  <513C  of  plants  is  re- 
flected in  the  collagen  of  the  animals  that 
consume  these  plants,  and  ultimately  can 
be  detected  at  higher  trophic  levels  when 
these  herbivores  are  preyed  upon  by  carni- 
vores. The  fractionations  of  C  and  N  iso- 
topes between  diet  and  collagen  that  we 
discovered  match  previous  reports,  with 
one  exception.  The  fractionation  of  N 
between  plant  and  herbivore  collagen  was 
much  lower  than  expected.  Finally,  al- 
though our  observations  must  be  supported 
by  more  extensive  sampling,  there  are  sta- 
tistically significant  secular  trends  in  the 
<515N  of  Amboseli  browsers  and  mixed 
feeders  and  carnivores,  whereas  grazers 
are  invariant.  The  substantial  isotopic  trends 
shown  by  Amboseli  elephants  may  indi- 
cate that  stable  isotopes  will  be  of  limited 
utility  in  tracing  the  source  of  elephant 
ivory  in  changing  habitats. 


References 

Ambrose,  S.  H.,  and  M.  J.  DeNiro,  The  isotopic 
ecology  of  East  African  mammals,  Oecologia, 
69,  395-406,  1986. 

Behrensmeyer,  A.  K.,  Taphonomic  and  ecologic 
information  from  bone  weathering,  Paleobio., 
4, 150-162,  1978. 

DeNiro,  M.  J.,  and  S.  Epstein,  Influence  of  diet  on 
the  distribution  of  carbon  isotopes  in  animals, 
Geochim.  Cosmochim.  Acta,  42,  495-506, 
1978. 

DeNiro,  M.  J.,  and  S.  Epstein,  Influence  of  diet  on 
the  distribution  of  nitrogen  isotopes  in  ani- 
mals, Geochim.  Cosmochim.  Acta,  45,  341- 
351,  1981. 

Hare,  P.  E.,  M.  L.  Fogel ,  T.  W.  Stafford,  Jr.,  A.  D. 
Mitchell,  and  T.  C.  Hoering,  The  isotopic 
composition  of  carbon  and  nitrogen  in  indi- 
vidual amino  acids  isolated  from  modern  and 


fossil  proteins,  J.  Archaeol.  Sci,  in  press,  199 1 . 

Heaton,  T.  E.,  J.  C.  Vogel,  G.  von  la  Chevallerie, 
and  G.  Collett,  Climatic  influence  on  the  iso- 
topic composition  of  bone  nitrogen,  Nature, 
322,  822-823,  1986. 

Raven,  J. .,  The  application  of  mass  spectrometry 
to  biochemical  and  physiological  studies,  in 
The  Biochemistry  of  Plants,  Vol.  13,  Aca- 
demic Press,  Inc.,  New  York,  pp.  127-180, 
1987. 

Schoeninger,  M.  J.,  and  M.  J.  DeNiro,  Nitrogen 
and  carbon  isotopic  composition  of  bone  col- 
lagen from  marine  and  terrestrial  animals, 
Geochim.  Cosmochim.  Acta,  48,  625-639 
1984. 

Sealy,  J.C.,  N.  J.  van  der  Merwe,  J.  A.  Lee  Thorp, 
and  J.  L.  Lanham,  Nitrogen  isotopic  ecology 
in  southern  Africa:  implications  for  environ- 
mental and  dietary  tracing,  Geochim. 
Cosmochim.  Acta,  57,2707-2717,  1987. 

Tieszen,  L.  L.,  and  T.  W.  Boutton,  Stable  carbon 
isotopes  in  terrestrial  ecosystem  research,  in 
Stable  Isotopes  in  Ecological  Research, 
Rundel,  P.W.,  J.R.  Ehleringer,  and  K.  A.  Nagy , 
eds.,  Springer- Verlag,  New  York,  pp.  167- 
195,  1988 

Tuross,  N.,M.  L.  Fogel,  andP.  E.  Hare,  Variability 
in  the  preservation  of  the  isotopic  composition 
of  collagen  from  fossil  bone,  Geochim. 
Cosmochim.  Acta,  52,  929-935,  1988. 

van  der  Merwe,  N.  J.,  Natural  variations  in  13C 
concentration  and  its  effect  on  environmental 
reconstruction  using  13C/12C  ratios  in  animal 
bones,  in  The  Chemistry  of  Prehistoric  Human 
Bone,  Price,  T.D.,  ed.,  Cambridge  Univ.  Press, 
New  York,  pp.  105-125,  1989. 

van  der  Merwe,  N.  J.,  J.  A.  Lee-Thorp,  J.  F. 
Thackeray,  A.  Hall-Martin,  F.  J.  Kruger, 
H.Coetzee,  R.  H.  V.  Bell,  and  M.  Lindeque, 
Source-area  determination  of  elephant  ivory 
by  isotopic  analysis,  Nature,  346,  744-746, 
1990. 

Vogel,  J.  C,  Isotopic  assessment  of  the  dietary 
habits  of  ungulates,  S.  Afr.  J.  Sci.,  74, 298-30 1 , 
1978. 

Vogel,  J.  C,  B.  Eglington,  and  J.  M.  Auret,  Isotope 
fingerprints  in  elephant  bone  and  ivory,  Na- 
ture, 346,  747-749,  1990. 


172 


CARNEGIE  INSTITUTION 


Rapid  Racemization  of  Aspartic  Acid 

in  mollusk  and  ostrich  eggshells : 

A  New  Method  for  Dating  on 

a  Decadal  Time  Scale 

Glenn  A.  Goodfriend,  David  W.  von 
Endt*  and  RE.  Hare 

Epimerization  of  L-isoleucine  to  D- 
alloisoleucine  has  been  extensively  ana- 
lyzed in  mollusk  shells  as  a  means  of  deter- 
mining relative  or  absolute  ages,  primarily 
of  Pleistocene  samples.  More  recently, 
epimerization  in  Pleistocene  ostrich  egg- 
shells has  been  studied  (Brooks  et  al. ,  1 990). 
Racemization  of  a  number  of  other  amino 
acids  has  been  studied  in  mollusk  shells, 
leucine  being  the  most  widely  studied  (e.g., 
Wehmiller,  1984;  reviewed  in  Goodfriend, 
1991).  Until  recently,  research  on  aspartic 
acid  racemization  in  mollusks  had  been 
limited  to  some  Pleistocene  marine  samples 
from  the  West  Coast  of  the  U.  S. 
(Kvenvolden  et  al.,  1979).  But  recent  stud- 
ies on  aspartic  acid  racemization  in  desert 
land  snails  have  turned  up  an  interesting 
pattern:  the  initial  rate  of  racemization  is 
extremely  high;  the  rate  then  slows  down 
progressively  with  increasing  time  or  D/L 
ratio.  This  pattern  has  been  demonstrated 
in  both  a  radiocarbon-dated  series  of  Holo- 
cene  shells  (Goodfriend,  1 99 1 )  as  well  as  in 
heating  experiments  of  modern  shells 
(Goodfriend  and  Meyer,  1991). 

The  very  rapid  initial  rate  of  racemiza- 
tion presents  the  possibility  of  using  aspar- 


Conservation     Analytical    Laboratory, 
Smithsonian  Institution,  Washington,  D.C.  20560 


tic  acid  racemization  as  a  high-resolution 
dating  method  for  young  materials.  This  is 
of  particular  interest  because  radiocarbon 
generally  cannot  be  used  for  dating  of  post- 
1650  A.D.  samples  (see  radiocarbon  cali- 
bration curve  of  Stuiver  and  Pearson  ( 1 986) 
for  this  period).  On  the  other  hand,  evalua- 
tion of  the  precision  of  aspartic  acid  racem- 
ization dating  in  this  time  range  is  difficult, 
because  of  the  unavailability  of  radiocar- 
bon as  an  independent  measure  of  age.  For 
this  reason,  we  turned  to  museum  mollusk 
collections  as  a  source  of  material  of  known 
age  with  which  to  evaluate  the  age  predic- 
tive ability  of  D/L  aspartic  acid  ratios.  In 
addition,  we  examined  aspartic  acid  race- 
mization in  three  samples  of  ostrich  egg- 
shells, to  see  if  the  phenomenon  of  rapid 
initial  racemization  also  occurs  in  this  ma- 
terial. 


Materials  and  Methods 

Seven  samples  of  the  land  snail 
Triodopsis  multilineata  from  Iowa  and 
Kansas  were  obtained  from  the  collection 
of  the  U.  S.  National  Museum  of  Natural 
History.  The  dates  of  collection  of  these 
samples  range  from  1881  to  1949.  Most  of 
these  could  be  seen  to  have  been  collected 
alive  because  of  the  presence  of  some  re- 
mains of  the  animals  inside  the  shells.  In 
addition,  two  modem  samples  (1990  and 
1991)  of  another  Triodopsis  species,  T. 
tridentata,  were  collected  in  the  Chevy 
Chase  district  of  Washington,  D.  C.  Samples 
of  modem  ostrich  eggs  hatched  at  the  Front 
Royal,  Virginia,  breeding  farm  of  the  Na- 
tional Zoological  Park  in  1978  and  at  Dolly 


GEOPHYSICAL  LABORATORY 


173 


Farms  (Vicksburg,  Mississippi)  in  1990 
were  obtained.  A  bead  fashioned  out  of 
ostrich  eggshell,  excavated  from  a  fort  at 
Oudepost,  South  Africa  (occupied  ca.  1652- 
1660  A.D.;  Schrire  et  al.,  1990)  was  also 
obtained  for  analysis.  (The  authors  are  in- 
debted to  R.  Herschler  and  P.  Greenhall  of 
the  U.  S.  National  Museum  of  Natural 
History  for  providing  the  material  of 
Triodopsis  multilineata  used  in  this  study. 
Ostrich  egg  samples  were  kindly  provided 
by  A.  Brooks,  C.  Schrire,  J.  Kokis,  the 
National  Zoological  Park,  and  R.  Shafer  of 
Dolly  Farms.) 

The  periostracum  (the  outer  organic 
layer,  covering  or  partly  covering  the 
mollusk  shells)  was  ground  off  the  shell 
samples  by  abrasive-tipped  bits  using  a 
motorized  hand-held  tool.  Organic  mate- 
rial on  the  surface  of  the  ostrich  egg  pieces 
was  similarly  removed.  Samples  were  then 
subjected  to  a  short  dip  in  dilute  HC1, 
washed  three  times  in  distilled  water,  and 
dried  under  vacuum.  Hydrolysis  and 
derivatization  of  the  amino  acids  to  their  N- 
trifluoracetyl  isopropyl  ester  derivatives 
were  carried  out  as  described  in  Goodfriend 
(1991).  The  D/L  amino  acid  ratios  were 
analyzed  by  gas  chromatography  using  a 
Hewlett-Packard  model  5790.  The  values 
are  reported  as  the  ratio  of  the  areas  of  the 
D  and  L  peaks,  as  calculated  by  a  Hewlett- 
Packard  model  3  3 94 A  integrator.  For  the 
Triodopsis  samples,  analyses  were  based 
on  preparations  made  from  small  pieces  of 
three  individual  shells  (total  weight:  13-19 
mg);  a  piece  from  a  single  shell  was  ana- 
lyzed in  the  case  of  the  T.  tridentata  and 
ostrich  egg  samples.  At  least  two  analyses 
of  each  snail  shell  preparation  were  carried 


out  and  the  mean  of  these  is  reported.  The 
standard  error  of  these  sample  means  aver- 
aged 0.001 1.  The  ostrich  eggshell  samples 
were  analyzed  once. 


Results 

The  Triodopsis  shells  show  a  progres- 
sive increase  in  the  D/L  aspartic  acid  ratio 
with  increasing  age  (Fig.  95).  The  initial 
value  of  0.041  to  0.044  represents  either 
racemization  induced  by  the  preparation 
procedure  or  the  occurrence  of  small 
amounts  of  D-aspartic  acid  in  the  modern 
shells.  From  this  initial  value,  the  D/L  ratio 
increases  up  to  0.093  in  the  110-year-old 
specimens  (collected  in  1881).  Although 
the  modem  specimens  are  of  a  different 
species  than  the  others,  different  mollusk 
species  within  the  same  genus  generally 
show  the  same  rates  of  epimerization  (Lajoie 
et  al.,  1980),  so  it  is  expected  that  the 
modem  T.  tridentata  values  are  represen- 
tative of  those  of  modem  T  multilineata.  A 
simple  linear  regression  of  the  D/L  ratios 
on  the  age  yields  the  equation 


D/L  =  0.000458  (age)  +  0.00431, 

which  indicates  an  average  racemization 
rate  of  l%/22  yr.  (In  this  and  subsequent 
analyses,  a  single  anomalous  sample  (D/L 
of  0. 12  for  a  1949  sample)  was  left  out;  this 
high  value  may  have  been  the  result  of 
heating  of  the  shell  to  extract  the  bodies  or 
to  dry  the  shells  after  extraction  of  the 
bodies.)  The  correlation  coefficient  between 
the  D/L  ratio  and  age  is  0.979.  An  estimate 


174 


CARNEGIE  INSTITUTION 


0.10 


-i — i — i — i — i- 


■  T.  multilineata 
•  T.  tridentata 


1990  1970  1950  1930  1910  1890 

year  A.  D. 

Fig.95.  DA-  aspartic  acid  ratios  in  shell  samples  of 
two  species  of  land  snails  of  the  genus  Triodopsis. 
The  year  of  collection  of  the  samples  is  indicated 
on  the  horizonal  axis. 


of  the  error  of  an  age  predicted  from  the  D/ 
L  ratio  of  a  specimen  was  calculated  from 
the  data  as  the  square  root  of  the  mean 
square  error  of  a  regression  of  age  on  D/L 
ratio.  A  value  of  9.5  yr  was  thus  obtained 
and  indicates  that  the  year  of  collection  of 
an  undated  shell  sample  can  be  estimated 
with  approximately  this  degree  of  precision 
based  on  analysis  of  its  D/L  aspartic  acid 
ratio.  This  estimate  assumes  that  the  scatter 
of  the  points  about  the  regression  line  is 
uniform  over  the  range  of  values  analyzed, 
whereas  it  appears  that  the  scatter  is  greater 
at  higher  ratios,  which  is  as  expected  if  the 
scatter  is  due  to  variation  in  the  average 
racemization  rate  of  different  samples.  Thus 
the  error  of  age  estimations  would  be  lower 
than  the  calculated  value  for  more  recent 
ages  and  higher  for  older  ages. 

The  few  results  available  for  the  ostrich 
egg  samples  (Table  28)  also  suggest  a  high 
aspartic  acid  racemization  rate  in  this  ma- 


terial. A  net  racemization  of  about  0.095 ,  or 
a  rate  of  about  1%  racemization  per  35 
years,  is  indicated  for  the  eggshell  sample 
about  330  years  old. 


Discussion 

The  results  confirm  the  occurrence  of  a 
very  high  rate  of  aspartic  acid  racemization 
in  museum  mollusk  material,  as  expected 
based  on  earlier  studies  of  desert  land  snails. 
Because  the  samples  show  a  regular  pattern 
of  increasing  D/L  ratios  with  increasing 
age,  aspartic  acid  racemization  may  be 
useful  as  a  dating  method  for  materials  on 
a  decadal  time  scale.  For  study  of  museum 
collections,  this  may  have  several  applica- 
tions. In  biogeographical  studies,  it  is  often 
of  importance  to  know  when  a  specimen  or 
set  of  specimens  was  found  at  a  particular 
location,  since  distributions  may  change 
over  short  time  scales.  Aspartic  acid  race- 
mization analysis  could  be  used  to  deter- 
mine the  approximate  time  of  collection  of 
undated  samples  in  museum  collections.  It 
could  also  be  used  to  determine  if  speci- 
mens were  alive  (or  freshly  dead)  when 
collected,  or  whether  they  represent  older, 
dead-collected  material.  Older  records  of 
distributions  could  be  obtained  from  such 
material.  The  method  may  also  be  applied 


Table  28.  D/L  aspartic  acid  ratios  in  some  ostrich 
eggshell  samples. 


Year 


Source 


D/L 


1990  Dolly  Farms  0.044 

1978  Natl.  Zoological  Park       0.057 

ca.  1652-1660    Oudepost,  S.  Afr.  0.129 


GEOPHYSICAL  LABORATORY 


175 


to  dating  of  recent  deposits  in  nature  and 
should  provide  good  time  resolution  for  the 
post- 1650  A.D.  period  not  covered  by  ra- 
diocarbon. Such  applications  require  an  in 
situ  rate  calibration  based  on  independentiy- 
dated  material  of  the  same  species.  An  in 
situ  calibration  is  required  since  different 
museum  collections  or  different  field  sites 
will  differ  in  their  average  temperatures.  In 
museum  collections,  this  calibration  can  be 
obtained  from  other  specimens  of  known 
collection  dates.  In  the  field,  radiocarbon 
dates  from  pre- 1650  A.D.  samples  or  dat- 
ing by  association  with  archeological  arti- 
facts are  the  most  likely  sources  of  cali- 
bration dates.  Possible  problems  with  ap- 
plication of  the  method  to  museum  mate- 
rials may  arise  if  samples  have  been  sub- 
jected to  prolonged  heating  or  high  tem- 
peratures during  processing  or  storage. 

Aspartic  acid  racemization  has  been 
applied  previously  to  dating  of  human  teeth 
(e.g.,  Helfman  and  Bada,  1976;  Ohtani  et 
al.,  1988),  where  it  shows  a  high  rate  of 
racemization  (1%  per  13  yr  in  dentine; 
Helfman  and  Bada,  1976).  However,  this 
high  rate  occurs  at  body  temperature,  or 
about  37°C.  The  projected  rate  at  room 
temperature  (about  2 1  °C)  would  be  14  times 
slower,  or  1%  per  180  yr  (assuming  an 
activation  energy  of  30  kcal/mol).  Thus 
very  rapid  racemization  of  aspartic  acid  in 
young  materials  seems  to  be  limited  to 
biogenic  carbonate  materials,  such  as  mol- 
lusk  shells  and  bird  eggshells;  biogenic 
phosphates,  as  represented  by  teeth,  show  a 
considerably  slower  rate.  One  may  expect 
the  phenomenon  of  very  rapid  initial  race- 
mization of  aspartic  acid  to  be  found  also  in 
other  biogenic  carbonates,  such  as  fora- 


miniferal  tests  and  coral  skeletons. 

Some  possible  modifications  of  ana- 
lytical procedures  could  lead  to  even  higher 
temporal  resolution.  For  example,  it  has 
been  found  that  the  free  amino  acid  fraction 
in  mollusks  is  always  more  highly 
epimerized  than  the  total  amino  acid  frac- 
tion that  is  obtained  from  hydrolysis  (e.g., 
Miller  and  Hare,  1980).  Although  D/L  en- 
antiomer  ratios  have  never  been  measured 
in  the  free  amino  acid  fraction,  it  might  be 
expected  that  this  would  yield  similar  re- 
sults, i.e.,  that  the  DA-  aspartic  acid  ratio 
may  increase  faster  in  the  free  amino  acid 
fraction  than  in  the  total.  It  has  been  sug- 
gested that  the  rapid  initial  rate  of  aspartic 
acid  racemization  may  actually  be  the  re- 
sult of  the  racemization  of  asparagine  rather 
than  aspartic  acid  per  se  (Goodfriend,  1 99 1 ); 
asparagine  is  converted  to  aspartic  acid 
during  hydrolysis,  so  what  is  measured  as 
"aspartic  acid"  is  actually  the  sum  of  the 
aspartic  acid  and  asparagine  originally 
present  in  the  sample.  Development  of 
methods  for  measuring  the  D/L  ratio  of 
asparagine  may  result  in  better  time  resolu- 
tion for  dating  applications. 


References 

Brooks,  A.  S.,  P.  E.  Hare,  J.  E.  Kokis,  G.  H.  Miller, 
R.  D.  Ernst,  and  F.  Wendorf,  Dating  Pleisto- 
cene archeological  sites  by  protein  diagenesis 
in  ostrich  eggshell,  Science,  248, 60-64, 1990. 

Goodfriend,  G.  A.,  Patterns  of  racemization  and 
epimerization  of  amino  acids  in  land  snail 
shells  over  the  course  of  the  Holocene, 
Geochim.  Cosmochim.  Acta,  55,  293-302, 
1991. 

Goodfriend,  G.  A.,  and  V.  R.  Meyer,  A  compara- 
tive study  of  amino  acid  racemization/ 
epimerization  kinetics  in  fossil  and  modern 
mollusk  shells,  Geochim.  Cosmochim.  Acta, 
in  press. 


176 


CARNEGIE  INSTITUTION 


Helfman,  P.  M.,  and  J.  L.  Bada,  Aspartic  acid 
racemisation  in  dentine  as  a  measure  of  age- 
ing, Nature,  262,  279-281,  1976. 

Kvenvolden,  K.  A.,  D. .  Blunt,  andH.  E.  Clifton, 
Amino-acid  racemization  in  Quaternary  shell 
deposits  at  Willapa  Bay,  Washington, 
Geochim.  Cosmochim.  Acta,  43,  1505-1520, 
1979. 

Lajoie,  K.  R.,  J.  F.  Wehmiller,  and  G.  L.  Kennedy, 
Inter-  and  intrageneric  trends  in  the  apparent 
racemization  kinetics  of  amino  acids  in  Qua- 
ternary mollusks,  mBiogeochemistry  of  Amino 
Acids,  P.  E.  Hare,  T.  C  Hoering,  and  K.  King, 
Jr.,  eds.,  John  Wiley  and  Sons,  New  York,  pp. 
305-340,  1980. 

Miller,  G.  H.  and  P.  E.  Hare,  Amino  acid  geochro- 
nology:  integrity  of  the  carbonate  matrix  and 
potential  of  molluscan  fossils,  mBiogeochem- 
istry of  Amino  Acids,  P.  E.  Hare,  T.  C.  Hoering, 
and  K.  King,  Jr.,  eds.,  John  Wiley  and  Sons, 
New  York,  pp.  415-443,  1980. 

Ohtani,  S.,  S.  Kato,  H.  Sugeno,  H.  Sugimoto,  T. 
Marumo,  M.  Yamazaki,  and  K.  Yamamoto,  A 
study  on  the  use  of  the  amino-acid  racemiza- 
tion method  to  estimate  the  ages  of  unidenti- 
fied cadavers  from  their  teeth,  Bull.  Kanagawa 
Dental  College,  16,  11-21,  1988. 

Schrire,  C,  J.  Deetz,  D.  Lubinsky,  and  C. 
Poggenpoel,  The  chronology  of  Oudepost  I, 
Cape,  as  inferred  from  an  analysis  of  clay 
pipes,  J.  Archaeol.  ScL,  17,  269-300,  1990. 

Stuiver,  M.  and  G.  W.  Pearson,  High-precision 
calibration  of  the  radiocarbon  time  scale,  AD 
1950-500  BC,  Radiocarbon,  28,  805-838, 
1986. 

Wehmiller,  J.  F.,  Relative  and  absolute  dating  of 
Quaternary  mollusks  with  amino  acid  racem- 
ization: evaluation,  applications  and  questions, 
in  Quaternary  Dating  Methods,  W.  C. 
Mahaney,  ed.,  Elsevier,  Amsterdam,  pp.  171- 
193,  1984. 


A  Burning  Question:  Differences 

between  Laboratory-Induced 

and  Natural  Di agenesis  in  Ostrich 

Eggshell  Proteins* 

A. S.  Brooks,  RE.  Hare,  J.E.  Kokis,  and 
K.  Durana 

In  earlier  papers  (Brooks  et  ai,  1990; 
Kokis  et  al.,  1990),  we  demonstrated  the 
utility  of  the  D-alloisoleucine/L-isoleucine 
(A/I)  ratio  in  ostrich  eggshell  for  estimat- 
ing the  age  of  archaeological  specimens. 
Of  the  biogenic  carbonates  and  phosphates 
tested  so  far,  ostrich  eggshell  most  nearly 
approximates  a  closed  system  with  little 
loss  of  either  water  or  protein  breakdown 
products  to  the  environment.  Although 
ostrich  eggshell  is  a  common  material  in 
archaeological  sites  located  in  the  arid  and 
semiarid  regions  of  the  Old  World,  human 
activity  and  natural  factors  at  these  sites 
may  produce  anomalous  epimerization  ra- 
tios in  two  ways:  by  heating  (camp fires, 
brush  fires)  and  by  stratigraphic  mixing 
through  ancient  excavations  (pits,  burials, 
burrows,  etc.)  into  underlying  deposits 
which  may  also  contain  eggshell  from  hu- 
man occupation  debris.  For  any  archaeo- 
logical horizon,  it  is  important  to  be  able  to 
distinguish  between  heating  and  strati- 
graphic  admixture,  especially  of  older  ma- 
terials. Heating  to  certain  higher  tempera- 
tures may  indicate  the  presence  of  human- 
controlled  fire,  whereas  stratigraphic  ad- 
mixture may  call  into  question  the  interpre- 


This  work  is  supported  by  NSF  Grant  BNS- 
9011657 


GEOPHYSICAL  LABORATORY 


177 


tation  of  other  materials  at  the  site,  e.g., 
human  fossils.  In  addition,  if  temperature 
differentially  affects  two  decomposition 
reactions  because  they  have  different  acti- 
vation energies,  we  may  be  able  to  use 
these  two  reactions  to  determine  simulta- 
neously both  time  and  temperature.  In  this 
paper,  we  describe  results  from  laboratory 
heating  of  ostrich  eggshell  fragments  at 
controlled  temperatures  for  varying  peri- 
ods. Amino  acid  compositions  of  these 
heated  samples  were  compared  to  amino 
acid  compositions  of  archaeological 
samples  from  the  last  80,000  years. 


Materials  and  Experimental  Methods 

Heating  experiments  were  conducted 
in  a  heated  aluminum  block  with  a  Model 
71 A  Temperature  Controller  (RFL  Indus- 
tries, Inc.,  Boonton,  New  Jersey).  Tem- 
perature readings  were  within  ±0.2° C. 
Samples  of  eggshell  fragments  were 
weighed  and  dropped  into  pre -heated  tubes 
placed  in  the  aluminum  block.  Samples 
were  then  processed  for  free  and  total  amino 
acids  as  described  by  Brooks  et  al.  (1990). 

Three  different  sample  series  of  ostrich 
eggshell  fragments  were  heated  in  the  labo- 
ratory, and  two  sets  of  archeological  samples 
were  analyzed.  (Modem  ostrich  eggshell 
samples  were  provided  by  the  National 
Zoological  Park,  Washington,  D.C.  and  R. 
Shafer  of  Dolly  Farms,  Vicksburg,  Missis- 
sippi. Archaeological  eggshell  samples 
were  provided  by  O.  Bar-Yosef,  A.S. 
Brooks,  J.  Deacon,  M.  Mehlman,  and  W.E. 
Wendt.) 


Laboratory-Heated  Samples: 

(1)  A  series  heated  dry  at  300°C  for 
incremental  time  periods  of  15  min.,  30 
min.,  1  hr.,2hrs.,4hrs.,  8hrs.,  16hrs.,and 
32hrs. 

(2)  A  series  heated  dry  for  one  hour 
each  at  temperatures  of  160°C,  200°C, 
240°C,  280°C,  320°C,  and  360°C. 

(3)  A  series  heated  for  incremental 
time  periods  in  water  vapor  at  157°C  at 
times  ranging  from  0.5  to  256  hours. 
Archaeological  Samples: 

(4)  A  stratified  series  of  archaeological 
samples  from  a  tropical  zone  site  (7uGi, 
Botswana),  in  which  the  A/I  ratio  increases 
regularly  with  age  and  depth  to  above  1.0. 

(5)  A  group  of  pieces  with  anomalous 
A/I  ratios  from  archaeological  sites  (7iGi, 
Boomplaas,  Mumba  Shelter,  Qafzeh, 
Apollo  11).  These  pieces  either  had  no 
significant  A/I  peaks  or  had  ratios  which 
were  much  higher  than  others  from  the 
same  or  underlying  levels. 

For  each  series  we  measured  the  ratios 
of  A/I  peak  areas  subtracting  0.015  from 
each  ratio  to  correct  for  laboratory-induced 
epimerization.  We  also  measured  peak  ar- 
eas of  aspartic  acid  (Asp),  glutamic  acid 
(Glu),  glycine  (Gly),  alanine  (Ala),  and 
ammonia  (NH3),  as  well  as  serine  (Ser), 
threonine  (Thr),  and  arginine  (Arg). 


Results 

In  the  heating  experiments,  a  sequence 
of  changes  in  amino  acid  composition  and 
concentrations  occurs  that  can  be  described 
as  a  series  of  stages. 


178 


CARNEGIE  INSTITUTION 


Stage  0:  Modern  unheated.  Four  major 
amino  acids  (Glu,  Gly,  Asp,  Ala)  occur  in 
comparable  amounts.  Other  amino  acids, 
including  Ser,  Arg,  Thr,  and  He,  are  also 
present  at  lower  levels.  No  significant 
amounts  of  alloisoleucine  are  found.  The 
level  of  NH3  relative  to  the  four  major 
amino  acids  is  insignificant. 

Stage  1.  Light  heating  (1  hour  at  160°- 
200°  C).  The  four  major  amino  acids  persist 
in  comparable  amounts,  whereas  Ser,  Arg, 
and  Thr  diminish  to  trace  levels. 
Alloisoleucine  increases  with  length  of  heat- 
ing. The  NH3  level  is  elevated. 

Stage  2.  Moderate  heating  (1  hour  at 
200°C-280°C).  The  four  major  amino  ac- 
ids no  longer  remain  at  comparable  levels; 
glutamic  remains  relatively  constant  while 
the  others  decrease.  Serine,  threonine,  and 
arginine  diminish  to  only  trace  levels  or  are 
completely  absent.  A/I  values  are  not  al- 
ways possible  to  determine  due  to  the  ap- 
pearance of  interfering  peaks.  The  level  of 
NH3  steadily  increases  from  200°C  to  280°C 
and  over  time. 

Stage  3.  Strong  heating  (1  hour  at 
300°C-360°C).  The  four  major  amino  ac- 
ids diminish  to  only  trace  levels.  Some  Glu 
persists  after  other  amino  acids  disappear. 
NH3  is  the  predominant  peak.  Some  new 
peaks  appear  at  this  stage,  which  are  tenta- 
tively identified  as  y-amino-butyric  acid 
(GABA),  and  some  amines,  possibly 
methyl,  ethyl,  and  propyl.  Alloisoleucine 
and  isoleucine  levels  are  too  low  to  calcu- 
late with  confidence.  Interestingly,  there 
appears  to  be  some  synthesis  of  amino 
acids  at  these  low  levels. 


In  the  archaeological  series  the  same 
four  stages  are  observed.  Samples  that  show 
good  correlations  with  other  age  estimates, 
such  as  radiocarbon  dating,  exhibit  only 
Stage  0  or  1  patterns.  Some  of  the  anoma- 
lous samples  from  archaeological  sites  that 
do  not  correlate  with  the  other  age  esti- 
mates show  stage  2  or  stage  3  patterns,  with 
high  NH3  levels  and  the  presence  of  amines 
and  probably  GABA.  Other  anomalous  ar- 
chaeological samples  exhibit  little  change 
from  stage  0  or  early  stage  1  patterns;  these 
show  no  evidence  of  heating  and  are  pre- 
sumed to  have  derived  from  underlying 
levels  by  stratigraphic  admixture. 


Discussion 

Differences  were  immediately  apparent 
on  inspection  of  the  chromatograms  for  the 
archaeological  series  compared  to  the 
heated  series.  In  the  archaeological  series, 
little  significant  decrease  was  noted  in  the 
four  "stable"  amino  acids  studied,  even 
while  A/I  ratios  increased  to  over  1 .0.  In 
addition,  there  was  little  build-up  of  NH3. 
In  the  heated  samples,  in  contrast,  at  tem- 
peratures as  low  as  200°C  for  1  hour,  or  at 
157°C  for  32  hours,  there  is  a  net  increase 
in  NH3  and  a  decrease  in  aspartic  concen- 
tration relative  to  the  more  stable  concen- 
tration of  glutamic  acid.  The  archaeologi- 
cal samples  from  7cGi  differ  from  all  the 
pieces  from  series  2  heated  at  200°C  or 
above,  all  the  pieces  from  series  1  (300°C), 
and  all  pieces  from  series  3  (157°C)  heated 
32  hours  or  more  in  two  respects:  (1)  heated 


GEOPHYSICAL  LABORATORY 


179 


pieces  exhibit  a  higher  concentration  of 
NH3  than  of  aspartic  acid,  and  (2)  no  ar- 
chaeological piece  has  more  than  50%  as 
much  NH3  as  aspartic  acid,  except  for  the 
anomalous  pieces. 

Of  the  anomalous  archaeological  pieces 
studied,  three  high  A/I  ratio  pieces  from 
7iGi  were  found  in  the  top  Later  Stone  Age 
horizons  where  A/I  ratios  normally  ranged 
from  0.1  to  0.5  and  radiocarbon  calibra- 
tions suggested  an  age  in  the  last  35,000 
years.  Two  of  these  pieces  clearly  fit  with 
the  earlier  Middle  Stone  Age  series  and  are 
presumably  derived  from  below  by  human 
or  animal  disturbance.  However,  one  of  the 
three  which  provided  an  infinite  radiocar- 
bon age  (>40,000  yr  B.R)  must  have  been 
heated  in  antiquity,  as  it  matched  the  amino 
acid  composition  of  the  strongly  heated 
(early  Stage  3)  laboratory  samples.  Two 
anomalous  pieces  from  the  top  levels  at 
another  site  (Boomplaas)  gave  A/I  ratios 
higher  than  the  pieces  from  the  bottom 
levels  whose  estimated  age  was  80,000 
B.R  One  of  these  pieces  has  been  dated  by 
TAMS  14C  to  5220  ±  70  yr  B.R  It  exhibits 
an  NH3-to-aspartic  acid  ratio  greater  than 
1.0,  and  is  presumed  to  have  been  heated. 

Other  anomalous  archaeological  pieces 
contained  high  NH3  concentrations  but 
insufficient  amounts  of  alloisoleucine  or 


isoleucine  to  establish  the  A/I  ratio.  As 
mentioned  above,  experimental  pieces  sub- 
jected to  stronger  heating  conditions  (320°C 
for  one  hour,  or  300° C  for  four  or  more 
hours)  exhibited  high  NH3  and  g-amino- 
butyric  acid  peaks  as  well  as  other  very 
small  peaks.  A/I  ratios  in  these  strongly 
heated  pieces  could  not  be  measured  pre- 
cisely, but  did  not  appear  to  progress  much 
above  0.7.  Examination  of  the  small  peaks 
suggests  that  at  the  higher  temperatures  in 
the  heating  experiments,  serine  as  well  as 
some  other  amino  acids  are  being  synthe- 
sized. Archaeological  samples  from  Apollo 
11  in  Namibia,  Mumba  Shelter  in  Tanza- 
nia, and  Qafzeh  Cave  in  Israel  also  con- 
form to  this  latter  pattern  and  were  almost 
certainly  heated  to  relatively  high  tempera- 
tures in  antiquity. 


References 

Brooks,  A.S.,P.E.  Hare,  J.  E.  Kokis,  G.  H.  Miller, 
R.  D.  Ernst,  andF.  Wendorf,  Dating  pleistocene 
archaeological  sites  by  protein  diagenesis  in 
Ostrich  eggshell,  Science,  248,  60-64,  1990. 

Kokis,  J.  E.,  A.  S.  Brooks,  andP.  E.  Hare,  Chronol- 
ogy a  nd  aminostratigraphy  of  Middle  and  Late 
Stone  Age  sites  from  Sub-saharan  Africa:  A 
comparison  of  protein  diagenesis  and  radio- 
carbon dating  of  ostrich  eggshell,  Geological 
Society  of  America  Abstracts  With  Programs, 
22,  A145-146,  1990, 


GEOPHYSICAL  LABORATORY 


181 


Publications 

Reprints  of  the  numbered  publications  listed  below  are  available,  except  where  noted,  at  no  charge 
from  the  Librarian,  Geophysical  Laboratory,  5251  Broad  Branch  Road,  N.W,  Washington,  D.C. 
20015-1305,  U.S.A.  Please  give  reprint  number(s)  when  ordering.  Youmay  also  request  to  be  placed 
on  the  Laboratory's  mailing  list  to  receive  periodic  notifications  of  recent  publications. 


Angel,  R.  J.,  N.  L.  Ross,  L.  W.  Finger,  and  R.  M. 
Hazen,  Ba3CaCuSi60n:  A  new  {1B,1s(i,oo)} 
{4Si60i7}  chain  silicate,  Acta  Crystallogr. 
C46,  2028-2030,  1990  (G.L.  Paper  2190). 

Angel,  R.  J.,  R.  K.  McMullen,  and  C.  T.  Prewitt, 
Substructure  and  superstructure  of  mullite  by 
neutron  diffraction,  Am.  Mineral.,  76,  332- 
342,  1991  (G.L.  Paper  2216). 

B adding,  J.  V.,  H.  K.  Mao,  and  R.  J.  Hemley, 
High-pressure  synchrotron  X-ray  diffraction 
of  Cs  IV  and  Cs  V,  Solid  State  Commun.,  77, 
801-805,  1991  (G.L.  Paper  2208). 

Bebout,  G.  E.,  Field-based  evidence  for 
devolatilization  in  subduction  zones:  Implica- 
tions for  arc  magmatism,  Science,  251,  413- 
416,  1991  (G.L.  Paper  2206). 

Bebout,  G.  E.,  Geometry  and  mechanisms  of  fluid 
flow  at  15  to  45  kilometer  depths  in  an  early 
cretaceous  accretionary  complex,  Geophys. 
Res.  Lett.,  18,  923-926,  1991  (G.L.  Paper 
2217). 

Chamberlain,  C.  P.,  J.  M.  Ferry,  and  D.  Rumble, 
III,  The  effect  of  net-transfer  reactions  on  the 
isotopic  composition  of  minerals,  Contrib. 
Mineral.  Petrol.,  105,  322-336,  1990  (G.L. 
Paper  2192;  no  reprints  available  for  distribu- 
tion). 

Cifuentes,  L.  A.,  L.  E.  Schemel,  and  J.  H.  Sharp, 
Qualitative  and  numerical  analysis  of  the  ef- 
fects of  river  inflow  variations  on  mixing 
patterns  in  estuaries,  Estuarine  Coastal  Shelf 
Sci.,  30,  41 1-427,  1990  (G.L.  Paper  2198;  no 
reprints  available  for  distribution). 

Coffin,  R.  B.,  D.  J.  Velinsky,  R.  Devereux,  W.  A. 
Price,  and  L.  A.  Cifuentes,  Stable  carbon  iso- 
tope analysis  of  nucleic  acids  to  trace  sources 
of  dissolved  substrates  used  by  estuarine  bac- 


teria, Appl.  Environ.  Microbiol.,  56,  2012- 
2020,  1990  (G.L.  Paper  2191). 

Cohen,  R.  E.,  Bonding  and  elasticity  of  stishovite 
Si02  at  high  pressure:  linearized  augmented 
plane  wave  calculations,  Am.  Mineral.,  76, 
733-742,  1991  (G.  L  Paper  2220). 

Fei,  Y.,  H.  K.  Mao,  and  B.  O.  My  sen,  Experimen- 
tal determination  of  element  partitioning  and 
calculation  of  phase  relations  in  the  MgO- 
FeO-SiC>2  system  at  high  pressure  and  high 
temperature,  /.  Geophys.  Res.,  96,  B2,  2157- 
2169,  1991  (G.L.  Paper  2205). 

Finger,  L.  W.,  R.  M.  Hazen,  and  C.  T.  Prewitt, 
Crystal  structures  of  Mgi2SUOi9(OH)2  (Phase 
B)  and  Mgi4Si5<I>24  (Phase  AnhB),  Amer. 
Mineral,  76,  1-7,  1991  (G.L.  Paper  2219). 

Hare,  P.  E.,  M.  L.  Fogel,  T.  W.  Stafford,  Jr.,  A.  D. 
Mitchell,  and  T.  C.  Hoering,  The  isotopic 
composition  of  carbon  and  nitrogen  in  indi- 
vidual amino  acids  isolated  from  modern  and 
fossil  proteins,  /.  Archaeol.  Sci.,  18, 277-292, 
1991  (G.L.  Paper  2215). 

Hanfland,  M.,  R.  J.  Hemley,  and  H.  K.  Mao, 
Optical  absorption  measurements  of  hydrogen 
at  megabar  pressures,  Phys.  Rev.  B,  43,  8767- 
8770  1991  (G.L.  Paper  2213). 

Hazen,  R.  M.,  Crystal  structures  of  high-tempera- 
ture superconductors,  in  Physical  Properties 
of  High-Temperature  Superconductors  II,  D. 
M.  Ginsberg,  ed.,  Chapter  3,  pp.  121-198, 
World  Scientific,  New  Jersey,  1990  (G.L.  Pa- 
per 2158;  no  reprints  available  for  distribu- 
tion). 

Hazen,  R.  M.,  and  J.  S.  Trefil,  Science  Matters: 
Achieving  Scientific  Literacy,  Doubleday,  New 
York,  1991  (G.L.  Paper  2195)  (Available  at 
your  local  bookstore  or  if  you  prefer  direct 
from  Doubleday) 


182 


CARNEGIE  INSTITUTION 


Hazen,  R.  M.,  and  J.  S.  Trefil,  Achieving  geologi- 
cal literacy,  /.  Geol.  Educ,  39,  28-30,  1991 
(G.L.  Paper  2196) 

Hazen,  R.  M,  J.  Zhang,  and  J.  Ko,  Effects  of  Fe/ 
Mg  on  the  compressibility  of  synthetic 
wadsleyite:P-(Mg1.J^ejt)2Si04(x<Q.25),P/ry5. 
Chem.  Minerals ,17,  416-419,  1990  (G.L.  Pa- 
per 2197). 

Hemley,  R.  J.,  H.  K.  Mao,  L.  W.  Finger,  A.  P. 
Jephcoat,  R.  M.  Hazen,  and  C.  S.  Zha,  Equa- 
tion of  state  of  solid  hydrogen  and  deuterium 
from  single-crystal  X-ray  diffraction  to  26.5 
GPa,  Phys.  Rev.  B,  42, 6458-6470, 1990  (G.L. 
Paper  2180). 

Hemley,  R.  J.,  andH.  K.  Mao,  Critical  behavior  in 
the  hydrogen  insultator-metal  transition,  Sci- 
ence, 249,  391-393,  1990  (G.L.  Paper  2184). 

Hemley,  R.  J.,  H.  K.  Mao,  and  M.  Hanfland, 
Spectroscopic  investigations  of  the  insulator- 
metal  transition  in  solid  hydrogen,  in  Molecu- 
lar Systems  under  High  Pressure  (Proceed- 
ings of  the  II  Archimedes  Workshop  on  Mo- 
lecular Solids  under  Pressure  Catania,  Italy, 
28-31  May  1990)  R.  Pucci,  and  G.  Piccitto, 
eds.,  pp.  223-243,  Elsevier,  New  York,  1991 
(G.L.  Paper  2189). 

Hemley,  R.  J.,  H.  K.  Mao,  and  J.  F.  Shu,  Low- 
frequency  vibrational  dynamics  and  structure 
of  hydrogen  at  megabar  pressures,  Phys.  Rev. 
Lett.,  65, 2670-2673, 1990  (G.L.  Paper  2201). 

Hemley,  R.  J.,  and  J.  D.  Kubicki,  Deep  mantle 
melting,  Nature,  349,  283-284,  1991  (G.L. 
Paper  2209). 

Hemley,  R.  J.,  M.  Hanfland,  andH.  K.  Mao,  High- 
pressure  dielectric  measurements  of  solid  hy- 
drogen to  170  GPa,  Nature,  350,  488-491, 
1991  (G.L.  Paper  2218). 

Kubicki,  J.  D.,  G.  E.  Muncill,  and  A.  C.  Lasaga, 
Chemical  diffusion  in  melts  on  the 
CaMgSi206-CaAl2Si2C>8  join  under  high  pres- 
sures, Geochim.  Cosmochim.  Acta,  54,  2709- 
2715,  1990  (G.L.  Paper  2199). 

Kubicki,  J.  D.,  and  A.  C.  Lasaga,  Molecular  dy- 
namics and  diffusion  in  silicate  melts,  in  Dif- 
fusion, Atomic  Ordering,  andMass  Transport, 
J.  Ganguly,  ed.,  pp.  1-50,  Advances  in  Physi- 
cal Geochemistry  Series,  Springer- Verlag, 
New  York,  1990  (G.L.  Paper  2202;  no  reprints 
available  for  distribution). 


Kubicki,  J.  D.,  and  A.  C.  Lasaga,  Molecular  dy- 
namics simulations  of  pressure  and  tempera- 
ture effects  on  MgSiC>3  and  Mg2Si04  melts 
and  glasses,  Phys.  Chem.  Minerals,  77,  661- 
673,  1991  (G.L.  Paper  2224). 

Kudoh,  Y.,  C.  T.  Prewitt,  L.  W.  Finger,A. 
Darovskikh,  and  E.  Ito,  Effect  of  iron  on  the 
crystal  structure  of  (Mg,Fe)SiC>3  perovskite, 
Geophys.  Res.  Lett.  ,  17,  1481-1484,  1990 
(G.L.  Paper  2179). 

Kushiro,  I.,  and  B.  O.  Mysen,  Experimental  stud- 
ies of  the  system  Mg2Si04-Si02-H2  at  pres- 
sures 10"2-10"10  bar  and  temperatures  to 

o 

1650  C:  Application  to  condensation  and 
vaporization  processes  in  the  primitive  solar 
nebula,  in  Progress  in  Metamorphic  andMag- 
matic  Petrology,  (D.  S.  Korzhinskiy  Memo- 
rial Volume),  L.  L.  Perchuk,  ed.,  Chapter  16, 
pp.  4 1 1  -433 ,  Cambridge  University  Press,  New 
York,  1991  (G.L.  Paper  2214). 

Liu,  X.,  and  C.  T.  Prewitt,  High- temperature  dif- 
fraction study  of  LnCoC>3  perovskites:  A 
high-order  electronic  phase  transition.  J.  Phys. 
Chem.  Solids,  52, 441-448,  1991  (G.L.  Paper 
2204). 

Mao,  H.  K.,  R.  J.  Hemley,  and  M.  Hanfland, 
Infrared  reflectance  measurements  of  the  in- 
sulator-metal transition  in  solid  hydrogen, 
Phys.  Rev.  Lett.,  65,  484-487,  1990  (G.L. 
Paper  2186). 

Mao,  H.  K.,  Y.  Wu,  L.  C.  Chen,  J.  F.  Shu,  and  A. 
P.  Jephcoat,  Static  compression  of  iron  to  300 
GPa  and  Feo.8Nio.2  alloy  to  260  GPa:  Implica- 
tions for  composition  of  the  core,  /.  Geophys. 
Res.,95,  B13,  21,737-21,742,  1990  (G.L.  Pa- 
per 2194). 

Mao,  H.  K.,  R.  J.  Hemley,  Y.  Fei,  J.  F.  Shu,  L.  C 
Chen,  A.  P.  Jephcoat,  Y.  Wu,  and  W.  A. 
Bassett,  Effect  of  pressure,  temperature,  and 
composition  on  lattice  parameters  and  density 
of  (Fe,Mg)Si03-perovskites  to  30  GPa,  /. 
Geophys. Res.,96,B5,S069-S019, 1991  (G.L. 
Paper  2212). 

Mao,  H.  K.,  and  R.  J.  Hemley,  Optical  transitions 
in  diamond  at  ultrahigh  pressures,  Nature, 
351,  721  724,  1991  (G.L.  Paper  2222). 

Mysen,  B.  O.,  Volatiles  in  magmatic  liquids,  in 
Progress  in  Metamorphic  and  Magmatic  Pe- 
trology (D.  S.  Korzhinskiy  Memorial  Vol- 
ume), L.  L.  Perchuk,  ed.,  Chapter  17,  pp.  435- 


GEOPHYSICAL  LABORATORY 


183 


475,  Cambridge  University  Press,  New  York, 
1991  (G.L.  Paper  2172).  " 

My  sen,  B.  O.,  Relationships  between  silicate  melt 
structure  and  petrologic  processes,  Earth-Sci- 
ence Reviews,  27,  281-365,  1990  (G.L.  Paper 
2174). 

Mysen,  B.  O.,  Effect  of  pressure,  temperature,  and 
bulk  composition  on  the  structure  and  species 
distribution  in  depolymerized  alkali  alumino- 
silicate  melts  and  quenched  melts,/.  Geophys. 
Res.,  95,  BIO,  15,733-15,744, 1990  (G.L.  Pa- 
per 2181). 

Mysen,  B.  O.,  Interaction  between  water  and  melt 
in  the  system  CaAl2-04-Si02-H20,  Chem. 
Geol,  88,  223-243,  1990  (G.L.  Paper  2188). 

Nagahara,  H. ,  I.  Kushiro,  and  B .  O.  Mysen,  Vapor- 
ization and  condensation  experiments  in  the 
system  olivine-hydrogen,  in  Dynamic  Pro- 
cesses of  Material  Transport  and  Transforma- 
tion in  the  Interior  of  the  Earth,  S.  Marumo, 
ed.,  pp.  473-490,  Terra  Pub.,  Tokyo,  1990 
(G.L.  Paper  2225;  no  reprints  available  for 
distribution). 

Parise,  J.  B.,  Y.  Wang,  A.  Yeganeh-Haeri,  D.  E. 
Cox  and  Y.  Fei,  Crystal  structure  and  thermal 
expansion  of  (Mg,Fe)Si03  perovskite, 
Geophys. Res. Lett., 17,2089-2092,\990(G.L. 
Paper  2207;  no  reprints  available  for  distribu- 
tion). 

Pickett,  W.  E.,  R.  E.  Cohen,  and  H.  Krakauer, 
Lattice  instabilities,  isotope  effect,  and  high- 
Tc  superconductivity  in  La2-xBaxCu04)JP/ry1y. 
Rev.  Lett.,  67,  228-231,  1991. 

Ross,  N.  L.,  and  R.  M.  Hazen,  High-pressure 
crystal  chemistry  of  MgSi03  perovskite,  Phys. 
Chem.  Minerals,  17, 228-237,  1990  (G.L.  Pa- 
per 2176). 

Ross,  N.  L.,  J.  F.  Shu,  R.  M.  Hazen,  and  T. 
Gasparik,  High-pressure  crystal  chemistry  of 
stishovite,  Am.  Mineral.,75,  739-747,  1990 
(G.L.  Paper  2185). 

Ross,  N.  L.,  and  K.  Leinenweber,  Single  crystal 
structure  refinement  of  high-pressure  ZnGe03 
ilmenite,  Z.  Kristallogr.,  191,  93-104,  1990 
(G.L.  Paper  2203;  no  reprints  available  for 
distribution). 

Stafford,  T.  W.,  Jr.,  P.  E.  Hare,  L.  Currie,  A.  J.  T 
Jull,  and  D.  Donahue,  Accuracy  of  North 
American  human  skeleton  ages,  Quaternary 
Research,  34,  111-120,  1990  (G.L.  Paper 


2175). 

Stafford,  T.  W.,  Jr.,  P.  E.  Hare,  L.  Currie,  A.  J.  T. 
Jull,  andD.  J.  Donahue,  Accelerator  radiocar- 
bon dating  at  the  molecular  level,  /.  Archaeol. 
Sci.,  18,  35-72,  1991  (G.L.  Paper  2193). 

Stathoplos,  L.,  and  P.  E.  Hare,  Amino  acids  in 
planktonic  foraminifera:  Are  they  phyloge- 
netically  useful?  in  Origin,  Evolution,  and 
Modern  Aspects  ofBiomineralization  in  Plants 
and  Animals,  Proceedings  of  the  Fifth  Inter- 
national Symposium  on  Biomineralization,  R. 
E.Crick,ed.,pp.329-338,PlenumPubl.Corp., 
New  York,  1989  (G.L.  Paper  2166;  no  reprints 
available  for  distribution). 

Velinsky,  D.  J.,  M.  L.  Fogel,  J.  F.  Todd,  and  B.  M. 
Tebo,  Isotopic  fractionation  of  dissolved  am- 
monium at  the  oxygen-hydrogen  sulfide  inter- 
face in  anoxic  waters,  Geophys.  Res.  Lett.,  18, 
649-652,  1991  (G.L.  Paper  2211). 

Yoder,  H.  S.,  Jr.,  Heat  transfer  during  partial 
melting:  An  experimental  study  of  a  simple 
binary  silicate  system,/.  Volcanol.  Geotherm. 
Res.  43,  1-36,  1990  (G.L.  Paper  2182). 

Zeitler,  P.  K.,  B.  Barreiro,  C.  P.  Chamberlain,  and 
D.  Rumble,  IE,  Ion-microprobe  dating  of  zir- 
con from  quartz- graphite  veins  at  the  Bristol, 
New  Hampshire,  metamorphic  hot  spot,  Geol- 
ogy, 18,  626-629,  1990  (G.L.  Paper  2187;  no 
reprints  available  for  distribution). 

Zhang,  J.,  D.  Ye,  and  C.  T.  Prewitt,  Relationship 
between  the  unit-cell  volumes  and  cation  radii 
of  isostructural  compounds  and  the  additivity 
of  the  molecular  volumes  of  carbonates,  Am. 
Mineral.,76, 100-105, 1991  (G.L.  Paper  22 10). 

Zheng,  Z  Z.,  D.  X.  Gu,  Y,  Xin,  D.  O.  Pederson,  L. 
W.  Finger,  C  G.  Hadidiacos,  andR.  M.  Hazen, 
A  new  1212-type  phase:  Cr-substituted 
TlSr2CaCu207  with  Tc  up  to  about  110  K, 
Modern  Phys.  Lett.,  5,  635-642,  1991  (G.  L. 
Paper  2223;  no  reprints  available  for  distribu- 
tion). 


GEOPHYSICAL  LABORATORY 


185 


Personnel 

July  1, 1990  to  June  30, 1991 


Research  Staff 

Charles  T.  Prewitt,  Director 
Francis  R.  Boyd,  Jr. 
Ronald  E.  Cohen1 
Larry  W.  Finger 
Marilyn  L.  Fogel 
John  D.  Frantz 
P.  Edgar  Hare 
Robert  M.  Hazen 
Russell  J.  Hemley 
Thomas  C.  Hoering 
T.  Neil  Irvine 
Ho-Kwang  Mao 
Bjorn  O.  My  sen 
Douglas  Rumble  III 
David  Virgo 
HattenS.  Yoder,  Jr. 

Postdoctoral  Associates 

Zhaoxin  Gong2 
David  Palmer3 
Ellen  K.  Wright4 

Research  Associates 

Jingzhu  Hu 
Jinfu  Shu 

Postdoctoral  Fellows 

John  V.  B  adding5 
Gray  E.  Bebout 
James  Brenan6 
Yingwei  Fei7 


Michael  Hanfland 
David  B.  Joyce8 
Paul  L.  Koch9 
James  D.  Kubicki10 
Charles  Meade11 
Craig  M.  Schiffries12 
Hiroko  Takahashi13 
Willem  L.  Vos14 
Jinmin  Zhang15 

Predoctoral  Associate 

Julie  Kokis16 

Research  Interns 

Craig  Bates17 
Jon  Cramer18 
Karen  Durana19 
Howard  Lu20 
Alistaire  M.  Moore21 
Nicole  Y.  Morgan22 

Supporting  Staff 

Andrew  J.  Antoszyk,  Shop  Foreman 
Bobbie  L.  Brown,  Instrument  Maker 
Stephen  D.  Coley,  Sr.,  Instrument  Maker 
David  J.  George,  Electronics  Technician 
Christos  G.  Hadidiacos, 

Electronics  Engineer 
Marjorie  E.  Imlay,  Assistant  to  the  Director 
Lavonne  Lela,  Librarian23 
Yunye  Luo,  Library  Technician24 
Harvey  J.  Lutz, 

Technician/Mail  Supervisor 


186 


CARNEGIE  INSTITUTION 


Mary  M.  Moore, 

Word  Processor  Operator 

— Receptionist 
Lawrence  B.  Patrick, 

Maintenance  Supervisor25 
David  Ratliff,  Jr., 

Maintenance  Technician26 
Pedro  J.  Roa, 

Maintenance  Technician27 
Susan  A.  Schmidt, 

Coordinating  Secretary 
John  M.  Straub, 

Business  Manager 
Mark  Vergnetti, 

Instrument  Maker28 
Stephanie  Vogelpohl, 

Administrative  Assistant29 


J.  Michael  Palin,  Yale  University 
Nicolai  P.  Pokhilenko,  Inst.  Mineralogy  & 

Petrology,  Novosibirsk,  USSR 
Robert  Popp,  Texas  A.  and  M. 
Guoyin  Shen, 

University  of  Uppsala,  Sweden 
Bradley  Tebo, 

Scripps  Institution  of  Oceanography 
Noreen  C.  Tuross,  Smithsonian  Institution 
K.  Vedam,  Pennsylvania  State  University 
David  von  Endt,  Smithsonian  Institution 
YanWu, 

University  of  California,  Berkeley 

Adjunct  Senior  Research  Scientist 
Peter  M.  Bell 


Visiting  Investigators 


Emeritus 


Rateb  M.  Abu-Eid, 

Kuwait  Institute  for  Scientific  Research 
Constance  Bertka, 

Arizona  State  University 
Alison  Brooks, 

George  Washington  University 
Robert  T.  Downs, 

Virginia  Polytechnic  Institute 

&  State  University 
Glenn  A.  Goodfriend, 

Weizmann  Institute  of  Science,  Israel 
Matthew  Hoch,  University  of  Delaware 
Hans  G.  Huckenholz, 

Munich  University,  Germnay 
Donald  G.  Isaak, 

Naval  Research  Laboratory 
James  G.  Kirklin, 

Johns  Hopkins  University 
Kevin  Mandernack, 

Scripps  Institution  of  Oceanography 


Hatten  S.  Yoder,  Jr.,  Director  Emeritus 
Felix  Chayes,  Petrologist  Emeritus 


Appointed  Sept.  1,  199U 

2  Appointed  Jan.  28,  1991 

3  Appointed  Oct.  1,  1990 

4  To  June  30,  1991 

5  Accepted  position  as  Assistant  Professor,  The 
Pennsylvania  State  University 

6  Appointed  Nov.  15,  1990 

7  Accepted  position  as  Associate  Staff  Member, 
Geophysical  Laboratory 

8  To  June  30,  1991 

9  Appointed  Sept.  1,  1990 
°To  December  30,  1990 

1  Appointed  Gilbert  Fellow  July  1,  1990 
2To  Sept.  30,  1990 

3  Appointed  March  1,  1991 

4  Appointed  April  1,  1991 
5To  June  30,  1991 
Appointed  July  1,  1990 
7FromJune24,  1991 
8FromMay  15,  1991 
9FromJune  1,  1991 

20From  June  24,  1991 
21  From  June  24,  1991 
22From  June  24,  1991 


GEOPHYSICAL  LABORATORY  187 

23  Also  associated  with  the  Department  of 
Terrestrial  Magnetism  (DTM) 

24  Also  associated  with  DTM 

25  Also  associated  with  DTM 

26  Also  associated  with  DTM 

27  Also  associated  with  DTM 

28  To  Nov.  30,  1990 

29AppointedJune3,  1991