Maïté Van Rampelbergh

146
Faculteit Wetenschappen Analytical, Environmental & Geo-Chemistry AMGC Speleothems as tools to reconstruct paleoclimates in tem- perate (Belgium) and semi-arid (Socotra, Yemen) regions during the Mid- to Late Holocene Ph.D. dissertation presented to obtain the degree of Doctor of Science from the Vrije Uni- versiteit Brussel by Maïté Van Rampelbergh Promotor: Prof. Dr. Philippe Claeys Prof. Dr. Eddy Keppens Co-Promotor: Dr. Sophie Verheyden November 2014

Transcript of Maïté Van Rampelbergh

Page 1: Maïté Van Rampelbergh

Faculteit WetenschappenAnalytical, Environmental & Geo-ChemistryAMGC

Speleothems as tools to reconstruct paleoclimates in tem-

perate (Belgium) and semi-arid (Socotra, Yemen) regions

during the Mid- to Late Holocene

Ph.D. dissertation presented to obtain the degree of Doctor of Science from the Vrije Uni-versiteit Brussel by

Maïté Van Rampelbergh

Promotor: Prof. Dr. Philippe ClaeysProf. Dr. Eddy Keppens

Co-Promotor: Dr. Sophie Verheyden

November 2014

Page 2: Maïté Van Rampelbergh
Page 3: Maïté Van Rampelbergh

  i  

Acknowledgements    

This  thesis  is  the  result  of  four  years  of  reading,   learning,  travelling,  doing  field  work,   meeting   great   people   and   working   hard,   which   would   not   have   been  possible  without  the  great  team  that  surrounded  me.      First  of  all  I  want  to  thank  my  two  amazing  advisors,  Prof.  Dr.  Eddy  Keppens  and  Prof.  Dr.  Philippe  Claeys,  who,  from  the  first  day  we  met  at  the  VUB,  triggered  my  passion   and   huge   interest   in   Geology   and   past   climates.   The   scientist   and  Geologist  I  have  become  today  is  the  result  of  their  education  filled  with  passion  and  patience.  Thank  you  for  believing  in  me  and  giving  me  the  chance  to  do  this  PhD.  You  both  put  a  lot  of  time  and  effort  in  this  project  and  I  really  appreciated  the  way  you  supported  me.  Thank  you  also   for   feeding  me  with   the  occasional  cheesecakes   (Philippe)   and   home   made   cakes   (Eddy),   which   softened   the  rougher   scientific   periods.   I   really   enjoyed  working  with   both   of   you   and  will  certainly  miss  it.    Second   I   would   like   to   thank   Dr.   Sophie   Verheyden,   the   co-­‐advisor   and   brain  behind  this  whole  project.  Sophie,  without  you  I  could  not  have  brought  this  PhD  to   this   level.   Thank   you   for   teaching   me   how   to   reconstruct   climate   from  speleothems   with   such   a   huge   passion   and   motivation.   I   cannot   thank   you  enough   for   the   many   hours   you   have   put   in   discussing   my   results   and  proofreading  my  papers.  I  really  enjoyed  working  with  you.    I  would   like   to   say   a   big   ‘thank   you’   to  Mr.   Van  Dierendonck,   not   only   for   the  financial  support,  but  also  for  his  enormous  interest  in  my  work  and  his  friendly  visits  at  our  department.    Thank  you   to  Michael  Korntheuer,  David  Verstraeten,  Claire  Mourgues  and  Luc  Deriemaker   for   the  help   and   the  measurements   in   the   lab   and   for   keeping   the  infamous  “Kiel  Device”  running.  You  always  made  time  for  me  when  needed  and  were  always  ready  to  answer  my  questions.      My   thanks   also   go   to   the   ‘Domaine   des   Grottes   de   Han’   and   Guy   Evrard   for  allowing  me  to  core  the  Proserpine  and  letting  me  sample  the  drip  site  every  two  weeks.  A  special   thank  you  goes  to  Etienne  Lannoy,  guide  at   the  Han-­‐sur-­‐Lesse  cave,  who  during   the  4  years  of  my  project,   sampled  rainwater  and  drip  water  every  two  weeks  and  always  showed  a  big  interest  in  my  work.  It  was  always  a  pleasure  to  come  and  pick  up  my  samples  at  your  house.      

Page 4: Maïté Van Rampelbergh

  ii  

 Many  thanks  to  my  colleagues  for  the  support,  advice  and  the  good  times  during  the   coffee   breaks   and   BBQ’s:   Dr.   Steven   Goderis,   Dr.   Virginie   Renson,   Kevin  Debondt,  David  De  Vleeschouwer,  Joke  Belza,  Christina  Makarona,  Aurélie  Sorel,  Janos   Kodolanyi,   Claudio   Ventura   Bordenca,   Matthew   Hubert,   Lidia   Pittarello,  Sean  McKibbin,  Harry  Zekollari,  Niels  De  Winter  and  many  others.  I  will  miss  you  guys!    I  would  like  to  thank  my  parents  for  giving  me  the  opportunity  to  study  Geology  and  for  always  supporting  me  in  my  decisions.  I  could  not  have  reached  this  far  without  your  love.  Also  a  big  ‘thank  you’  to  my  great  family  and  friends  for  their  encouragements  and  drinks  when  I  needed  them  the  most:  Maud,  Vincent,  Ellen,  Tim,   Hanneke,   Philippe,   Yannick,   Nick,   Marc,   Aline,   Lana,   Kim,   Eyra,   Wouter,  Thomas,  Bart,  Daan,  Arnaud,  Monia,  Wiet,  Justine,  Laurens  and  many  others.    Finally   I   want   to   thank   Robbert,   the   man   and   recent   husband   of   my   life   and  dreams.   This   PhD   is   the   results   of   your   never-­‐ending   support   and   love.   You  always   believed   in   me   and   filled   me   with   the   confidence   needed   to   proceed.  Thank  you  for  sharing  my  joy  during  the  happy  moments  but  also  for  getting  me  through  the  harder  times,  I  could  not  have  done  this  without  you.  The  way  you  supported  me   the   last  months   really   impressed  me,   I  know   I  am  safe  with  you  and  that  you  will  keep  making  me  laugh  even  in  the  hardest  times.  I  love  you.                                            

Page 5: Maïté Van Rampelbergh

  iii  

 

Table  of  contents          Acknowledgements                   i  Summary                                              v  Samenvatting                                              vii  Chapter  1  -­‐  General  Introduction                  1  Chapter  2  -­‐  Speleothems  and  Climate              11  2.1  Speleothems                    11  2.2  Dating  of  speleothems                  14  

2.2.1  Dating  techniques                15  2.2.2  Establishing  an  age-­‐depth  model            18  

2.3  δ18O  and  δ13C  in  Speleothems                19  2.3.1  Factors  that  determine  the  δ18O  of  the  drip  water        19  2.3.2  Factors  that  determine  the  δ13C  of  the  drip  water        22  2.3.3  Equilibrium  fractionation  factors              24  2.3.4  Disequilibrium  fractionation,  new  insights          26  2.3.5  Disequilibrium  fractionation,  effects  and  tests        30  2.3.6  Prior  Calcite  Precipitation  (=  PCP)            32  2.3.7  Summary  δ18O  and  δ13C  signals  in  speleothems        33  

2.4  Mg  and  Sr  in  speleothems                36  2.4.1  The  Mg  and  Sr  distribution  coefficients            36    2.4.2  The  Mg/Ca  and  Sr/Ca  ratio  in  the  precipitating  solution      39  

2.5  Climate  reconstructions  from  speleothems              42  2.5.1  Quantitative  paleotemperature  estimates            42  2.5.2  Semi-­‐empirical  climate  relationships            44  

  2.5.3  Quantification  of  the  isotope  effects  in  the  meteorological                            cycle                    45  

2.5.4  Speleothems  as  tools  to  reconstruct  continental  climates                                    46  Chapter  3  -­‐  Socotran  speleothems  reveal  monsoon  changes  during                                                    the  Mid-­‐  to  Late-­‐Holocene              65  Chapter  4  -­‐  Monitoring  of  the  Proserpine  stalagmite          83  Chapter  5  -­‐  A  500-­‐year  speleothem  multiproxy  record  from            

                   the  Han-­‐sur-­‐Lesse  cave,  Belgium                                    103  Chapter  6  -­‐  General  Conclusions  and  Perspectives                                  131      

Page 6: Maïté Van Rampelbergh

 iv  

 

Page 7: Maïté Van Rampelbergh

  v  

Summary    In  the  past  60  years,  speleothems  have  successfully  contributed  to  paleoclimate  reconstruction   and   provided   important   insights   in   climate   teleconnections.  Although   large-­‐amplitude   long-­‐term   global   variations   are   well   documented,  better  insights  in  the  smaller-­‐scaled  and  shorter-­‐term  variations  such  as  during  the   Holocene   and   the   last   millennium   are   still   needed.   To   allow   such  reconstructions,  the  behavior  of  proxy  signals  measured  in  speleothems  needs  to  be   understood   in  more   details.   In   this   PhD,   speleothems   from   two   contrasting  climate   regions,   being   temperate   (Belgium)   and   semi-­‐arid   (Socotra,   Yemen)  systems,  are  studied  for  multiple  proxies  at  high  resolution  going  up  to  seasonal  scales   to   investigate   their   potential   as   tool   to   reconstruct   climate   variations  during   respectively,   the  most   recent   500   years   and   the  Mid-­‐   to   Late  Holocene.  Speleothems  from  Socotra  Island,  located  in  the  northern  Indian  Ocean,  indicate  detailed   variations   of   the   less-­‐known   winter   subsystem   of   the   Indian   Ocean  Monsoon  over  the  last  6  000  years.  Monitoring  results  of  a  fast  growing  (up  to  2  mm/y)   stalagmite   (called   ‘Proserpine’)   in   the   Han-­‐sur-­‐Lesse   cave   in   Belgium  provide   new   insights   on   how   δ18O,   δ13C,   growth   rate   and   calcite   fabric   link   to  climate  variations  at  a  seasonal  scale.  Applying  these  findings  on  the  Proserpine  speleothem  allows  reconstructing  climate  variations  during  the  last  500  years  up  to  seasonal  scale  in  terms  of  wetter  and  warmer  or  colder  and  dryer  winters.    Chapter   1   highlights   the   important   position   of   speleothems   within   the   broad  climate   science   area   and   describes   the   need   for  more   detailed   high-­‐resolution  reconstructions   to   allow   refining   the   climate   models   to   more   regional   scales.  Chapter  2  provides  an  elaborated  state-­‐of-­‐the  art  of  the  processes  and  the  theory  behind  speleothem  based  climate  reconstructions.  At  the  end  of  this  chapter,  the  history  of  speleothem  science  with  the  milestone  papers  are  discussed  together  with   important  recent  methodological   insights,  which  will   further  reinforce  the  position  of  speleothems  within  the  field  of  climate  science.  Chapter  3  reports  the  first  part  of  this  PhD  research  that  focuses  on  using  speleothems  to  reconstruct  monsoon   variations   in   the   northern   Indian   Ocean.   Four   speleothems   from   the  eastern   side   of   Socotra   Island   (Yemen)   are   studied   for   their  δ18O,  δ13C,  Mg/Ca  and   Sr/Ca   composition   to   investigate   how   the   Indian   Ocean   Monsoon   (IOM)  evolved  in  that  area  over  the  last  6  000  years.  The  4  records  provide  unique  new  insights  in  the  northeast  winter  IOM  subsystem,  which  is  difficultly  recorded  in  archives   since   it   is   often   overwritten   by   the   stronger   summer   IOM   variations.  The   results   show   that   the   northeast  winter   IOM   subsystem   evolves   differently  than  its  southwest  summer  counterpart  over  the  last  6  000  years  with  no  links  to  the  North   Atlantic   climate   records.  Chapter  4   and  Chapter  5   report   the   second  part   of   this   PhD   research   where   the   Proserpine   stalagmite   from   the   Han-­‐sur-­‐

Page 8: Maïté Van Rampelbergh

 vi  

Lesse   cave   is   studied   to   investigate   how   speleothem   proxies   record   climate  variations  at  seasonal  resolution  in  temperate  climates  such  as  Belgium.  First,  a  one   year   (=   2013),   biweekly   cave  monitoring   of   the   Proserpine   growth   site   is  carried  out  in  Chapter  4  to  investigate  how  δ18O,  δ13C,  layer  thickness  and  calcite  fabric  changes  can  be  used  to  reconstruct  the  paleoclimate  at  seasonal  scale.  The  most  important  conclusions  of  this  work  are  that  seasonal  climate  variations,  i.e.  seasonal   variations   in   the   climate   parameters   (in   this   thesis,   effective  precipitation   and   temperature),   are   quickly   transferred   into   the   cave   and   that  they  are  successfully  recorded  in  the  speleothem  calcite.  Seasonal  δ18O  variations  reflect   temperature   variations,   whereas   seasonal   δ13C,   layer   thickness   and  growth  rate  variations  reflect  changes  in  effective  precipitation.  These  acquired  insights  are  used   in  Chapter  5   to   interpret   seasonally   resolved  δ18O,  δ13C,   layer  thickness   and   calcite   fabric   changes   in   the   Proserpine   stalagmite   over   the   last  500  years   to   investigate  climate  variations  during   the   last  part  of   the  Little   Ice  Age  (LIA,  ±  1300-­‐1850)  and  the  anthropogenic  period  (most  recent  150  years).  Decadal  and  multi-­‐decadal  changes  in  the  measured  proxies  reconstruct  winter  precipitation   intensities   (and   temperatures)   and   indicate   the   occurrence   of  different   climatic   events   that   correspond   with   known   European   variations.  Seasonal   variations   are   well   recorded   in   the   stalagmite   and   provide   new  knowledge  on  seasonal  temperature  and  effective  precipitation  over  the  studied  period   in   Northern   Continental   Europe.   This   work   illustrates   that,   provided   a  good   knowledge   of   the   cave   system   combined   with   a   multiproxy   approach;  speleothems   from   temperate   climates   successfully   record   lower-­‐amplitude  and  seasonal  variations.            

Page 9: Maïté Van Rampelbergh

  vii  

Samenvatting    Tijdens  de   jongste  60   jaar  hebben  speleothemen  bewezen  succesvol   te  kunnen  worden  ingezet  bij  het  bestuderen  van  klimaatveranderingen  uit  het  verleden  en  hebben   ze   vooral   bijgedragen   tot   het   verwerven   van   nieuwe   inzichten   in   de  teleconnecties   tussen   verschillende   klimaatsystemen.   Hoewel   de  klimaatgemeenschap   reeds   een   grondige   kennis   bezit   op   het   vlak   van   globale,  sterk   uitgesproken   klimaatveranderingen,   is   er   nog   steeds   nood   aan   het   beter  begrijpen   van   de  minder   uitgesproken   veranderingen   op   regionale   schaal.   Het  bestuderen  van  dergelijke  ‘zwakkere’  klimaatvariaties  op  een  hoge  resolutie  kan  gebeuren  aan  de  hand  van  speleothemen  maar  vereist  een  grondige  kennis  van  de   relatie   tussen   de   veranderingen   in   het   klimaat   en   de   veranderingen   in   de  meetbare  proxies.  In  deze  doctoraatsthesis  wordt,  door  het  meten  van  meerdere  proxies   op   een   hoge   tijdsresolutie   (tot   seizoenaal)   in   speleothemen   van   een  gematigde   (België)   en   een   semi-­‐aride   (Socotra,   Jemen)   klimaatregio,   het  potentieel   van   deze   archieven   geëvalueerd   voor   het   reconstrueren   van  klimaatveranderingen  tijdens  respectievelijk  de  jongste  500  jaar  en  het  midden-­‐  tot   laat-­‐Holoceen.   De   resultaten   van   de   speleothemen   van   het   eiland   Socotra,  gelegen   in   het   noorden   van  de   Indische  Oceaan,   geven   gedetailleerde   variaties  weer   van   het   minder   gekende   winter-­‐subsysteem   van   de   Indische   Oceaan  Moesson  (IOM)  tijdens  de  afgelopen  6  000  jaar.  De  resultaten  van  de  monitoring  van  een  snel  groeiende  (tot  2  mm/jaar)  stalagmiet  (de   ‘Proserpine’),   in  de  grot  van   Han-­‐sur-­‐Lesse   in   België,   bieden   nieuwe   inzichten   over   hoe   δ18O   en   δ13C-­‐signalen,   groeisnelheid  en   calcietstructuur   relateren  met  klimaatveranderingen  tot   op   seizoenale   schaal   in   gematigde   klimaatregio’s.   Het   toepassen   van   deze  resultaten  op  de  proxy-­‐reeksen  gemeten   in  de  Proserpine,  over  de   jongste  500  jaar,   laat  toe  klimaatveranderingen  tot  op  seizoenale  schaal  te  reconstrueren  in  termen  van  veranderingen  in  winterneerslag  en  –temperatuur.    Hoofdstuk  1  schetst  de  belangrijke  positie  van  klimaatreconstructies  aan  de  hand  van  speleothemen  binnen  het  bredere  domein  van    de  klimaatwetenschappen  en  beschrijft   de   behoefte   aan   meer   gedetailleerde   hoge-­‐resolutie  klimaatreconstructies   die   het   verfijnen   van   klimaatmodellen   naar   regionale  schaal  mogelijk  maken.  Hoofdstuk  2  geeft  een  uitgebreide  ‘state-­‐of-­‐the-­‐art’  van  de  theorie   en   van   de   belangrijke   processen   die   spelen   bij   het   gebruik   van  speleothemen   in   het   bestuderen   en   reconstrueren   van   het   klimaat.  Het   laatste  deel   van   dit   hoofdstuk   beschrijft   de   geschiedenis   van   de  speleotheemwetenschap,   de   mijlpaalstudies   en   de   recente   methodologische  ontwikkelingen   die   mits   verdere   ontwikkeling   zullen   bijdragen   tot   betere   en  meer  gedetailleerde  klimaatreconstructies  aan  de  hand  van  speleothemen.    

Page 10: Maïté Van Rampelbergh

 viii  

Hoofdstuk  3  beschrijft  het  eerste  luik  van  dit  doctoraatsonderzoek  dat  zich  richt  op  het  gebruik  van  speleothemen  om  variaties  van  de   IOM  te   reconstrueren   in  het  noordelijk  deel  van  de  Indische  Oceaan.  Vier  speleothemen,  bemonsterd  op  de  oostelijke  kant  van  het  eiland  Socotra  (Jemen)  worden  onderzocht  naar  hun  δ18O,  δ13C,  Mg/Ca  en  Sr/Ca-­‐samenstelling  om  na  te  gaan  hoe  de  IOM,  tijdens  de  afgelopen   6   000   jaar,   is   geëvolueerd   in   dit   gebied.   De   4   stalagmieten   bieden  unieke   nieuwe   inzichten   in   het   noordoostelijk   winter   IOM   subsysteem,   dat  moeilijk   te   bestuderen   is   omdat   het   in   klimaatarchieven   vaak   wordt  overschreven  door  de  sterkere  zomer  IOM  variaties.  De  resultaten  tonen  aan  dat,  over   de   jongste   6   000   jaar,   het   noordoostelijk  winter   IOM   subsysteem   op   een  andere   manier   evolueerde   dan   zijn   zuidwesten   zomer   IOM   tegenhanger.  Bovendien  vertoont  het  noordoosten  winter  IOM  subsysteem  geen  link  met  het  Noord-­‐Atlantisch   klimaat.   Hoofdstuk   4   en   5   omvatten   het   tweede   luik   van   dit  doctoraatsonderzoek  waarin  de  Proserpine  stalagmiet  uit  de  grot  van  Han-­‐sur-­‐Lesse  wordt  bestudeerd  om  te  achterhalen  hoe  speleotheem-­‐proxies  seizoenale  klimaatveranderingen,   i.e.   seizoenale   veranderingen   van   klimaatparameters   (=  in   deze   studie   effectieve   neerslag   en   temperatuur),     opnemen   in   gematigde  klimaten   zoals  België.  Hoofdstuk  4   beschrijft   de   eerste   fase   van  dit   luik  waarin  een   tweewekelijkse   monitoring   van   de   Proserpine-­‐groeisite   gedurende   een  periode  van  één  jaar  wordt  uitgevoerd  om  te  onderzoeken  hoe  variaties  in  δ18O,  δ13C,   laagdikte   en   calcietstructuur   veranderen   in   functie   van   de   seizoenen.  Resultaten   tonen   dat   seizoenale   klimaatveranderingen   snel   worden  overgebracht  naar  de  grot  en  dat  ze  op  betrouwbare  wijze  worden  opgenomen  in   het   speleotheemcalciet.   Seizoenale   δ18O   variaties   weerspiegelen  temperatuurschommelingen,   terwijl   seizoenale   δ13C,   laagdikte-­‐   en  groeisnelheidvariaties,   veranderingen   in   effectieve   neerslag   weerspiegelen.   In  Hoofdstuk   5   worden   deze   verworven   inzichten   gebruikt   om   veranderingen   in  δ18O,   δ13C,   laagdikte   en   calcietstructuur,   gemeten   op   seizoenale   schaal,   in   de  Proserpine  over  de  afgelopen  500  jaar  te  interpreteren  en  zo  een  beter  inzicht  te  krijgen  in  seizoenale  klimaatveranderingen  tijdens  het  laatste  deel  van  de  Kleine  IJstijd  (±  1300-­‐  1850)  en  de  antropogene  periode  (laatste  150  jaar).  Tienjarige  en  honderdjarige  variaties   in  de  gemeten  proxies  weerspiegelen  veranderingen   in  neerslagintensiteit   (en   temperatuur)   tijdens   de   winters   en   stemmen   overeen  met  gekende  variaties  in  het  Europese  klimaat.  Seizoenale  klimaatveranderingen  in   de   stalagmiet   verschaffen   nieuwe   inzichten   in   seizoenale   variaties   in  temperatuur  en  effectieve  neerslag  tijdens  de  bestudeerde  periode.  Het  werk  uit  luik  2  van  dit  doctoraat  illustreert  dat  een  goede  kennis  van  het  grotsysteem,  in  combinatie  met   een  multiproxy-­‐aanpak,   het  mogelijk  maakt   om   speleothemen  uit  een  gematigd  klimaat  succesvol  te  gebruiken  bij  de  reconstructie  van  lagere  amplitude  en  seizoenale  klimaatvariaties.  

Page 11: Maïté Van Rampelbergh

Chapter  1:  General  Introduction  

  1  

Chapter  1        

General  Introduction    “The  bad  weather  of  last  month  is  the  consequence  of  climate  change”  a  statement  often  heard  on  television,  read  in  newspapers  or  made  by  friends  and  family.      Are  these  changes  really  due  to  climate  change?  Is  the  extremely  warm  summer  of   2003   in   Europe   the   consequence   of   climate   change   or   is   it   a   normal  consequence   of   the   combination   of   different   climate   factors   such   as   the  North  Atlantic   Oscillation   (NAO),   the   El   Niño   Southern   Oscillation   (ENSO),   solar  insolation,   etc.   Will   the   ongoing   climate   change   cause   a   general   warming   in  Europe,   or   will   it   lead   to   more   extreme   seasons   with   warmer   summers   and  colder  winters?  Or  will  winters  become  warmer  and  wetter?      To  answer  such  questions,  climate  studies  are  carried  out  to  better  understand  the   factors   that   influence   climate   variations   for   different   climate   regions.   The  results  from  such  studies  are  used  to  establish  and  refine  climate  models,  which  are  used  to  simulate  past  but  also  future  climate  parameters  such  as  temperature  or   precipitation   intensity.   Figure   1   illustrates   the   most   recent   temperature  reconstruction  based  on  a  combination  of  such  climate  models  and  is  published  in   the   IPCC’s   Fifth   Assessment   Report   (AR5)   (IPCC,   2013).   The   results   clearly  indicate  the  future  global  temperature  increase  that  is  related  to  the  increase  in  greenhouse  gases  in  the  atmosphere,  a  process  already  suggested  by  Arrhenius  in   1896   (Arrhenius,   1896).   To   optimize   and   refine   such   climate  models,  more  detailed   climate   studies   from   the   continents   are   necessary.   Polar   and   ocean  regions  are  well-­‐covered  through  the  ice  core  (Dansgaard  et  al.,  1993;  Thompson  et   al.,   2013)   and   oceanic   sediment   core   records   (Heinrich,   1988;   Shacklet.N,  1968;   Shackleton   and  Vincent,   1978;   Zachos   et   al.,   2001).   Lake   records   can  be  used  to  reconstruct  continental  climate  evolutions,  but   they  do  not  provide  the  high-­‐resolution   and   good   dating   needed   to   refine   the   climate   models   to   the  yearly  or  seasonal  scale  (except  for  the  varved  lake  sediments  that  allow  yearly  resolved  time  series).  Other  archives  such  as  tree-­‐rings  and  corals  can  deliver  the  needed  high-­‐resolution  climate  data  but  they  are  often  seasonally  biased.    

Page 12: Maïté Van Rampelbergh

Chapter  1:  General  Introduction    

 2  

 Figure   1.   Combined   results   of  multiple   climate  models   for   the   period   between  1950   and   2100   for   the   change   in   global   annual   mean   surface   temperature  relative   to   1986–2005   (adapted   after   IPCC,   2013).   The   results   are   given   for  different  scenarios  or  RPC’s  used  during  the  modeling.    Speleothems  such  as  stalagmites  have  the  huge  potential  to  provide  the  missing  high-­‐resolution   paleoclimate   continental   information   needed   to   refine   the  climate   models   to   regional   scales.   Caves   containing   speleothems   are   found   in  almost  every  continent  (Fig.  2)  delivering  a  good  coverage  of  the  continental  area  to   reconstruct   the   climate.   Through   the   well-­‐established   U-­‐series   dating  technique   (Cheng   et   al.,   2000;   Edwards   et   al.,   1987),   speleothems   can   be   very  precisely  be  dated  back  to  600  kyr  BP.    In  contrast  to  14C-­‐dating,  U/Th-­‐ages  are  absolute   and   do   not   need   to   be   calibrated   and   consequently   deliver   ages  with  small  uncertainty  intervals.  Furthermore,  more  recent  speleothems  that  display  seasonal  or  annual  laminations  can  provide  age  models  based  on  layer  counting  that  delivers  exact  timing  of  the  measured  events  (Fairchild  et  al.,  2001;  Mattey  et  al.,  2008;  Treble  et  al.,  2003;  2005;  Van  Rampelbergh  et  al.,  in  review).      Precisely   dated   speleothems   have   already   played   key   roles   in   dating   large  climatic   transitions   such   as   the  Dansgaard-­‐Oescher   cycles   (Genty   et   al.,   2003),  short-­‐lived  abrupt  millennial  scale  events  such  as  the  8.2  kyr  cold  event  (Cheng  et   al.,   2009b),   glacial   inceptions  and   terminations   (Fig.  3)   (Cheng  et   al.,   2009a;  Fleitmann  et  al.,  2009;  Yuan  et  al.,  2004),    important  geological  events  such  as  the  formation   of   the   Grand   Canyon   (Polyak   et   al.,   2008)   or   delivered   long,  continuous  and  well-­‐dated  records  such  as  the  Devils  Hole  Cave,  Nevada,  record  covering  the  last  560  kyr  (Winograd  and  Ludwig,  1996).    

Page 13: Maïté Van Rampelbergh

Chapter  1:  General  Introduction  

  3  

Figure   2.   Karst   regions   of   the   world   indicated   in   pink   on   the   world   map  (www.circleofblue.org).  12.5  %  of  the  continents  (excluding  Antarctica  and  large  parts   of   Greenland   and   Iceland)   are   covered   by   carbonate-­‐rocks   outcrops   and  indicate   that   speleothems   form   an   important   archive   to   reconstruct   the  terrestrial  climate.      Another   advantage   of   speleothems   as   paleoclimatic   archive   is   that   they   often  display   continuous   growth   over   long   time   intervals   (10   to   100   000   years)  making  it  possible  to  provide  continuous  paleoclimate  reconstructions  covering  different  glacial-­‐interglacial  cycles.  Furthermore,  speleothems  are  physically  and  chemically   robust   due   to   their   relatively   protective   cave   environment.   Any  calcite   alteration   that   may   have   occurred   is   easily   detectable   usually   by   clear  macroscopic   calcite   fabric   changes.   Finally,   with   the   recent   analytical   and  technical  advances,  different  possible  geochemical  or  physical  parameters  such  as  stable  isotopes  of  oxygen  and  carbon,  trace  elements  or   layers  thickness  can  be  measured   at   decadal   to   sub-­‐seasonal   resolution.   These   parameters   contain  information   about   the   climate   above   the   cave   during   their   deposition   such   as  temperature,   rainfall   amount,   type   of   vegetation,   soil   productivity   and   glacier  extend.   A   good   understanding   how   these   parameters   reflect,   through   their  transfer   functions,   climate   variations   at   the   surface   allows   detailed  reconstructions.    Hendy  and  Wilson  suggested  in  1968  that  the  δ18O  composition  of  speleothems  deposited  in  equilibrium  with  their  drip  water  could  be  used  to  reconstruct  the  paleo-­‐temperature.  This  was  the  onset  of  the  growing  success  of  speleothems  in  reconstructing  the  paleoclimate  as  illustrated  by  the  exponential  increase  of  the  amount  of   published  papers   since   then   (see   ISI  Web  of   Science).  However,   the  early  promise  of  reconstructing  temperature  curves  from  the  δ18O  composition,  have  proven  to  be  difficult  to  fulfill  due  to  the  fact  that  most  speleothems  do  not  

Page 14: Maïté Van Rampelbergh

Chapter  1:  General  Introduction    

 4  

reflect   perfect   isotopic   equilibrium   conditions   (Dorale   and  Liu,   2009;   Fairchild  and   Baker,   2012).   However,   even   when   affected   by   disequilibrium   effects,  different   studies   have   proven   that   when   carefully   studied   and   correctly  interpreted   the   signals   can   deliver   important   climate   information   such   as   on  shifts   in   the   intensity  of   the  Monsoons   (Cai  et  al.,  2012;  Fleitmann  et  al.,  2007;  Yuan  et  al.,  2004)  or  variations  of  the  NAO  variation  cycles  (Proctor  et  al.,  2000;  Trouet   et   al.,   2009).   The   advantage   of   the   increasing   amount   of   speleothem  records  is  that  continental  and  inter-­‐continental  comparison  between  the  proxy-­‐time  series  (McDermott  et  al.,  2011)  and  with  other  paleoclimate  archives  (Bar-­‐Matthews  et  al.,  2003;  Genty  et  al.,  2003;  Van  Rampelbergh  et  al.,  2013;  Wang  et  al.,   2001)   become   possible.   Their   increased   temporal   resolution   and   reliable  absolute   chronology   also   allow   the   records   to   be   compared   to   climate   forcing  such  as  solar  insolation  that  is  necessary  to  understand  the  climate  mechanisms  (Neff  et  al.,  2001;  Wang  et  al.,  2008).  With  the  advances  in  analytical  techniques  such  as  laser  absorption  spectroscopy  on  speleothem  fluid  inclusions  (Affolter  et  al.,  2014)  or  the  clumped  isotope  techniques  (Kluge  et  al.,  2008),  the  speleothem  community   is   getting   closer   to   extract   temperature   signals   from   speleothems  such  as  first  suggested  by  Hendy  and  Wilson  in  1968.    Although  speleothems  may  seem  the  perfect  archive  to  reconstruct  the  climate,  the  increasing  amount  of  speleothem  studies  indicates  that  a  variety  of  climatic,  environmental   and   hydrological   parameters   influence   the   geochemical   and  physical   properties   of   speleothems   making   their   interpretation   as   a   simple  climate  proxy  not  straightforward.  More  and  more,   the  speleothem  community  realizes  that  each  cave  has  its  own  unique  geological  and  environmental  setting  that   needs   to   be   well   understood   before   stable   isotopes,   trace   elements,  crystallographic  changes,  layer  thickness  records  or  other  proxies  can  be  used  to  reconstruct  the  climate.  This  clearly  emphasizes  that  a  profound  understanding  is   needed   of   how   the   measured   proxy   in   the   selected   cave   is   reflecting   the  climate.  This  can  be  achieved  with  long-­‐term  monitoring  programs  with  frequent  cave  visits.  However,   such  programs  are   logistically,   technically   and   financially  demanding  making  it  difficult  for  research  groups  to  carry  them  out.  When  cave  monitoring  programs  are  not  feasible  due  to  the  remote  location  of  the  cave  or  due  to  logistic  difficulties  with  installing  the  measurement  material  in  the  field,  a  multiproxy   or   if   possible   a  multiple   stalagmite   analysis   of   the   selected   cave   is  necessary  to  get  a  good  understanding  of  the  cave  system  and  how  to  interpret  the  measured  proxies  in  terms  of  climate.    

Page 15: Maïté Van Rampelbergh

Chapter  1:  General  Introduction  

  5  

 Figure  3.  Comparison  of  different  speleothem  records  with  the  NGRIP  ice  record  summarized   in   Fleitmann   et   al.   (2009).   The   curves   indicate   the   value   of  speleothems  not  only  for  the  absolute  dating  of  transitions,  but  also  to   indicate  how   transitions   or   shorter   climate   events   such   as  Heinrich   events   affected   the  terrestrial  climate.    In  this  PhD  thesis,  speleothems  from  two  different  climatic  settings  are  studied,  the   semi-­‐arid   tropical   region   of   the   northern   Indian   Ocean   and   the   more  temperate   region   of   the   Belgium.   The   first   part   of   the   research   focuses   on  analyzing   four   speleothems   from   the   island  of   Socotra   in  Yemen   to   investigate  how  the   Indian  Ocean  Monsoon  has  evolved   in   the  Holocene.  Because   frequent  cave  visits  were  not  possible  due  to  the  remoteness  of  the  caves,  no  monitoring  program   was   carried   out.   When   deposited   in   equilibrium,   the   δ18O   signal   in  speleothems   from   tropical   and   subtropical   areas,   such   as   Socotra,   is   mostly  influenced  by   the   “amount   effect”,   describing   the   inverse   relationship  between  the   amount   of   precipitation   and   its   oxygen   isotopic   composition   (Dansgaard,  1964;   Rozanski   et   al.,   1992)   facilitating   the   interpretation   of   the   measured  

Page 16: Maïté Van Rampelbergh

Chapter  1:  General  Introduction    

 6  

signals  to  a  climate  signal.  However  to  be  100%  certain  that  the  amount-­‐effect  is  responsible   for   the   variations   in   the   sampled   speleothems,   a   multiproxy  approach  was   done   for   each   sample.   Different   speleothems   covering   the   same  time   periods  were   sampled   in   the   same   cave   to   test   the   reproducibility   of   the  results.   Speleothems   were   sampled   in   two   different   caves   to   obtain   a   good  understanding  of  the  local  climate  evolution  of  Socotra.  The  multi-­‐proxy,  multi-­‐speleothem   and   inter-­‐cave   comparison   leads   to   a   profound   understanding   of  how   the   proxies   reflect   climate.   Comparing   the   obtained   Holocene   climate  variations  derived  from  the  studied  speleothems  with  other  records  in  the  area  such   as   ocean   and   lake   cores   or   speleothem   records   from   Southern   Oman  delivers   a   more   profound   knowledge   of   the   mechanisms   of   the   Indian   Ocean  Monsoon  around  the  Northern  Indian  Ocean.      In  the  second  part  of  this  PhD,  a  cave  monitoring  is  combined  with  a  seasonally  resolved  speleothem  record  from  the  Han-­‐sur-­‐Lesse  cave,  Belgium.  The  problem  with  most  proxies  used  in  speleothems  from  mid-­‐latitudes  is  that  they  cannot  be  interpreted   in   terms  of  a  single  climate  parameter  such  as   temperature  and/or  precipitation  due   to   the  multiple  potential   influences   (e.g.   source,   temperature,  seasonal   changes   in   rainfall,   groundwater   residence   time,   etc.)   on   the  speleothem  geochemistry  and/or  morphology  (Baker  et  al.,  2011).  Therefore  the  approach  to  reconstruct  a  climate  signal   from  a  Belgian  speleothem  is  different  compared   to   the   speleothem   climate   reconstruction   in   the   sub-­‐tropics.   To  investigate   how   the   climate   parameters   such   as   temperature,   precipitation   or  soil   activity   are   linked   to   the   cave   parameters,   a   biweekly   cave-­‐monitoring  program  was  carried  out   for  a  period  of  one  year.   In  a   second  phase,   the  most  recent  500  years  of  a  2  000  year  old  seasonally  laminated  stalagmite,  called  the  Proserpine   stalagmite,   from   the   Han-­‐sur-­‐Lesse   cave   was   analyzed   for   its   δ18O  and   δ13C   composition,   layer   thickness   and   calcite   fabric   changes   at   a   seasonal  scale.  Together  with   the  cave  monitoring  results   the  different  analyzed  proxies  deliver  information  on  how  the  climate  evolved  up  to  seasonal  scale  during  the  most  recent  500  years.      The   results   obtained   from   this   PhD   provide   information   that   can   be   used   to  refine   climate  models   around   the   northern   Indian   Ocean   and   around  Western  Europe.  For   the  monsoon  regions  our  work  showed  that  not  only   the  summer-­‐monsoon  affects  the  continental  regions  around  the  northern  Indian  Ocean,  but  that   a   winter   monsoon   also   has   to   be   taken   into   account   and   that   they   both  evolve   differently   over   time.   The   work   in   Belgium   elaborated   the   general  knowledge  of  how  climate  variations  are  linked  to  variations  in  cave  parameters.  These  findings  can  be  used  to  translate  speleothem  proxies  recovered  from  mid-­‐latitude  locations  to  climate  signals.            

Page 17: Maïté Van Rampelbergh

Chapter  1:  General  Introduction  

  7  

References    Affolter,  S.,  Fleitmann,  D.,  Leuenberger,  M.,  2014.  New  online  method   for  water  isotope   analysis   of   speleothem   fluid   inclusions   using   laser   absorption  spectroscopy  (WS-­‐CRDS).  Climate  of  the  Past  10,  1291-­‐1304.  

Arrhenius,   S.,   1896.   On   the   Influence   of   Carbonic   Acid   in   the   Air   upon   the  Temperature  of  the  Ground  Philosophical  Magazine  and  Journal  of  Science  Series  5,  237-­‐276.  

Baker,   A.,   Wilson,   R.,   Fairchild,   I.J.,   Franke,   J.,   Spoetl,   C.,   Mattey,   D.,   Trouet,   V.,  Fuller,   L.,   2011.   High   resolution   delta   O-­‐18   and   delta   C-­‐13   records   from   an  annually   laminated   Scottish   stalagmite   and   relationship   with   last   millennium  climate.  Glob.  Planet.  Change  79,  303-­‐311.  

Bar-­‐Matthews,  M.,  Ayalon,  A.,  Gilmour,  M.,  Matthews,  A.,  Hawkesworth,  C.J.,  2003.  Sea-­‐land   oxygen   isotopic   relationships   from   planktonic   foraminifera   and  speleothems   in   the   Eastern   Mediterranean   region   and   their   implication   for  paleorainfall  during  interglacial  intervals.  Geochimica  Et  Cosmochimica  Acta  67,  3181-­‐3199.  

Cai,   Y.,   Zhang,   H.,   Cheng,   H.,   An,   Z.,   Edwards,   R.L.,   Wang,   X.,   Tan,   L.,   Liang,   F.,  Wang,   J.,   Kelly,   M.,   2012.   The   Holocene   Indian   monsoon   variability   over   the  southern   Tibetan   Plateau   and   its   teleconnections.   Earth   and   Planetary   Science  Letters  335,  135-­‐144.  

Cheng,  H.,  Edwards,  R.L.,  Broecker,  W.S.,  Denton,  G.H.,  Kong,  X.,  Wang,  Y.,  Zhang,  R.,  Wang,  X.,  2009a.  Ice  Age  Terminations.  Science  326,  248-­‐252.  

Cheng,  H.,  Edwards,  R.L.,  Hoff,   J.,  Gallup,  C.D.,  Richards,  D.A.,  Asmerom,  Y.,  2000.  The  half-­‐lives  of  uranium-­‐234  and  thorium-­‐230.  Chemical  Geology  169,  17-­‐33.  

Cheng,   H.,   Fleitmann,   D.,   Edwards,   R.L.,   Wang,   X.F.,   Cruz,   F.W.,   Auler,   A.S.,  Mangini,   A.,   Wang,   Y.J.,   Kong,   X.G.,   Burns,   S.J.,   Matter,   A.,   2009b.   Timing   and  structure  of  the  8.2  kyr  BP  event  inferred  from  delta  O-­‐18  records  of  stalagmites  from  China,  Oman,  and  Brazil.  Geology  37,  1007-­‐1010.  

Dansgaard,  W.,  1964.  Stable  Isotopes  In  Precipitation.  Tellus  16,  436-­‐468.  

Dansgaard,   W.,   Johnsen,   S.J.,   Clausen,   H.B.,   Dahljensen,   D.,   Gundestrup,   N.S.,  Hammer,   C.U.,   Hvidberg,   C.S.,   Steffensen,   J.P.,   Sveinbjornsdottir,   A.E.,   Jouzel,   J.,  Bond,  G.,   1993.  Evidence   for   general   instability  of  past   climate   from  a  250-­‐Kyr  Ice-­‐Core  record  Nature  364,  218-­‐220.  

Dorale,   J.A.,   Liu,   Z.,   2009.   Limitations   of   the   Hendy   test   criterai   in   judging   the  paleoclimatic  suitability  of  speleothems  and  the  need   for  replication.   Journal  of  Cave  and  Karst  Studies  71,  73-­‐80.  

Edwards,   R.L.,   Chen,   J.H.,   Wasserburg,   G.J.,   1987.   238U-­‐234U-­‐230Th-­‐232Th  systematics  and  the  precise  measurement  of   time  over   the  past  500  000  years.  Earth  and  Planetary  Science  Letters  81,  175-­‐192.  

Page 18: Maïté Van Rampelbergh

Chapter  1:  General  Introduction    

 8  

Fairchild,   I.J.,   Baker,   A.,   2012.   Speleothem   Science:   From   process   to   past  environment.  Blackwell  Publishing  ltd.  

Fairchild,  I.J.,  Baker,  A.,  Borsato,  A.,  Frisia,  S.,  Hinton,  R.W.,  McDermott,  F.,  Tooth,  A.F.,  2001.  Annual   to  sub-­‐annual   resolution  of  multiple   trace-­‐element   trends   in  speleothems.  Journal  of  the  Geological  Society  158,  831-­‐841.  

Fleitmann,  D.,  Burns,  S.J.,  Mangini,  A.,  Mudelsee,  M.,  Kramers,   J.,  Villa,   I.,  Neff,  U.,  Al-­‐Subbary,   A.A.,   Buettner,   A.,   Hippler,   D.,  Matter,   A.,   2007.  Holocene   ITCZ   and  Indian   monsoon   dynamics   recorded   in   stalagmites   from   Oman   and   Yemen  (Socotra).  Quaternary  Science  Reviews  26,  170-­‐188.  

Fleitmann,  D.,  Cheng,  H.,  Badertscher,  S.,  Edwards,  R.L.,  Mudelsee,  M.,  Goektuerk,  O.M.,  Fankhauser,  A.,  Pickering,  R.,  Raible,  C.C.,  Matter,  A.,  Kramers,  J.,  Tuysuz,  O.,  2009.   Timing   and   climatic   impact   of   Greenland   interstadials   recorded   in  stalagmites  from  northern  Turkey.  Geophysical  Research  Letters  36.  

Genty,  D.,  Blamart,  D.,  Ouahdi,  R.,  Gilmour,  M.,  Baker,  A.,   Jouzel,   J.,  Van-­‐Exter,  S.,  2003.   Precise   dating   of   Dansgaard-­‐Oeschger   climate   oscillations   in   western  Europe  from  stalagmite  data.  Nature  421,  833-­‐837.  

Heinrich,  H.,  1988.  Origin  and  consequences  of  cyclic  ice  rafting  in  the  northeast  atlantic-­‐ocean  during  the  past  130  000  years.  .  Quaternary  Research  29,  142-­‐152.  

IPCC,   2013.   IPCC   2013:   Summary   for   policy   makers.,   In:   Stocker,   T.F.,   Qin,   D.,  Plattner,   G.K.,   Tingor,   M.,   Allen,   S.K.,   Boschung,   J.,   Nauels,   A.,   Xia,   Y.,   Bex,   V.,  Midgley,   P.M.   (Eds.),   Climate   Change   2013:   The   Physical   Science   Basis.  Contribution   of   Working   Group   I   to   the   Fift   Assessment   Report   of   the  Intergovernmental   Panel   on   Climate   Change.   Cambridge   University   Press,  Cambridge,  United  Kingdom  and  New  York,  USA.  

Kluge,  T.,  Marx,  T.,   Scholz,  D.,  Niggemann,  S.,  Mangini,  A.,  Aeschbach-­‐Hertig,  W.,  2008.  A  new  tool  for  palaeoclimate  reconstruction:  Noble  gas  temperatures  from  fluid   inclusions   in   speleothems.   Earth   and   Planetary   Science   Letters   269,   407-­‐414.  

Mattey,  D.,   Lowry,  D.,  Duffet,   J.,   Fisher,  R.,  Hodge,   E.,   Frisia,   S.,   2008.  A  53   year  seasonally  resolved  oxygen  and  carbon  isotope  record  from  a  modem  Gibraltar  speleothem:   Reconstructed   drip   water   and   relationship   to   local   precipitation.  Earth  and  Planetary  Science  Letters  269,  80-­‐95.  

McDermott,   F.,   Atkinson,   T.C.,   Fairchild,   I.J.,   Baldini,   L.M.,   Mattey,   D.P.,   2011.   A  first   evaluation   of   the   spatial   gradients   in   delta   O-­‐18   recorded   by   European  Holocene  speleothems.  Glob.  Planet.  Change  79,  275-­‐287.  

Neff,   U.,   Burns,   S.J.,   Mangini,   A.,   Mudelsee,   M.,   Fleitmann,   D.,   Matter,   A.,   2001.  Strong  coherence  between  solar  variability  and  the  monsoon  in  Oman  between  9  and  6  kyr  ago.  Nature  411,  290-­‐293.  

Polyak,   V.,   Hill,   C.,   Asmerom,   Y.,   2008.   Age   and   evolution   of   the   Grand   Canyon  revealed   by   U-­‐Pb   dating   of   water   table-­‐type   speleothems.   Science   319,   1377-­‐1380.  

Page 19: Maïté Van Rampelbergh

Chapter  1:  General  Introduction  

  9  

Proctor,   C.J.,   Baker,   A.,   Barnes,   W.L.,   Gilmour,   R.A.,   2000.   A   thousand   year  speleothem   proxy   record   of   North   Atlantic   climate   from   Scotland.   Climate  Dynamics  16,  815-­‐820.  

Rozanski,   K.,   Araguasaraguas,   L.,   Gonfiantini,   R.,   1992.   Relationship   between  long-­‐term   trends   of   18O   isotope   composition   of   precipitation   and   climate  Science  258,  981-­‐985.  

Shacklet.N,   1968.   Depth   of   pelagic   foraminifera   and   isotopic   changes   in  pleistocene  oceans  Nature  218,  79-­‐&.  

Shackleton,  N.J.,  Vincent,  E.,  1978.  Oxygen  and  carbon   isotope  studies   in   recent  foraminifera  from  Southwest  Indian  Ocean  Marine  Micropaleontology  3,  1-­‐13.  

Thompson,  L.G.,  Mosley-­‐Thompson,  E.,  Davis,  M.E.,  Zagorodnov,  V.S.,  Howat,  I.M.,  Mikhatenko,  V.N.,  Lin,  P.N.,  2013.  Annually  Resolved  Ice  Core  Records  of  Tropical  Climate  Variability  over  the  Past  similar  to  1800  Years.  Science  340,  945-­‐950.  

Treble,  P.,  Shelley,   J.M.G.,  Chappell,   J.,  2003.  Comparison  of  high  resolution  sub-­‐annual   records   of   trace   elements   in   a   modern   (1911-­‐1992)   speleothem   with  instrumental  climate  data  from  southwest  Australia.  Earth  and  Planetary  Science  Letters  216,  141-­‐153.  

Treble,   P.C.,   Chappell,   J.,   Shelley,   J.M.G.,   2005.   Complex   speleothem   growth  processes  revealed  by  trace  element  mapping  and  scanning  electron  microscopy  of  annual  layers.  Geochimica  Et  Cosmochimica  Acta  69,  4855-­‐4863.  

Trouet,   V.,   Esper,   J.,   Graham,   N.E.,   Baker,   A.,   Scourse,   J.D.,   Frank,   D.C.,   2009.  Persistent   Positive   North   Atlantic   Oscillation   Mode   Dominated   the   Medieval  Climate  Anomaly.  Science  324,  78-­‐80.  

Van   Rampelbergh,   M.,   Fleitmann,   D.,   Verheyden,   S.,   Cheng,   H.,   Edwards,   L.,   De  Geest,  P.,  De  Vleeschouwer,  D.,  Burns,  S.J.,  Matter,  A.,  Claeys,  P.,  Keppens,  E.,  2013.  Mid-­‐   to   late   Holocene   Indian   Ocean   Monsoon   variability   recorded   in   four  speleothems   from  Socotra   Island,  Yemen.  Quaternary  Science  Reviews  65,  129-­‐142.  

Van  Rampelbergh,  M.,  Verheyden,  S.,  Allan,  M.,  Quinif,  Y.,  Keppens,  E.,  Cheng,  H.,  Edwards,  L.R.,  Claeys,  P.,  in  review.  A  500-­‐year  seasonally  resolved  δ18O  and  δ13C   and   layer   thickness   record   from   a   speleothem   of   the   Han-­‐sur-­‐Lesse   cave,  Belgium.  Climate  of  the  Past.  

Wang,   Y.,   Cheng,   H.,   Edwards,   R.L.,   Kong,   X.,   Shao,   X.,   Chen,   S.,  Wu,   J.,   Jiang,   X.,  Wang,   X.,   An,   Z.,   2008.   Millennial-­‐   and   orbital-­‐scale   changes   in   the   East   Asian  monsoon  over  the  past  224,000  years.  Nature  451,  1090-­‐1093.  

Wang,  Y.J.,  Cheng,  H.,  Edwards,  R.L.,  An,  Z.S.,  Wu,  J.Y.,  Shen,  C.C.,  Dorale,  J.A.,  2001.  A   high-­‐resolution   absolute-­‐dated   Late   Pleistocene   monsoon   record   from   Hulu  Cave,  China.  Science  294,  2345-­‐2348.  

Page 20: Maïté Van Rampelbergh

Chapter  1:  General  Introduction    

 10  

Winograd,   I.J.,   Ludwig,   K.R.,   1996.   High   resolution   paleotemperature   proxy  record  for  the  last  interglaciation  based  on  Norwegian  speleothems  -­‐  Comment.  Quaternary  Research  45,  102-­‐102.  

Yuan,  D.,  Cheng,  H.,  Edwards,  R.L.,  Dykoski,  C.A.,  Kelly,  M.J.,  Meiliang,  Z.,  Jiaming,  Q.,  Yushi,  L.,  Yongjin,  W.,  Jiangyin,  W.,  Dorale,  J.A.,  Zhisheng,  A.,  Yanjun,  C.,  2004.  Timing,  duration,  and  transitions  of  the  last  interglacial  Asian  monsoon.  Science  304,  575-­‐578578.  

Zachos,  J.,  Pagani,  M.,  Sloan,  L.,  Thomas,  E.,  Billups,  K.,  2001.  Trends,  rhythms,  and  aberrations  in  global  climate  65  Ma  to  present.  Science  292,  686-­‐693.  

 

 

Page 21: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  11  

Chapter  2      

Speleothems  and  Climate    

2.1  Speleothems      Speleothems   such   as   stalagmites,   stalactites   or   flowstones   are   natural   cave  deposits  that  mostly  consist  of  calcium  carbonate  (CaCO3).  The  development  and  growth   of   speleothems   in   karst   regions   have   been   described   extensively   in  several   publications   (Dreybrodt,   1988;   Fairchild   et   al.,   2006;   Lachniet,   2009;  Fairchild  and  Baker,  2012).    When  rainwater  falls  only  a  fraction  will  effectively  infiltrate  into  the  soil  due  to  evaporation   and   plant   uptake.   In   the   soil,   the   higher   pCO2   (typically   0.1   atm)  compared  to  the  water  pCO2  (typically  10-­‐4  atm)  acidifies  the  water  according  to  reaction  [1]  (Fig.  1):    CO2  +  H2O  à  H2CO3           [1]    until  equilibrium  with  respect  to  the  partial  CO2  pressure  in  the  soil  is  attained.      After  passing  through  the  soil  zone,  the  water  enters  the  epikarst.  The  epikarst  is  the  upper  surface  of  the  bedrock  characterized  by  solution  features  in  the  vadose  zone  where  water  may  be  stored  and  mixed  (Williams,  2008).  In  the  epikarst,  the  corrosive   water   comes   in   contact   with   the   carbonate   bedrock   (mostly  marine  limestone)  and  dissolves   it   to  become  enriched   in  calcium  carbonate  according  to  reaction  [2]  (Fig.  1):    CaCO3  +  H2CO3  à  Ca2+  +  2HCO3-­‐                                 [2]            

Page 22: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 12  

This  zone  is  referred  as  the  ‘dissolution’  zone  of  the  epikarst  (Fig.  1)  (Fairchild  et  al.,  2006).  The  amount  of  dissolved  calcite  depends  on  the  amount  of  dissolved  CO2  in  the  solution  (the  acidity  of  the  water)  and  whether  dissolution  occurs  in  an  open  or  closed  system  (see  2.3.2)  (Dreybrodt,  1988).    

Figure   1.   The   dissolution   and   precipitation   regimes   of   the   karst   environment.  Water  gets  enriched  in  CO2  in  the  soil  zone,  becomes  more  acidic  and  dissolves  the   calcite   bedrock   (dissolution   zone).   In   the   cave,   the   lower   cave   air   pCO2  compared  to  the  water  pCO2  forces  the  drip  water  to  degas.  The  water  becomes  supersaturated   with   respect   to   calcite   and   CaCO3   is   deposited   to   form  speleothems  (adapted  after  Fairchild  et  al.  2006).    When  vadose  water  enters  the  cave,  where  pCO2    levels   are   lower   than   in   the   limestone   crack   system,   CO2   degasses   from   the  water,  the  drip  water  becomes  supersaturated  with  respect  to  calcite,  and  calcite  is  precipitated  in  the  form  of  speleothems  according  to  the  reaction  [3]  (Fig.  1):    Ca2+  +  2HCO3-­‐  à  CaCO3  +  H2O  +CO2         [3]      The   CO2   degassing   of   the   water   occurs   immediately   within   10   seconds   after  entering   the   cave   atmosphere   through   molecular   diffusion   (Dreybrodt   and  Scholz,  2011).  When  the  drip  water  falls  on  top  of  a  stalagmite,  further  CO2  is  lost  due   to   diffusion   within   a   few   seconds   (Dreybrodt,   1980)   and   supersaturation  with   respect   to   calcite   increases   even     further.   It   is   likely   that   a  portion  of   the  drip  water  is  lost  during  the  impact  of  the  drop  on  the  surface  of  the  stalagmite  

Page 23: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  13  

due  to  splashing  effects.  The  zone,  where  speleothems  are  formed  is  referred  as  the  ‘precipitation’  zone  (Fig.  1)  (Fairchild  et  al.,  2006).      Around  the  drip  hanging  on  the  ceiling,  calcite  deposits  in  the  shape  of  a  rim.  The  consecutive  rims  at  the  ceiling  will   lead  to  a  straw  shaped  calcite  form  referred  as   “soda  straws”   (Fig.  2).  When   too   long,   these  straws  become  clogged  and   the  drip  starts  to  flow  along  the  surface  of  the  straw.  Over  time,  when  more  calcite  is  deposited,   the   soda   straw   evolves   into   a   stalactite.   This   explains   why  paleoclimate   reconstructions   cannot   be   based   on   stalactites.   The   center   of   the  stalactite  is  hollow  due  to  the  soda  straw  and  the  consecutively  deposited  calcite  layers   are   not   nicely   following   each   other.   However   in   stalagmites,   the  consecutive   falling   drips   will   cause   the   layers   to   be   deposited   on   top   of   each  other  with   the   youngest   layers   at   the   top  of   the   stalagmite   (Fig.2).   Stalagmites  are  consequently  better  suited  for  paleoclimate  reconstructions.    

 Figure  2.  Soda  straws  (left)  and  the  different  location  of  speleothem  deposits  in  caves  (after  Fairchild  et  al.  2006).    The  speleothem  growth  rate  depends  on  different  factors  such  as  (i)  drip  water  calcium   ion   concentration,   (ii)   discharge   amounts,   and   (iii)   the   pCO2   gradient  between  the  cave  air  and  drip  water  rendering  the  estimation  of  an  average  rate  difficult  (Genty  and  Quinif,  1996;  Baker  et  al.,  1998;  Genty  et  al.,  2001b;  Frisia  et  al.,   2003).   The   main   factor   driving   growth   rate   changes   is   (i)   the   drip   water  calcium  ion  concentration  (Genty  et  al.,  2001b),  which  can  increase  due  to  longer  residence  times  of  water   in  the  epikarst,  when  soil  pCO2  values  are  higher.  The  drip   water   calcium   ion   concentration   may   decrease   when   calcite   precipitates  from  the  water  before  entering  the  cave  (Prior  Calcite  Precipitation,  see  2.3.6  for  more  details  about  this  process),  a  process  mostly  occurring  during  drier  periods  when   aerated   zones   in   the   epikarst   become   more   important   (e.g.   Van  Rampelbergh  et   al.,   in   review)     .  Discharge   changes   (ii)  may  also  affect   growth  rate,  with  higher  discharge  causing  faster  calcite  deposition.  This  effect   is  more  pronounced   for   slow   dripping   sites,   where   the   amount   of   water   provided   to  precipitate  a  stalagmite  is  low.  A  stronger  pCO2  gradient  between  the  drip  water  

Page 24: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 14  

and  the  cave  air   (iii)  will  cause  stronger  CO2  degassing  of   the  drip  water  when  entering  the  cave,  which  causes  higher  calcite  supersaturation  of  the  drip  water.      Generally,  stalagmites  are  considered  to  grow  on  average  at  10-­‐100  μm/y  in  cool  temperate   climates   and   at   300-­‐500  μm/y   in   sub-­‐tropical   climates   (Fairchild   et  al.,  2006).  However,   in  reality  a   large  variety  of  growth  rates  are  observed.  For  example,   growth   rates   up   to   2   mm/y   may   occur   such   as   is   observed   in   the  Proserpine  stalagmite  from  the  Han-­‐sur-­‐Lesse  cave,  Belgium  (Van  Rampelbergh  et  al.,   in  review).   In  such  fast  growing  speleothems  annual  or  seasonal   layering  may  be  detected  if  growth  parameters  such  as  calcium  ion  concentration  or  drip  discharge   vary   annually   or   seasonally   (Genty   and   Quinif,   1996).   Layering   in  stalagmites  mostly   consists   of   alternating   dark   and  more   compact   layers  with  whiter   and   more   porous   layers   and   is   mostly   an   indicator   of   the   seasonal  changes  in  cave  climate  and  chemistry  (Treble  et  al.,  2005b;  Mattey  et  al.,  2008;  Van  Rampelbergh  et  al.,  2014).  Fluorescent  annual  lamina,  rich  in  organic  matter,  may  also  occur  in  speleothems  and  are  related  to  changes  in  organic  matter  flux  from   the   surface   (Frisia  et   al.,   2000;  Proctor  et   al.,   2000;  Fairchild  et   al.,   2001;  Genty  et  al.,  2001b),  or  to  seasonal  flushing  of  soil-­‐derived  elements  and  particles  (Frisia  et  al.,  2000).    The  time  needed  for  the  rainwater  to  reach  the  cave,  referred  to  as  the  transition  time,   depends   on   the   ‘flow-­‐type’   of   the   epikarst.   The   flow-­‐type   of   karst  groundwater  encountered  in  caves  vary  from  very  slow  dripping  seepage  flow  to  rather  high-­‐discharge  types,  including  shaft  flow  and  subcutaneous  flow  (Smart  and  Friedrich,  1987).  Drip  sites  feeding  actively  forming  stalagmites  are  typically  of   the   seepage   flow   type   (i.e.   low  discharge,   low   variability),   the   seasonal-­‐drip  type  (i.e.  low  discharge,  but  seasonal  variability),  or  the  vadose  flow  type  (higher  discharge,  commonly  giving  rise  to  flowstone  deposits).      Summarized,   three   major   sources   are   important   for   the   formation   of  speleothems:   rainwater,   soil   activity   and   the   presence   of   a   carbonate   bedrock.  The   calcite   layer   that   is   formed   consequently   reflects   the   geochemical   and  chemical  composition  of  the  water,  soil  and  bedrock  at  the  time  of  its  formation.        2.2  Dating  of  speleothems    The  development  of   techniques   for  precise  and  accurate  dating  of   speleothems  on   small   amounts   of   samples   has   allowed   speleothem   science   to   become   so  prominent  in  the  recent  years  (Fig.  3)  (Fairchild  and  Baker,  2012).  Speleothems  can  now  be  absolutely  dated  with  very  small  error  bars,  down  to  0.1  %,  for  the  ages  making  them  valuable  archives  in  dating  large  climatic  events.    

Page 25: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  15  

 

   Figure  3.  a)  An  overview  of  the  different  speleothem  dating  techniques  and  their  corresponding   timescales   (y-­‐axis).   When   visible,   layering   can   be   used   to   date  young   speleothems.   14C   can   be   used   in   young   speleothems   to   detect   the   1964  ‘bomb  peak’  as  anchor  point  for  the  age  model.  The  most  commonly  used  dating  technique   in   speleothems   is   the   U/Th-­‐dating   technique   that   delivers   absolute  ages   with   very   small   error   ranges.   The   recently   developed   U/Pb-­‐dating  technique   can  be  used   to  determine   speleothem  samples  older   than   the  U/Th-­‐age   limit   of   600   kyr.   b)   Methodological   advances   in   dating   techniques   have  increased   the   precision   of   ages   obtained   in   speleothems.   The   most   used  speleothem   dating   technique   is   based   on   ICP-­‐MS   (Inductively   Coupled   Plasma  Mass   Spectrometer)   determined   U/Th-­‐ages.   Picture   from   Fairchild   and   Baker  (2012).    2.2.1.  Dating  techniques    U-­‐series  or  U/Th-­‐dating   is  the  most  commonly  used  technique  when  working  with  speleothems.  The  U/Th  method  allows  precise  dating  of  the  samples  back  to  600  kyr  (Fig.  3a)  (Edwards  et  al.,  1987;  Cheng  et  al.,  2000;  Cheng  et  al.,  2009a;  2009b).  The  technique  is  based  on  the  decay  of  the  parent  isotope  238U  to  230Th.  Uranium   has   the   property   to   be   soluble   in   water   while   thorium   is   not.   Both  

Page 26: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 16  

elements   are   present   in   the   carbonate   bedrock.   The   drip   water   will   dissolve  uranium,  while  Th  will  link  to  bigger  colloids  or  organic  molecules.    When  calcite  is  deposited  from  the  drip  water  in  the  cave,  the  radioactive  238U  gets  enclosed  in  the   crystal   structure   at   concentrations   typically   varying   between   0.05   and   0.5  ppm.   From   there,   238U   decays   radioactively   according   to   its   series   until   stable  206Pb   is  produced.  For  U/Th-­‐ages,  only  a  part  of   the  decay  series   is  used,  being  the   decay   of   234U   to   230Th.   This   dating   method   is   different   from   other   dating  techniques   since   it   does   not   use   the   stable   end   of   the   decay   chain   but   rather  looks  at  the  balance  between  mother  234U  and  the  produced  radioactive  daughter  230Th.   The   amount   of   230Th   measured   in   the   speleothem   calcite   is   thus   an  indicator  of  the  age  of  the  calcite.  However,  ‘detrital  230Th’  can  be  incorporated  in  the  deposited  calcite  together  with  impurities  causing  artificial  older  U-­‐Th  ages.  The  measured  U/Th-­‐age  can  be  corrected  by  measuring  the  amount  of  232Th  that  is  used  as  proxy  for  the  detrital  230Th  in  the  sample.  Together  with  an  estimation  of   the   230Th/232Th   ratio   of   the   detrital   phase,   the   measured   U/Th-­‐age   can   be  corrected   for   its   amount   of   ‘detrital   230Th’.   The   230Th/232Th   ratio   can   be  estimated   using   the   isochron   techniques   or   better   by   measuring   the   ratio  directly  in  the  chemically  separated  detrital  phase  (Dorale  et  al.,  1998).  The  main  problem  when   using   the   isochron   technique   is   that   only   few   estimates   of   the  230Th/232Th  ratio  exists  and  that  it  is  high  likely  that  this  ratio  varies  from  site  to  site  and  through  time  at  a  given  site,  such  that   the  232Th-­‐corrected  ages  can  be  associated   with   extremely   large   errors.   This   is   especially   the   case   in   young  speleothems   or   with   samples   with   low   U-­‐concentrations   where   little   230Th   is  present   from  decay  of   uranium.  Dating   results  with   too  high   232Th   content   are  mostly   not   used   to   establish   the   age   model   of   the   stalagmite.   Although   the  problem   of   ‘detrital   230Th’   may   be   present   in   speleothems,   most   studied  speleothem   samples   are   not   contaminated   and   can   be   U/Th   datet   with   high  precision.    The  development  of   the  U/Th-­‐dating  technique  started   in  the  1970s  with   α-­‐spectrometry   and   has   been   further   developed   using   TIMS   (Thermal  Ionization  Mass  Spectrometer)  and   finally   ICP-­‐MS   (Inductively  Coupled  Plasma  Mass  Spectrometer)  based  measurements  (Fig.  3b)   that  allow  to  determine   the  ages   of   small   samples   between   100   and   400   mg   depending   on   the   U-­‐content  (only   ~2.5ng   of   U   is   needed).   Precision   goes   up   to   5   %   and   much   worse   for  detrital  contaminated  samples.      

Seasonal   layers,   if   visible,   can   be   used   to   establish   the   age   model   for  speleothems  covering  the  last  millennia  by  counting  the  layers  back  in  time  (Fig.  3a).  This  technique  has  the  large  advantage  to  be  non-­‐destructive  for  the  sample.  Due   to  small   changes   in  cave  climatology  between  seasons,   layers  deposited   in  stalagmites  are  often  seasonal  with  one  darker  and  one   lighter   layer  deposited  every   year   (Genty   and   Quinif,   1996).   Clear   seasonal   layering  mostly   occurs   in  speleothems  with  significant  high  growth  rates  (>1  mm/y)  and  has  proven  to  be  very  efficient  in  providing  absolute  time  chronologies  for  the  measured  proxies  

Page 27: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  17  

(Mattey  et  al.,  2008;  Van  Rampelbergh  et  al.,   in  review).  Not  only  visible   layers  can  be  used  for  counted  chronologies,  trace  elemental  cycles  are  often  observed  in  speleothems  with  growth  rates  higher  than  0.05  mm/y  and  can  also  be  used  to  provide  counting  ages  (Fairchild  et  al.,  2001;  Treble  et  al.,  2003;  2005b;  Smith  et  al.,   2009).   However,   not   all   stalagmites   display   layering   and   if   present,   the  layering   is  often   interrupted  by  hiatuses  or  parts  where   the   layering   is  hard   to  define,   making   it   difficult   to   estimate   when   to   restart   the   counting.   Therefore  counted  age  models  always  have  to  be  calibrated  with  U-­‐series  ages.      Radiocarbon  or  14C-­‐dating  can  also  be  applied  on  speleothems  and  was  the  first  dating  technique  used  to  date  speleothem  samples  in  the  1960’s  (Broecker  et  al.,  1960).   However,   the   problem  with   this   technique   is   that   the   amount   of   ‘dead  carbon’  coming  from  dissolution  of  the  host  rock  or  old  organic  matter  remains  in   the   epikarst   needs   to   be   estimated   to   correct   the   systematically   too   old  measured   14C-­‐ages.   This   dead   carbon   proportion   typically   varies   around   15%,  but  needs  to  be  estimated  for  each  sample  by  comparing  the  14C-­‐ages  with  U/Th-­‐ages   (Genty   and   Massault,   1997).   For   example,   14C   measurements   in   Scottish  stalagmites  suggest  a  dead  carbon  proportion  between  22  and  38  %  that   is  the  results  of  the  ageing  of  the  soil  organic  matter  related  to  peat  bog  development  above  the  cave  (Genty  et  al.,  2001a).  Even  when  the  ‘dead  carbon’  proportion  of  the   analyzed   sample   is   known,   14C   ages   have   to   be   calibrated   against   the  dendrochronologic   calibration   curve   causing   the   obtained   ages   to   have   larger  uncertainties   compared   to   the  U/Th   or   layer   counting   ages.   The  measured   14C  value   and   its  2σ  uncertainty   range  are  projected  against   the  dendrochonologic  calibration  curve.  Because  the  calibration  curve  is  not  a  straight  line  it  is  possible  to  have  many  intercepts  on  the  calendar  year  axis,  each  with  its  own  probability  range.  Ages  are  reported  as  an  interval  with  their  own  probability  range,  which  depends   strongly   on   the   measured   14C   activity.   The   14C   ages   obtained   on  speleothems   are   thus   flawed   and  mostly   not   used   to   establish   speleothem  age  models.  The  14C-­‐ages  can  be  useful   to  date  young  samples  where  no   layering   is  visible  and  too  little  uranium  or  too  high  amounts  of  detrital  230Th  are  present.  For   such   samples   the   speleothem   14C   activity   can   be   compared  with   the   1964  ‘bomb  test’  peak  delivering  absolute  ages  (Genty  and  Massault,  1997).  This  has  for   example   been   done   to   determine   the   age   of   a   recent   speleothem   from   St  Michael’s  cave,  Gibraltar,  where  the   ‘bomb  peak’  was  used  as  spike  and  the  age  model  was  established  counting  annual  geochemical  cycles  (Mattey  et  al.,  2008).    In  1998,   the  age   limit   to  date   speleothems  back   to  600  kyr  was  broken  by   the  publication   of   Richards   et   al.   (1998)   where   they   described   that   older  speleothems  could  be  dated  with  the  U/Pb  technique,  which  allows  dating  back  to   >   400  Ma  with   a   precision   of   1   to   5  %.   However,  more   than   a   decade  was  necessary  to  develop  suitable  methods  to  determine  these  ratios  (Walker  et  al.,  2006;  Woodhead  et  al.,  2006).  The  main  problem  with  U/Pb  ages  is  the  very  low  

Page 28: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 18  

levels   of   Pb   in   speleothems   and   the   difficulty   in   obtaining   a   range   of  parent/daughter   isotope   ratios   for   isochron   reconstructions.   The   development  of   MC-­‐ICPMS   formed   a   breakthrough   is   this   discipline   making   it   possible   to  determine   isotope   ratios   with   very   low   concentrations.   However,   this   dating  technique   for   speleothems   is   still   under   development   due   to   problems   with  chemical  opening  of  the  system  in  old  sampled  where  U  may  be  lost  and  Th  and  Pb  may  be  injected  in  the  system.  Bajo  et  al.  (2012)  successfully  used  the  U/Pb  dating  technique  to  date  an  Early  to  Middle  Pleistocene  (±  800  ka)  speleothem  of  Corcia  Cave  (Italy).  Vaks  et  al.   (2013)  used  U/Pb  dated  speleothem  depositions  in   an   Israeli   cave   to   reconstruct   climate   variations   during   the   Pliocene   and  Pleistocene   (±   3   Ma).   Water   table   speleothem   U/Th-­‐ages   provided   age  estimations  for  the  incision  of  the  Grand  Canyon  (Polyak  et  al.,  2008).      2.2.2.  Establishing  an  age-­‐depth  model    Since   the  measured  climate  proxies  have  a  higher   resolution   (mm  to  µm)   than  the  age  points  (mm  to  cm),  ages  need  to  be  interpolated  in  an  age-­‐depth  model  based  on  the  obtained  ages.  Small  growth  rate  changes  or  hiatuses  between  two  consecutive   dating   points   can   induce   significant   chronological   errors   into   the  model.  Therefore  is  it  important  to  sample  the  dating  points  at  strategic  places  in  the   stalagmite   where   growth   rate   changes   or   hiatuses   are   expected.   Such  changes  or  hiatuses  are  often  indicated  by  changes  in  the  crystallography  of  the  sample  or  in  the  measured  proxy  records.  The  most  common  approach  to  obtain  an   age   model   is   to   linearly   interpolate   the   ages   between   two   dating   points.  However,  the  age  model  is  only  based  on  two  adjacent  data  points,  is  not  smooth  at  data  points   and  usually  has  no  quantification  of   the   age   error  between  data  points.   Least-­‐squares   polynomial   fits   have   the   tendency   to   display   too   much  curving  and  can  create  overshoots  (Scholz  et  al.,  2012).  Splines  can  also  be  used  (Hodge  et  al.,  2008),  but   they  require  too  much  modification  of   the  parameters  making   the   age  model   too   subjective.   Different   authors   have   tried   to   establish  models  that  provide  the  best  possible  ages  for  speleothems  such  as  the  MODAGE  by   Hercman   and   Pawlak   (2012)   or   COPRA   by   Breitenbach   (2012),   but   these  model  are  mostly  complex  and  difficult  to  use.  The  StalAge  age  model  established  by   Scholz   and   Hoffmann   (2011)   delivers   the   best   age-­‐depth   model   for  speleothems   currently   available   that   is   easy   to   use.   However,   the  weakness   of  the  age  model  is  that  it  tries  to  fit  straight  lines  through  three  adjacent  age-­‐points  averaging   out   possible   small   growth   rate   changes   between   points.   The   best  results   are   obtained   by   comparing   the   age-­‐depth   relation   obtained   by   StalAge  with  crystallographic  and  climate  proxy  variations  in  the  speleothem.  Models  to  establish  age-­‐depth  relationships   in  speleothems  are  useful  but  still  need   to  be  interpreted  critically  by  the  author.  The  location  of  the  age  points  is  still  a  very  important   factor   in   the   correctness   of   the   age   model   and   needs   to   be   well  thought  before  sampling  

Page 29: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  19  

2.3  δ18O  and  δ13C  in  Speleothems    2.3.1  Factors  that  determine  the  δ18O  of  the  drip  water      1.  δ18O  VARIATIONS  IN  THE  RAINWATER    The  water  dripping  from  a  cave  ceiling  originates  from  the  part  of  the  rainwater  that   is   infiltrated   in   the   soil,   which   is   the   difference   between   the   amount   of  precipitation   and   the   amount   of   water   lost   by   evaporation   and   plant  transpiration   (evapo-­‐transpiration).   The   δ18O   composition   of   rainwater   is  influenced   by   different   isotope   effects   being   the   latitude   effect,   the   continental  effect,   the   altitude   effect,   the   temperature   effect   and   amount   effect   (Dansgaard,  1964;   Rozanski   et   al.,   1992).   Derived   effects   are   the   seasonal   effect   due   to  seasonal   variations   of   temperature   and/or   amount,   and   the   source   effect   by  geographical  location  of  the  source  of  rainwater.    

 Figure  4.  Schematic  oxygen  (and  hydrogen)  isotope  fractionation  of  water  in  the  atmosphere  (adapted  after  Hoefs  (1997)).      The  latitude  effect  relates  to  the  gradual  depletion  of  the  δ18O  (and  also  δD)  in  the  atmospheric  water  during  its  successive  rainouts  (Fig.  4)  on  its  way  to  the  poles  (Rozanski   et   al.,   1992).   The   evolution   of   the   rainwater   δ18O   signal   due   to   the  successive   rainouts   of   the   clouds   follows   the   Rayleigh   distillation   model  (Rayleigh,   1896)   and   roughly   equals   -­‐0.18   ‰   per   latitudinal   degree   of   the  Northern   Hemisphere   (Rozanski   et   al.,   1992).   The   same   occurs   when   a   cloud  moves  inland  from  the  coast.  This  effect  is  referred  as  the  continental  effect  and  

Page 30: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 20  

describes   the   gradual   depletion   in  δ18O   (and   also  δD)   of   the   rainwater   further  away   from   its   water   source.   Recently,   McDermott   et   al.   (2011)   demonstrated  that   the  continental  effect  observed   in   the  European  precipitation   is  preserved  by  speleothems  δ18O  values.  The  altitude  effect  relates  to  the  fact  that  the  higher  the   altitudinal   location  of   the   rain,   the  more   the  water   becomes  depleted  with  generally  -­‐0.2  to  -­‐0.3  ‰  per  increase  of  100  m  (Lachniet  and  Patterson,  2006).  For   one   geographical   location   the   latitude,   altitude   and   the  distance   in-­‐land  or  continentality   determine   the   rainwater   isotopic   composition   and   can   be  estimated  constant  during  relatively  short  geological  periods  of  time.      The  rainwater   isotopic  composition  at  a  certain  geographical   location  will  vary  over  time  due  to  the  temperature  effect,  the  amount  effect  or  the  source  effect.  The  temperature   effect   relates   to   the   influence   of   temperature   on   the   δ18O   of  precipitation   through   time.   During   warmer   periods,   more   heavy   isotopes   will  evaporate   from   the   source   of   the   rainwater   (mostly   the   ocean)   and   cause  increased  rainwater  δ18O  values  (Rozanski  et  al.,  1992).  The  temperature  effect  on   the   rainwater   δ18O   signal   can   vary   between   +0.17   and   +0.9   ‰/1°C,  depending  on  the  geographical  location  (Dansgaard,  1964;  Rozanski  et  al.,  1992;  Mook,  2000;  Schmidt  et  al.,  2007).  Schmidt  et  al.  (2007)  suggested  a  dependence  of  +0.3  ‰/1°C  for  central  Europe.  When  colder  and  warmer  seasons  are  present  the   temperature   effect   will   cause   a   seasonal   effect   on   the   rainwater   isotopic  composition   with   isotopically   lighter   rainwater   during   the   colder   season.   The  seasonal   effect   is   the   smallest   at   the   equator,   where   temperature   remains  constant   throughout   the   year   but   is   more   pronounced   at   the   poles,   where  temperature  can  vary  more  strongly  between  winter  and  summer.  The  amount  effect   implies   that  more   rain   causes   the  water   to   be  more  depleted   in   isotopic  composition  (Rozanski  et  al.,  1992).  During  heavy  rainfalls,  the  water  drop  only  partially   exchanges   its   oxygen   isotopic   composition   with   surrounding   more  positive   air   preserving   its   more   negative   δ18O   composition   from   in   the   cloud  (Gat,   1996).   In   tropical   and   sub-­‐tropical   areas,   where   precipitation   intensities  display  large  variations  through  monsoonal  periods  the  amount  effect  will  have  the   largest   influence   on   the   isotopic   composition   of   the   rainwater.   For   one  geographical   location,   rainwater   can  also  originate   from  different   sources  with  different   isotopic   compositions   inducing   a   source   effect   leading   to   abrupt  changes  in  the  rainwater  isotopic  composition.      2.  δ18O  VARIATIONS  IN  THE  SOIL  ZONE    The   δ18O   of   the   water   in   the   soil   zone   is   expected   to   represent   the   amount-­‐weighted  mean  of  the  δ18O  of  infiltrating  rainwater  that  may  be  further  modified  by   evaporation.   In   (semi-­‐)arid   regions   the   soil-­‐   and   groundwater   δ18O   can  increase  compared  to  the  weighed  mean  of  the  rainwater  δ18O  due  to  substantial  

Page 31: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  21  

evaporation   of   water   from   surface   and   vadose   zone.   Also,   when   a   part   of   the  rainwater  evaporates  before  entering  the  soil,  such  as  is  often  the  case,  the  δ18O  signal   of   the   drip   water   can   increase.   The  magnitude   of   δ18O   increase   will   be  related   to   the   relative   humidity   in   the   soil   pores   and   the   evaporated   water  volume  (Lachniet,  2009).  However,  the  most  intense  rainstorms  that  commonly  have   low   δ18O   values   are   likely   to   dominate   recharge   into   the   soil   zone,   and  would   tend   to   counteract   the   isotopic   enrichment   associated  with   evaporation  (Dansgaard,   1964;   Rozanski   et   al.,   1992).   Plant   respiration   does   not   affect   the  δ18O  composition  of  water  (Gat,  1996).      3.  δ18O  VARIATIONS  IN  THE  EPIKARST    The  timing  and  amount  of  recharge  to  the  epikarst  is  an  important  control  on  the  resulting  drip  water  δ18O.  For  most  western  European  caves  such  as  the  Han-­‐sur-­‐Lesse   cave   (Belgium,   Bonniver,   2011)   or   the   Bunker   Cave   (western-­‐Germany,  Fohlmeister  et  al.  2012)  ,  recharge  occurs  in  winter  causing  the  δ18O  variations  in  the  speleothem  to  reflect  winter  precipitation  changes  (Fohlmeister  et  al.,  2012;  Van  Rampelbergh  et  al.,   in  review).   In  tropical  and  sub-­‐tropical  regions  such  as  Oman,  heavy  monsoon   rains   recharge   the  epikarst   and  only   a   small   amount  of  water   evaporates   causing   the   drip   water   δ18O   to   reflect   the   rainwater   annual  mean   δ18O   composition   (Fleitmann   et   al.,   2004).   If   evaporation   processes   are  more  important,  the  δ18O  composition  of  the  water  in  the  epikarst  can  increase  compared  to  the  annual  mean  rainwater  δ18O  (BarMatthews  et  al.,  1996).    The   transition   time   of   water   in   the   vadose   zone   determines   how   rainwater  signals  are  transferred  to  the  cave.  This  transition  time  can  be  estimated  by  lag  times  of  δ18O  values  (using  them  as  an  built-­‐in  tracer)  in  drip  waters  relative  to  rainfall   values   (Baldini   et   al.,   2006;   Cobb   et   al.,   2007),   chemical   variations   in  karst  waters  (Genty  and  Deflandre,  1998),  delay  response  in  precipitation  events  (Cobb   et   al.,   2007;   Van   Rampelbergh   et   al.,   2014)   from   fluorescent   tracers  (Bonniver,  2011),   fluorescent  organic  matter   in   the  drip  water   (Hartland  et  al.,  2010)  or  by  tritium  (3H)  tracing  (Kluge  et  al.,  2010).  Depending  on  the  epikarst  thickness  and  epikarst  flow  systems,  the  transfer  time  of  the  drip  water  from  the  soil   to   the   cave   varies   from  months   to   several   years   such   as   observed   in   the  shallow  (-­‐15   to   -­‐30m)  Ernesto  Cave,   Italy   (Miorandi  et  al.,  2010),  Bunker   cave,  Germany   (Kluge   et   al.,   2010)   and   Han-­‐sur-­‐Lesse   Cave,   Belgium   (Van  Rampelbergh  et   al.,   2014).  For  very  deep  cave   systems,   such  as  Monte  Corchia  cave   in   central   Italy   (-­‐1187m)   drip   water   residence   times   of   decades   can   be  observed  (Piccini  et  al.,  2008).  Transit  time  is  expected  to  be  most  rapid  for  the  conduits,  and  slowest  for  the  diffuse  seepage  flow.  In  a  general  sense  and  all  of  the   factors   being   equal,   the   thicker   the   overlying   limestone,   the   longer   the  potential   transit  time  and  the  more  complete  the  groundwater  mixing.  Systems  

Page 32: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 22  

with   shorter   transit   times   will   be   more   suitable   for   capturing   rapid,   high-­‐frequency  climate  events  (McDonald  et  al.,  2007),  whereas  very  slow  transit  time  with   substantial  mixing  will   be  more   suited   for   constraining   long-­‐term  climate  change.    The  saturation  state  of   the  drip  water  may  vary  over  time,   thus   influencing  the  timing  of  calcite  deposition  (Treble  et  al.,  2005a;  Baldini  et  al.,  2006).  Only  those  recharge  waters,  with  a  given  isotopic  composition  that  are  saturated  with  CaCO3  will   participate   to   the   deposition   of   speleothem   CaCO3.   It   is   this   isotopically  effective  recharge,  i.e.  the  recharge  water  with  a  certain  isotopic  signature  that  is  saturated  with  CaCO3,   that   is   relevant   to   the   interpretation  of   speleothem  δ18O  time  series  (Lachniet,  2009).  The  timing  of  isotopically  effective  recharge  may  be  forced  by  seasonal  variations  in  drip  water  and  cave  air  pCO2,  which  influences  drip  water   degassing   rates,   and   can   impart   a   seasonal   bias   to   the   speleothem  record   if   certain  months   produce  more   calcite   than   other   ones   (Baldini   et   al.,  2008;  Mattey  et  al.,  2008).  Ideally,  comprehensive  studies  of  infiltration  and  drip  water   geochemical   measurements   (δ18O,   pH,   pCO2,   calcite   saturation   indices)  over   the   course   of   several   years   should   be   completed   to   understand   the  saturation  variations  and  the  timing  of  drip  and  of  the  transfer  of  isotopic  signals  (Baldini  et  al.,  2006;  Mattey  et  al.,  2008).    The   δ18O   value   of   the   drip   water   is   a   function   of   seasonality   of   recharge   and  modification  within  the  soil  and  epikarst.  Typically  drip  waters  δ18O  variability  is  attenuated  relative   to  precipitation  δ18O  due   to  mixing   in   the  soil  zone  and   the  epikarst  (Mattey  et  al.,  2008;  Genty  et  al.,  2014;  Van  Rampelbergh  et  al.,  2014)    2.3.2  Factors  that  determine  the  δ13C  of  the  drip  water    The  carbon  isotopic  composition  of  the  drip  water  is  mainly  determined  by  the  δ13C  of  soil  CO2  and  limestone  and,  to  a  smaller  part,  by  the  dissolution  intensity  of   the   host   rock,   atmospheric   CO2   or   pCO2   in   the   cave   atmosphere   (Dreybrodt  and  Scholz,  2011).        Rainwater  equilibrates  with  the  atmospheric  pCO2  and  with  its  δ13C  composition.  The   increased   combustion   of   isotopically   light   organic   carbon   over   the   past  nearly  two  centuries  has  caused  the  δ13C  composition  of  the  atmospheric  CO2  to  decrease   due   to   the   ‘Suess-­‐effect’   (Suess,   1955)   from  ∼   -­‐6.4  ‰  between   1000  and  1800  AD  to  about  -­‐8.2  ‰  in  2011  (Cuntz,  2011).      The  major   sources   of   CO2   in   the   soil   zone   are   the   vegetation   and   the   decay   of  organic  matter  (Salomons  and  Mook,  1986).  The  soil  CO2  composition  is  mainly  driven  by  the  intensity  and  type  of  vegetation  (C3,  C4  or  CAM  plants)  above  the  

Page 33: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  23  

cave.  The  δ13C  values  of   soil  CO2  around   -­‐26  ‰  reflect   a  C3-­‐type  of   vegetation  that   is   typical   for   humid   climates  while   δ13C   values   of   soil   CO2  around   -­‐13  ‰  reflect   a   C4-­‐type   of   vegetation   that   is   typical   for  water   stressed   environments  (Salomons   and   Mook,   1986;   McDermott   et   al.,   2005).   During   periods   of  decreased  plant  CO2   input   in   the   total   soil  CO2   reservoir,   atmospheric  CO2   that  infiltrates  into  the  soil  can  influence  the  soil  δ13C  signature  by  several  ‰  (Baker  et  al.,  1997).  Within  the  soil,  the  high  soil  pCO2  (0.1  atm)  compared  to  the  water  pCO2   (0.001   atm)   forces   the   dissolution   of   soil   CO2   in   the  water.   This   process  goes  on  until  the  soil  CO2  partial  pressure  is  attained  in  the  water.  Consequently,  higher  soil  pCO2  values  will  cause  more  CO2  to  be  dissolved  in  the  water.      In   the  epikarst,   the  corrosive  water  dissolves   the  carbonate  bedrock.  Two  end-­‐member  models,  being  an  open  or  a  closed  system,  describe  how  this  dissolution  process   occurs   (Hendy,   1971;   Salomons   and   Mook,   1986).   In   an   open   system  model,   continuous   equilibration   occurs   between   the   seepage   water   and   the  ‘infinite’  reservoir  of  soil  CO2  during  the  dissolution.  Under  these  conditions,  the  δ13C  value  of  the  water  reflects  the  δ13C  composition  of  the  soil  CO2  (mainly  from  vegetation).   Under   closed   conditions,   the   percolating   water   becomes   isolated  from  soil  CO2  as  soon  as   the  carbonate  dissolution  starts.  The  δ13C  value  of   the  drip   water   is   influenced   by   the   δ13C   value   of   the   host   rock   typically   varying  around  0  ‰  (Salomons  and  Mook,  1986),   the  host   rock  nearly   always  being  a  marine  limestone.      Most  natural  cave  systems  are  likely  to  be  partially  open  causing  the  δ13C  value  of  the  drip  water  to  be  mainly  determined  by  the  isotopic  composition  of  the  soil  CO2  (Dreybrodt  and  Scholz,  2011).  Genty  and  Massault  (1997)  estimated  a  ca.  15  %  contamination  of   the  soil  water  δ13C  by   limestone  δ13C.  However,   this  %  can  vary  according  to  the  amount  of  dissolved  CaCO3,  which  relates  to  the  degree  of  the  open/closed  dissolution  system  and  of  the  pCO2  in  the  soil.  At  a  given  pCO2  value,   under   a   more   open   dissolution   system,   more   CaCO3   will   be   dissolved,  because  more  CO2   is  available  during  the  dissolution   in  comparison  to  a  closed  system  state  (Hendy,  1971).  Under  a  given  dissolution  system,  the  higher  the  soil  pCO2  value,  the  more  corrosive  the  water  and  the  more  CaCO3  can  be  dissolved  in  the   water.   More   intense   dissolution   of   isotopically   heavy   CaCO3   will   cause   an  increase  of  the  drip  water  δ13C.          In   the   cave,   the   δ13C   composition   of   the   drip   water   can   also   be   influenced   by  equilibration  with   the  δ13C  values  of   the  cave  air.  Dreybrodt  and  Scholz  (2011)  studied   the   possible   effect   of   equilibration   between   the   pCO2   of   the   dripwater  and   the   pCO2   of   the   cave   atmosphere.   Isotopic   exchange   between   the   carbon  isotopes  in  the  drip  water  and  the  carbon  isotopes  of  the  cave  air  CO2  drives  the  δ13C  of  the  drip  water  to  the  δ13C  value  of  the  cave  air  CO2.  Equilibration  between  

Page 34: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 24  

the  δ13C  of  the  drip  water  and  δ13C  of  the  cave  air  needs  about  1  h.  This  process  will,   however,   not   affect   the   drip  water   δ13C   if   the   interaction   time   is   4   times  shorter  than  the  time  needed  for  full  equilibration  (=  1h)  (Dreybrodt  and  Scholz,  2011).  Consequently,  only  drip  water,  which   falls  on   the   stalagmite  within   less  than   15   minutes   after   entering   the   cave   atmosphere,   will   have   a   δ13C  composition   that   is   not   affected   by   the   δ13C   composition   of   the   cave   air  (Dreybrodt  and  Scholz,  2011).      2.3.3  Equilibrium  fractionation  factors      For  a  given  δ18O  or  δ13C  composition  of  the  drip  water,  the  isotopic  composition  of   the   calcite   formed   in   equilibrium   will   be   determined   by   its   equilibrium  fractionation   factor.   The   oxygen-­‐isotope   equilibrium   fractionation   factor  between  water  and  the  precipitated  calcite  is  temperature  dependent  and  results  in   the   preferential   incorporation   of   18O   in   the   solid   phase.   The   equilibrium  fractionation  factor  of  the  δ13C  also  leads  to  the  preferential  incorporation  of  13C  in   the   calcite,   but   is   much   less   sensitive   to   temperature   than   in   the   case   of  oxygen.      Different   studies   and   authors   have   tried   to   determine   the   δ18O   water-­‐calcite  fractionation   factor   (Table   1).   The   mostly   used   laboratory   established  fractionation   factors   remain   the   ones   by  O'Neil   et   al.   (1969)   later  modified   by  Friedman   and   O'Neil   (1977),   the   relationship   of   Kim   and   O'Neil   (1997)   later  modified  by  Kim  et  al.  (2007),  the  results  of  Tarutani  et  al.  (1969)  and  Jimenez-­‐Lopez   et   al.   (2001)   (Table   1).   A   second   approach   to   determine   fractionation  factors   is   established  by  using   theoretical  models   as   the   ones   from  Horita   and  Clayton   (2007)   and   Chacko   and   Deines   (2008).   A   third   approach   consists   of  using   cave-­‐monitoring   data   to   make   an   average   of   the   in-­‐cave   observed  fractionation  factors  (Demeny  et  al.,  2010;  Tremaine  et  al.,  2011).  Tremaine  et  al.  (2011)  established  such  a   'cave  calcite'  relationship  by  doing  a  best   fit   through  the  data  on  a  large  number  of  modern  caves  at  different  latitudes,  altitudes  and  temperatures.     The   different   studies   display   different   reactions   for   the  fractionation  factor  and  different  temperature  dependencies  (Table  1).  However,  they  only  suggest  small  differences   for  the  enrichment  values  around  10°C  and  they  all  indicate  a  δ18O  increase  of  about  0.2  ‰  in  the  deposited  calcite  per  1°C  decrease  of  the  depositional  temperature.  This  is  the  value,  used  in  this  study.              

Page 35: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  25  

Author   Method   1000*lnα   dα/dT  (‰/°C)  

O'Neil  et  al.,  1969  modified   by  Friedman   &   O'Neil  1977  

Laboratory   2.78(106T-­‐2)-­‐2.89   -­‐0.24  

Kim   &   O'Neil   1997  modified   by   Kim   et  al.,  2007  

Laboratory   18.03(103T-­‐1)-­‐31.17   -­‐0.22  

Chacko   &   Deines  2008    constructed   relation  by   Tremaine   et   al.,  2011  

Theoretical  calculation   2.573(106T-­‐2)-­‐0.869   -­‐0.22  

Horita   &   Clayton  2007  

Calculations  compared   with  experimental  results  

0.952(106T-­‐2)+11.59(103T-­‐1)-­‐21.56   -­‐0.23  

Tremaine   et   al.,  2011  

Linear   best-­‐fit  through   large  number   of   cave  studies  

(16.01±0.65)*(103T-­‐1)-­‐(24.6±2.2)   -­‐0.17  

Demeny  et  al.,  2010  Cave   monitoring  results   Hungarian  cave  

17500*T-­‐1-­‐29.89   -­‐0.22  

Table   1.   A   selection   of   the   most   commonly   used   water-­‐calcite   oxygen  fractionation   factors.   Laboratory   and   theoretical   approaches   differ   from   the  relationships  found  in  cave  settings.    The  δ13C  of  calcite  deposited  in  equilibrium  with  the  δ13C  signal  of  the  dissolved  inorganic  carbon  (DIC)  in  the  water  is  determined  by  the  δ13C  of  the  DIC  and  the  fractionation  factor  between  the  DIC  and  the  calcite.  Different  studies  have  tried  to   determine   an   equilibrium   fractionation   factor   for   the  δ13C   between  DIC   and  calcite.  A  selection  is  listed  in  Table  2.  We  used  the  results  of  Emrich  et  al.  (1970)  and  Dulinski  and  Rozanski  (1990)  because  experimental  conditions  are  close  to  those  in  caves  and  because  they  are  based  on  a  compilation  of  data.  In  contrast  to  the  temperature  dependent  oxygen  fractionation  factor,  the  carbon  fractionation  factor   between   the   dissolved   inorganic   carbon   (DIC)   in   the   water   and  precipitated   calcite   is   much   less   temperature   dependent   (ca.   -­‐0.06   ‰/1°C  between  0  and  25  °C,  Table  2).              

Page 36: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 26  

Author   Method   ε   dα/dT  (‰/°C)  

Emrich  et  al.,  1970   Compilation  of  data   10.17  ±  0.18  (20°C)   -­‐0.063  ±0.008  (20°C)  

Dulinski   and  Rozanski   1990   and  references    

Calcite  precipitated  from   a   carbonate  solution   by  removal  of  CO2  

10.14  (10°C)  9.47  (15°C)   -­‐0.07  (5-­‐15°C)  

Romanek   et   al,  1992  

Calcite  precipitation   from  a   NaHCO3,   CaCl2  NaCl   solution   by  removal  of  CO2  

11.98  (±0.13)  -­‐0.12(±0.01)*T(°C)    

-­‐0.12  ±  0.1  (10-­‐40°C)  

Table   2.   Different   values   for   the   carbon   fractionation   factor   between   DIC   and  calcite   reported   in   enrichment   values   (ε)   at   the   experiment   temperature  with  their  temperature  dependency.    2.3.4  Disequilibrium  fractionation,  new  insights    In  a  pioneering  publication,  Hendy  (1971)  examined  the  effect  of  different  modes  of  calcite  deposition  on  speleothem  δ18O  and  δ13C  signals  and  concluded  that   if  the   loss   of   CO2   from   the   solution   is   slow,   the   precipitated   calcite   will   be   in  isotopic   equilibrium   with   the   solution.   In   this   case,   speleothem   δ18O   and   δ13C  values  will   only  depend  on   the   isotopic   composition  of   the  drip  water   and   the  corresponding   temperature   dependent   fractionation   factors.   If   however,  degassing   of   CO2   is   rapid,   kinetic   fractionation   will   occur   and   δ18O   and   δ13C  values  will  show  a  simultaneous  enrichment  along  individual  growth  layers.  The  ‘‘Hendy  test”  consists  of  analyzing  δ18O  and  δ13C  values  in  a  minimum  number  of  ~  10  samples   taken  within  a  single  growth   layer  and  at   increasing  distance  on  both  sides  from  the  growth  axis.  If  the  δ18O  and  δ13C  values  follow  a  co-­‐varying  increasing   trend,   this   observation   is   considered   to   be   the   result   of   kinetic  fractionation  during  calcite  deposition  in  the  studied  growth  layer.    Apart   from   the   equilibrium   and   kinetic   fractionation,   the   ‘Hendy’-­‐theory   also  describes   disequilibrium   fractionation,   being   intermediate   between   kinetic  fractionation   and   equilibrium   fractionation   (Hendy,   1971).   A   longer   residence  time   of   the   water   on   the   surface   of   the   stalagmite   leaves   more   time   for   CO2  degassing.   More   CO2   degassing   will   lead   to   a   gradual   enrichment   in   the   δ13C  signal  of  the  water  due  to  the  preferential  removal  of  light  carbon  isotopes.  The  δ18O   signal,   however,   remains   practically   unaffected   because   the   oxygen  reservoir   in   the  water   film   is   several  magnitudes   (104)   vaster   than   the   carbon  reservoir  (DIC),  and  the  CO2  escape  is  faster  than  the  evaporation  of  H2O  in  the  cave  environment  (Hendy,  1971).      

Page 37: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  27  

 This  disequilibrium  theory  has   later  been  proved   incorrect   for   the  evolution  of  the  δ18O  of  the  DIC  in  the  water  film.  A  recent  study  by  Beck  et  al.  (2005)  showed  that   the   commonly   used   exchange   time   between   the   oxygen   isotopes   of   the  water  and  the  oxygen  isotopes  of  the  DIC  was  too  low  by  one  order  of  magnitude.  Dreybrodt   (2008)   and   Dreybrodt   and   Scholz   (2011)   used   this   new   longer  exchange  time  to  model  the  disequilibrium  effects  on  the  δ18O  and  δ13C  signals  in  the  water   film  during   calcite   deposition   and  delivered   important   new   insights.  The   new   used   time   constants   by   Dreybrodt   and   Scholz   (2011)   and   Dreybrodt  (2011)  were   later  confirmed   in  experimental  setups  studying   the  chemical  and  isotopic  processes  in  the  water  covering  the  stalagmite  (Hansen  et  al.,  2013).  The  new  insights  are:    

1) There  is  no  such  thing  as  ‘fast’  or  ‘enhanced’  CO2-­‐degassing.    Calcite  deposition  from  water  that  enters  the  cave  can  be  divided  into  three  steps  (Fig.  5)   (Dreybrodt  and  Scholz,  2011;  Hansen  et  al.,  2013).  The   first  step   is   the  CO2-­‐outgassing   of   the   solution   by  molecular   diffusion   that   occurs   immediately  (mostly  within  10s)  after  the  drip  water  (at  pH  ≈  7)  enters  the  cave  (Hansen  et  al.,  2013).  In  a  second  step,  the  solution  equilibrates  with  the  low  pCO2  and  the  pH   increases   to   a   value   around   8.   This   step   is   one   order   of  magnitude   longer  compared  to  the  CO2  outgassing  and  takes  several  100s  (Zeebe  et  al.,  1999).  In  a  third  step,  CaCO3  precipitates  from  the  solution.  This  step  is  the  longest  and  is  in  the  order  of  several  1000s.  Precipitation  of  calcite  from  the  solution  during  the  third  step  converts  HCO3-­‐  into  CO2  that  degasses  from  the  solution.  CO2  degassing  from  on  the  water  film  covering  the  stalagmite  is  caused  by  precipitation  of  the  calcite  (Dreybrodt  and  Scholz,  2011;  Hansen  et  al.,  2013).      To   conclude,   first   CO2   outgassing   occurs   due   to   a   pressure   difference   between  the  drip  water  and  the  cave.  The  oversaturated  solution  starts  to  deposit  calcite  to  restore  the  chemical  equilibrium.  The  precipitation  of  calcite  from  the  solution  further  converts  HCO3-­‐   into  CO2.  This  produced  “new”  CO2  will  then  degas  from  the   solution.   Degassing   is   thus   caused   by   the   precipitation   of   calcite.   A   longer  residence   time   of   the   solution   on   the   surface   of   the   stalagmite   allows   more  calcite  to  be  deposited  and  thus  a  larger  production  of  CO2.  It  is  the  degassing  of  this  newly  produced  CO2  that  gradually  increases  the  δ13C  signal  of  the  DIC  in  the  solution.    

Page 38: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 28  

 Figure  5.  Evolution  of  the  concentration  of  the  [CO2]  and  the  [Ca]  in  the  solution  together  with  the  pH  and  timing  of  the  three  different  steps  from  the  entrance  of  the  drip  in  the  cave  atmosphere  to  the  deposition  of  calcite.  (Hansen  et  al.,  2013,  adapted   by   Dreybrodt,   presentation   at   the   S4   Speleothem   Summer   School,  Heidelberg,  2013).    

2) After  CO2  degassing,  the  δ18O  of  the  DIC  does  not  have  the  time  to  re-­‐equilibrate   isotopically   with   the   δ18O   in   the   water   before   calcite   is  deposited.    

 When  the  drip  enters  the  cave,  degassing,  driven  by  the  pCO2  difference  between  the  drip  water  and  the  cave  air,  causes  preferential  removal  of  both  the  light  12C  and  16O   isotopes   from  the  solution.  This  occurs  within  seconds  (Dreybrodt  and  Scholz,   2011;   Hansen   et   al.,   2013).   Removal   of   light   CO2  molecules   causes   the  δ13C   value   of   the   bicarbonate   in   the   solution   to   increase.   The   increase   can   be  described  as  a  Rayleigh  distillation  process  (Salomons  and  Mook,  1986)  or  as  a  more   recently   developed  model   referred   as   the   ‘kinetic   approach’   (Dreybrodt,  2008).  The  removal  of  light  16O  will  also  cause  the  δ18O  of  the  DIC  in  the  solution  to   increase.   However,   buffering   effects   due   to   the   isotopic   oxygen   exchange  between  the  water  in  the  solution  and  the  bicarbonate  in  the  solution  (hydration  and   hydroxylation)   will   restore   the   δ18O   signal   of   the   HCO3-­‐   to   its   initial  composition.   This   process   of   re-­‐equilibration   in   the   thin   water   film   on   the  surface  of  the  stalagmite  is  longer  as  first  expected  and  takes  between  6  000  and  

Page 39: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  29  

65   000   seconds,   depending   on   the   temperature   (Dreybrodt   and   Scholz,   2011).  Full  equilibrium  between  the  δ18O  in  the  water  and  the  δ18O  of  the  DIC  is  reached  after  4  times  the  exchange  time  (general  rule  handled  by  Dreybrodt  and  Scholz,  2011).  Restoring  the  oxygen  isotope  equilibrium  between  the  DIC  and  the  H2O  in  the   solution   is   consequently   a   long   process.   Calcite   deposition,   starts   about   ±  1000   s   after   degassing,   being   long   before   the   equilibrium   can   be   reached.   CO2  degassing  will   thus  always   cause   increased  δ18O  values   (Dreybrodt  and  Scholz,  2011).    

3) The  δ18O  signal  of  the  rainwater  will  only  be  reflected  in  the  calcite  if  the   residence   time   of   the   water   in   the   epikarst   is   long   (i.e   several  days).  

 Speleothems   reflect   the   δ18O   composition   of   the   DIC   in   the   water   from  which  they   are   precipitated.   Because   the   amount   of   water  molecules   is  much   higher  (105  times)  compared  to  the  amount  of  DIC  molecules,  the  δ18O  signal  of  the  DIC  was  commonly  expected  to  have  equilibrated  with  the  water  δ18O  signal  (Scholz  et  al.,  2009).  However,  at  a  temperature  of  10°C,  equilibrium  between  the  water  and   the  DIC   is   only   established   after  ±  1.6  days   (Dreybrodt   and   Scholz,   2011).  Consequently,   the   δ18O   of   the   DIC   will   only   reflect   the   rainwater   δ18O   if   the  residence  time  in  the  epikarst  is  sufficiently  long  (i.e.,  several  days)  to  equilibrate  the  isotopes  between  the  water  and  the  dissolved  carbon  species  (95  ‰  HCO3-­‐  at  pH  =  8)  (Dreybrodt  and  Scholz,  2011).  If  equilibration  is  not  completed,  the  δ18O  of  the  DIC  and  consequently  the  speleothem,  will  not  reflect  the  δ18O  of  the  water  (Dreybrodt   and   Scholz,   2011).   This   new   insight  will   also   have   implications   for  the  process  of  Prior  Calcite  Precipitation  (see  section  2.3.6).    

4) H2O  evaporation  of  the  water  film  covering  the  stalagmite  needs  to  be  significantly  high  before  having  an  effect  on  the  δ18O  of  the  deposited  calcite    

 Evaporation  of  the  thin  water  film  covering  the  surface  of  the  stalagmite  tends  to  be  high  when  relative  humidity  is  low,  the  drip  water  has  a  long  residence  time  on   the   surface   of   the   stalagmite   or   stalactite   tip   and/or   the   cave   is  windy   and  well   ventilated.   After   evaporation,   the   isotopic   composition   in   the   water   film  covering   the   stalagmite   is   enriched   in   18O.   The   δ18O   value   of   the   DIC,   mainly  consisting   of   HCO3-­‐   at   pH   ≈   8,   remains   initially   unaffected   by   evaporation.  However,   since   105   times   more   water   molecules   are   present   in   the   solution,  HCO3-­‐   that   is   in   disequilibrium   with   the   water   will   start   to   re-­‐equilibrate   its  oxygen   isotopic   composition   through   hydration   and   hydroxylation   processes.  However,  these  equilibration  processes  are  slow  and  can  take  up  to  several  days  (Beck  et  al.,  2005;  Dreybrodt  and  Scholz,  2011).  Since  precipitation  occurs  within  

Page 40: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 30  

±1000s  after  CO2  degassing,  the  HCO3-­‐  that  determines  the  isotopic  signal  of  the  deposited   calcite   will   not   have   the   time   to   equilibrate   with   the   light   water  isotopes  of  the  H2O.  The  formed  calcite  thus  reflects  the  isotopic  composition  of  the   water   before   evaporation.   However,   under   certain   boundary   conditions  evaporation   processes   may   cause   increased   δ18O   values   in   the   formed   calcite  (Deininger   et   al.,   2012).   If   the   relative  humidity   inside   the   cave   is  below  85  %  and   the   wind   velocity   above   the   solution   is   higher   than   0.2   m/s,   evaporation  influences   the   δ18O.   Since   most   caves   have   a   relative   humidity   above   95   %,  evaporative  effects  might  be  of  subordinated  order  (Deininger  et  al.,  2012).      2.3.5  Disequilibrium  fractionation,  effects  and  tests    Dependence  of  calcite  δ18O  and  δ13C  on  drip  interval  Both  the  δ18O  and  δ13C  signals  are  similarly  affected  by  the  drip  interval.  Under  a  slower  drip,  the  residence  time  of  the  water  in  the  thin  film  on  the  surface  of  the  stalagmite   is   longer   and   calcite   deposition   has   more   time.   The   CO2   produced  during  the  calcite  deposition  will  degas  from  the  solution  and  increase  both  the  δ18O   and   δ13C   in   the   solution.   The   δ18O   and   δ13C   of   the   deposited   calcite   will  therefore  increase  with  longer  drip  intervals.      Dependence  of  calcite  δ18O  and  δ13C  on  the  pCO2  gradient  The   rate   of   CO2   degassing   is   a   function   of   the   CO2   gradient   between   the   drip  water,  which  is  largely  determined  by  the  degree  of  biological  respiration  in  the  soil,  and  the  cave  atmosphere,  which  is  controlled  by  (seasonal)  cave  ventilation  (Mühlinghaus  et  al.,  2007;  Baldini  et  al.,  2008;  2009;  Scholz  et  al.,  2009).  Large  pCO2  gradients  between  drip  water  and  cave  atmosphere,  increase  the  amount  of  CO2  degassing   from   the   solution.   However,   the   time   constant   for   degassing   of  excess   CO2   from   the   solution   does   not   depend   on   the   difference   between   the  pCO2   of   the   solution   and   cave   air   (Dreybrodt   and   Scholz,   2011).   Degassing   is  always  fast  and  occurs  within  ±  10s  after  entering  the  cave  atmosphere  (Hansen  et  al.,  2013).  More  CO2  degassing  causes  the  δ18O  and  also  the  δ13C  composition  of   the   DIC   to   increase,   as   lighter   CO2   is   preferentially   lost   from   the   solution.  Calcite  deposition  onsets  before  CO2  hydration  and  hydroxylation  processes  can  re-­‐equilibrate   the   decreased   δ18O   value   of   the   HCO3-­‐   in   the   solution.   CO2  degassing  has  a   similar  effect  on   the  δ18O  and  δ13C  signals.  Stronger  degassing,  due   to   a   larger   pCO2   gradient,   will   cause   heavier   δ18O   and   δ13C   signals   in   the  solution  and  consequently  in  the  deposited  calcite.    Drip  intervals  approaching  zero  (i.e.  very  fast  drip  rates)  represent  an  exception  by   showing   no   relation   between   the   calcite  δ18O   and  δ13C   and   drip   interval   or  pCO2  gradient  variations.  The  isotope  ratio  of  the  solution  on  the  surface  of  the  stalagmite  is  close  to  the  isotope  ratio  of  the  impinging  drops  (Mühlinghaus  et  al.,  

Page 41: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  31  

2009).   Calcite   deposited   under   such   dip   sites   represents   fractionation   under  conditions  of   isotopic  equilibrium  (Mühlinghaus  et  al.,  2009).   In  such  cases   the  δ18O  and  δ13C  values  of  the  precipitated  calcite  are  determined  by  their   isotope  fractionation  factor  (which  is  temperature  dependent  for  the  δ18O).      Different   tests   can  be   carried  out   to   investigate   the  presence  of  disequilibrium  processes.  A  Hendy-­‐test,  where  several  samples  are  drilled  along  a  single  growth  layer,   may   indicate   disequilibrium   processes.   Better   than   the   Hendy-­‐test   is   to  calculate   the   expected   δ18O   and   δ13C   values   based   on   the   theoretical  “equilibrium”   water-­‐calcite   fractionation   factors   for   C   and   O   (Kim   and   O'Neil,  1997),   the   present-­‐day   cave   temperature   and   the   isotopic   compositions   of  seepage  waters  and   to  compare   these  results  with   the  measured  δ18O  and  δ13C  values.   However,   the   problem  with   this   test   is   that   cave   conditions  may   have  changed  in  the  past.  Present-­‐day  equilibrium  deposition  of  the  calcite  in  the  cave  is  certainly  not  a  guarantee  for  a  similar  condition  in  the  past.        Speleothems  growing  in  most  natural  cave  environments  often  display  δ18O  and  δ13C  values  that  are  higher  than  predicted  by  calculations  starting  from  the  drip  water  δ18O  and  δ13CDIC  composition  (Mickler  et  al.,  2006;  Tremaine  et  al.,  2011).  Only   a   number   of   cave   environments   such   as   the   fast   growing   Proserpine  stalagmite   at   the   Han-­‐sur-­‐Lesse   cave   (Van   Rampelbergh   et   al.,   2014;   Van  Rampelbergh   et   al.,   in   review),   drip   sites   in   Bunker   cave   in  western   Germany  (Riechelmann  et  al.,  2013),  Soreq  Cave  in  Israel  (BarMatthews  et  al.,  1996)  or  in  Scotland   (Fuller   et   al.,   2008)   have   been   shown   to   deposit   speleothem   calcite  according  to  the  equilibrium  fractionation  rules.  The  difficult  match  between  the  calculated  values  and  the  values  measured   in  natural  cave  environments  result  in  a   large  debate  on  whether  speleothems  can  reflect   isotopic  equilibrium.  The  used   fractionation   factors   to   estimate   if   calcite   is   deposited   in   isotopic  equilibrium   are   mostly   determined   by   artificial   (laboratory   or   theoretical)  setups   and   may   not   be   absolutely   similar   in   natural   cave   environments.  Tremaine  et  al.  (2009)  estimated  a   ‘speleothem  δ18O  fractionation  factor’  based  on  a   large  collection  of  speleothem  data.  Their  suggested  fractionation  factor   is  smaller   than   the   fractionation   factors   derived   from   theoretical   and   laboratory  results  (such  as  O’  Neil  et  al.  1969  or  Horita  &  Clayton,  2007)  indicating  that  the  theoretical  and  mathematical  results  tend  to  overestimate  the  natural  conditions.  A  larger  amount  of  detailed  cave  monitoring  studies  is  needed  to  provide  better  insights  in  the  water-­‐calcite  fractionation  factors  for  natural  cave  environments.  Other  than  the  discussion  on  the  fractionation  factor,  the  different  processes  and  timing  of  these  processes  affecting  the  drip  water  DIC,  which  largely  determine  the   speleothem   isotopic   composition,   are   still   not   fully   understood.   Recent  findings  on  the  longer  isotopic  exchange  times  between  HCO3-­‐  and  H2O  and  their  relation   with   speleothem   calcite   formation   (Dreybrodt   and   Scholz,   2011;  

Page 42: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 32  

Deininger  et  al.,  2012;  Hansen  et  al.,  2013)  brought  new  insights   in   the  correct  impact   of   disequilibrium   processes   caused   by   longer   residence   times   of   the  water  on  the  speleothem  surface  or  pCO2  gradients.  More  detailed  studies  on  the  isotopic  processes  in  the  solution  layer  from  which  calcite  is  formed  and  on  the  timing   of   these   processes   will   provide   better   insights   how   disequilibrium  processes   may   affect   the   isotopic   composition   of   the   drip   water   DIC   and  consequently  the  deposited  calcite.      2.3.6  Prior  Calcite  Precipitation  (=  PCP)    Prior  calcite  precipitation  (PCP)  is  a  common  process  occurring  in  karst  aquifers  (Fairchild   et   al.,   2000;   Fairchild   et   al.,   2006;   Verheyden   et   al.,   2008b;  Riechelmann   et   al.,   2011;   Tremaine   and   Froelich,   2013;   Rutlidge   et   al.,   2014).  When   downward   percolating   water   encounters   a   zone   with   lower   pCO2,  degassing  occurs  and  calcite  can  precipitate  within  the  epikarst  before  reaching  the  cave.  During  drier  periods,  PCP  is  enhanced  as  aerated  zones  increase  in  the  aquifer  and  residence  time  of  the  water  becomes  longer  (Fairchild  et  al.,  2000).      The  CO2-­‐degassing  from  the  solution  within  the  epikarst  will  cause  the   isotopic  composition  of  the  HCO3-­‐   in  the  solution  to  increase  due  to  the  removal  of   light  oxygen   and   carbon   isotopes.   However,   buffering   effects   of   hydration   and  hydroxylation  will   re-­‐equilibrate   the   increased   δ18O   of   the   DIC   in   the   solution  with  the  unaffected  δ18O  of  the  water.  This  re-­‐equilibration  process  can  take  up  to   several   days   (Dreybrodt   and   Scholz,   2011).   Only   water   that   stayed   in   the  epikarst  for  several  days  after  being  affected  by  PCP  will  display  δ18O  values  that  are   in  equilibrium  with   the   initial   rainwater   signal.  PCP  will   thus  always  cause  increased  δ13C  values  of   the  DIC   in   the  drip  water  while   the   effect   on   the  δ18O  depends  on  the  residence  time  of  the  water  after  PCP.  If  the  residence  time  after  PCP  is  longer  than  several  days,  the  drip  water  δ13C  will  increase  while  the  δ18O  signal   represents   the   initial   water   composition   before   PCP.   Comparing   the  isotopic  signal  of  soil  and  drip  water  can  indicate  the  occurrence  of  PCP.  Seasonal  variations   in   PCP   can   cause   seasonal   variations   in   the   δ13C   and   under   short  residence   times  of   the  water  also   in   the  δ18O   signals  of   the  drip  water  and   the  formed  speleothem  (Van  Rampelbergh  et  al.,  2014).                  

Page 43: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  33  

The   most   visible   expression   of   PCP   is   the   formation   of   a   stalactite   above   a  stalagmite.   During   this   form   of   PCP,   the   drip  water  will   first   deposit   stalactite  calcite  before  forming  a  stalagmite.  Deposition  of  stalactite  calcite  increases  both  the  δ18O  and  δ13C  values  of   the  DIC   in   the   remaining  solution.  This   isotopically  heavier   solution   will   fall   on   the   top   of   the   stalagmite   and   start   to   deposit  stalagmite  calcite.  The   formation  of   the  stalagmite  calcite  will   start   long  before  the   increased   δ18O   of   the   DIC   in   the   solution   can   re-­‐equilibrate  with   the   large  reservoir  of  unaffected  (=lighter)  δ18O  of  the  H2O  in  the  solution  (Dreybrodt  and  Scholz,  2011).  During  stalactite-­‐PCP,  both  the  δ13C  and  the  δ18O  value  of  the  DIC  in   the  solution  and  consequently   the  deposited  stalagmite  calcite  are   increased  compared  to  the  composition  of  the  vadose  water.    PCP  can  thus  be  detected  by  increased  δ13C  signals  and  unaffected  δ18O  for  long  residence  times  or  by  both  increased  δ18O  and  δ13C  signals  under  short  residence  times  of  the  water  in  the  epikarst.    However,  since  multiple  other  processes  than  PCP  can  cause  increased  δ13C  (and  δ18O),  combination  of  stable  isotope  records  with   trace   elemental   records   (mostly   Mg   and   Sr,   but   also   Ba   can   be   used)   is  advised   to   indicate   the   process   of   PCP.   Co-­‐varying   increased   δ13C   and   trace  elemental  ratios  with  distribution  coefficients  smaller  than  1  (e.g.  Mg  and  Sr)  are  the  mostly  used  tool  to  identify  PCP  (see  2.4.1).    2.3.7  Summary  δ18O  and  δ13C  signals  in  speleothems    As  discussed  in  the  previous  sections,  numerous  effects,  which  act  independently  or  influence  each  other  in  a  positive  or  negative  manner,  determine  the  final  δ18O  and   δ13C   composition   of   the   deposited   speleothem   calcite.   No   two   caves   are  similar;   therefore   interpreting   the  measured   speleothem  δ18O   and  δ13C   signals  requires   a  deep  understanding  of   the   studied   cave   system   to   link   the  obtained  data  to  climate  parameters.  Furthermore,  the  higher  the  temporal  resolution  of  the  δ18O   and  δ13C   time   series,   the  more   complex   the   interactions   become.   The  required   knowledge   of   the   cave   system   is   obtained   by   cave   monitoring  programs,  together  with  calcite  production  experiments.    

Page 44: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 34  

 Figure  6.  Oxygen  isotopes  in  speleothems      Figure   6   summarizes   the   different   factors   determining   the   speleothem   δ18O  signal.  For  semi-­‐arid  tropical  regions  such  as  Socotra  Island  (see  Chapter  3),  the  rainwater   δ18O   is   driven   principally   by   the   amount   effect.   Speleothems   calcite  growing  in  such  regions,  -­‐   if  not  disturbed  by  other  factors  such  as  for  example  disequilibrium  effects  or  mixing  of  water  in  the  epikarst  -­‐,  is  expected  to  reflect  changes  in  wetter  versus  drier  conditions  (Fairchild  and  Baker,  2012).   In  semi-­‐arid  tropical  regions,  speleothems  deposited  in  equilibrium  with  their  drip  water  should  roughly  vary  around  an  average  of  -­‐4  ‰.  In  temperate  regions,  such  as  in  Belgium  (Chapter  4  and  5),  rainwater  δ18O  averages  -­‐7  ‰  and  is  influenced  both  by  the  temperature  and  by  the  amount  effect.  In  such  climate  regions  speleothem  calcite   deposited   in   equilibrium   with   the   drip   water   at   an   average   cave  temperature  of  12°C  is  expected  to  be  ∼  1  ‰  heavier  compared  to  the  drip  water  δ18O.  Belgian  speleothems  are  thus  expected  to  display  values  around  -­‐6  ‰.      

Page 45: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  35  

 Figure  7.  Carbon  isotopes  in  speleothems    Figure  7  summarizes  the  different  factors  to  be  considered  when  reconstructing  climate  based  on  δ13C  signals  in  speleothems.  The  carbon  fractionation  factor  is  only   slightly   temperature   dependent   and   provided   that   deposition   occurs   in  isotopic  equilibrium,  speleothem  calcite  mainly  reflect  the  δ13C  of  soil  CO2,  which  is  controlled  by  the  vegetation  above  the  cave.  For  dominant  C3  vegetation  (δ13C  between  -­‐32  ‰  and  -­‐25  ‰),  soil  CO2  has  a  δ13C  value  between  -­‐28  ‰  and  -­‐21  ‰.   Considering   the   enrichment   factor   between   CO2   in   the   soil   and  DIC   in   the  water  (95%  HCO3-­‐  at  pH≈  7)  of  ∼  10  ‰  and  the  enrichment  factor  between  DIC  in  the  water  and  the  CaCO3  in  the  speleothem  of  0.2  ‰  to  0.9  ‰  (for  10  °C  and  25   °C   resp.),   speleothem   calcite   -­‐   without   bedrock   carbon   contamination-­‐  displays   δ13C   values   between   -­‐18   ‰   and   -­‐11   ‰.   For   C4   vegetation   (δ13C  between  -­‐14‰  and  -­‐10  ‰),  speleothem  calcite  displays  a  δ13C  of  between  0  ‰  and   +4  ‰.   Dissolution   of   bedrock   (δ13C   ≈   0  ‰)   increases   the   epikarst  water  δ13CDIC   composition   and   consequently   the   speleothem   δ13C.   For   speleothems  growing  under  a  C3-­‐vegetation  covered  soil  and  with  a  15  %  bedrock  input,  the  speleothem  δ13C  will   vary  between   -­‐15  ‰  (-­‐18  ‰  *  85%  +  0‰*15%)  and   -­‐9  ‰.   Under   a   C4-­‐vegetation   covered   soil   and   a   15   %   bedrock   input,   the  speleothem  δ13C  will  vary  between  0  ‰  and  5  ‰.  

Page 46: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 36  

 2.4  Mg  and  Sr  in  speleothems    Trace   elements   represent   a   large   portion   of   the   measurable   proxies   in  speleothems  and  have   increasingly  provided  useful   information  on  speleothem  based   paleoclimate   reconstructions.   They   can   be   used   either   in   chronology,  where   trace  elements  vary  rhythmically   (Smith  et  al.,  2009)  or  as  paleoclimate  proxy    (Borsato  et  al.,  2007;  Fairchild  and  Treble,  2009;  Tremaine  and  Froelich,  2013;   Van  Rampelbergh   et   al.,   2013;   Orland   et   al.,   2014;   Rutlidge   et   al.,   2014;  Verheyden  et  al.,  2014).      Most   studies   of   speleothems   focus   on   elements   that   form   divalent   cations   in  solution  and  which  substitute  for  Ca  in  the  carbonate  crystal  lattice,  particularly  Mg   and   Sr   (Fairchild   and   Treble,   2009).   In   this   study,   we   focused   on   the   two  trace   elemental   ratios   of   Mg/Ca   and   Sr/Ca.   For   these   species,   the   Mg   and   Sr  content  measured   in   the   stalagmite   depend   on   their   ratio   in   the   precipitating  solution  and  their  distribution  coefficient.      2.4.1  The  Mg  and  Sr  distribution  coefficients        As  a  mineral  grows   from  an  aqueous  solution  and   if  equilibrium  is  maintained,  trace  elements  will  partition  between  the  two  phases  in  a  characteristic  manner.  The   fundamental   law   that  controls   the  distribution  of  a   trace  element  between  coexisting   phases   is   usually   referred   as   the   Berthelot-­‐Nernst   distribution   law  (McIntire,   1963).   According   to   this   law   and   at   equilibrium,   the   ratio   of   the  concentration   of   the   trace   component   in   the   solid   to   its   concentration   in   the  liquid  is  a  constant.  A  more  convenient  form  of  the  distribution  law  refers  to  the  ratio  of  the  trace  element  (Me)  to  the  major  element  (Ca  in  the  case  of  calcite):    D  (distribution  coefficient)  =  (cMe/Ca)solid  /  (cMe/Ca)solution    with  cMe/Ca  being  the  concentration  ratios  of  the  trace  element  (in  this  study  Sr  and  Mg)  to  calcium  on  either  a  weight  or  molar  basis.      An   overview   of   the   average   distribution   coefficients   for   Mg   and   Sr   between  water  and  calcite  derived  from  laboratory  experiments  or  cave  observations  are  listed   in   Table   3.   Laboratory-­‐based   inorganic   calcite   precipitation   experiments  similar  to  natural  cave  precipitates  have  been  performed  under  well-­‐controlled  solution   chemistry   conditions   to   determine   trace   element   distribution   into  calcite  (Katz  et  al.,  1972;  Pingitore  and  Eastman,  1986;  Huang  et  al.,  2001),  and  others).  Huang  and  Fairchild  (2001)  examined  trace  elements  under  variable  but  controlled   [Ca2+],   [HCO3-­‐],   and   pCO2   conditions   that   closely   mimicked   natural  

Page 47: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  37  

cave   systematics.   Distribution   coefficients   determined   by   modern   natural  speleothem   calcite   farmed   under   active   drips   in   natural   cave   systems   are  starting   to   increase   (Holland   et   al.,   1964;   Gascoyne,   1983;   Huang   et   al.,   2001;  Stern   et   al.,   2005;   Tremaine   and   Froelich,   2013)   but   often   display   different  values   (Tremaine  and  Froelich,  2013).  Generally,   large  differences  between   the  distribution   coefficients   of   the   different   used   approaches   and   cave   sites   are  observed.    All  results  clearly  show  partition  coefficients  for  water-­‐calcite  solutions  smaller  than  1.  This   implies   that   the  Mg/Ca   and  Sr/Ca   ratios   of   the   calcite   are   smaller  than   those  of   the  precipitating  solution.   Inversely,  when  a  solution  precipitates  calcite,   its  Mg/Ca  and  Sr/Ca  ratios  increase  (this   is  of  major  importance  for  the  process  of  PCP  discussed  in  2.3.6).      A   number   of   laboratory   and   cave   studies   have   demonstrated   a   temperature  dependence   on   DMg,   which   can   be   of   interest   for   paleo-­‐temperature  reconstructions   (Katz,   1973;   Gascoyne,   1983;   Huang   et   al.,   2001).   At   higher  temperatures,   more   Mg   is   incorporated   in   the   calcite.   Gascoyne   (1983,   1992)  suggested  that,  if  the  Mg/Ca  of  the  precipitating  solution  remained  stable,  Mg/Ca  ratios   could   be   used   a   temperature   proxy.   Huang   and   Fairchild   (2001)  performed  an  experimental  investigation  in  cave  like  setups  to  try  to  determine  the   temperature   dependence   of   the   DMg   value.   They   obtained   a   rather   small  temperature  dependence  of  about  0.006/°C  in  the  range  of  7  to  15  °C.  However,  up   until   now   no   speleothem   Mg/Ca   ratios   have   successfully   been   able   to  reconstruct   temperature   variations.   The   reason   is   that   cave   temperature  variations   are   often   too   small   to   induce   significant   (=   detectable)   Mg/Ca  variations   in   stalagmites.   Indeed,   the   temperature   signals   are   typically  masked  by  noise   from  other   factors  with  much   larger  variation,   such  as   changes   in   the  Mg/Ca–ratio   of   the   precipitation   solutions.   The   use   of   Mg/Ca   in   speleothem  studies   has   consequently   largely   shifted   from   temperature   reconstructions   to  interpreting   variations   in   terms   of   hydrological   changes   (Fairchild   and   Treble,  2009).        

 

Page 48: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 38  

   

Table   3.   Average   distribution   coefficients   for   Mg   (DMg)   and   Sr   (DSr)   from   the  literature   for   inorganic   calcites  determine  by   laboratory   experiments   and   cave  observation  results  (adapted  from  Tremaine  et  al.,  2013  and  references  herein).    Laboratory   and   speleothems   studies   indicate   a   growth   rate   dependency   of   the  DSr  (Pingitore  and  Eastman,  1986;  Huang  et  al.,  2001;  Treble  et  al.,  2003;  2005b;  Gabitov  and  Watson,  2006).  However,   the  growth  rate  needs   to  be  higher   than  0.5  mm/y  to  affect  the  DSr  (Gabitov  and  Watson,  2006).  The  fact  that  DSr  is  only  influenced  for  growth  rates  higher  than  0.5  mm/y  relates  to  the  fact  that  Sr2+  can  be  incorporated  in  the  lattice  sites  to  replace  Ca2+  but  also  in  crystal  defect  site  (non-­‐lattice   sites)   (Pingitore   and   Eastmann,   (1986)).     Such   non-­‐lattice   sites  

Page 49: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  39  

become   more   frequent   under   high   growth   rates.   However,   increased   Sr-­‐concentrations  have  been  observed  together  with  increased  speleothem  growth  rates   that  were  smaller   than  0.5  mm/y  (Frisia  et  al.,  2003;  Treble  et  al.,  2003).  Fairchild   and  Treble   (2009)   suggested   an   alternative,   and  better   suited   theory  for  the  relation  between  Sr-­‐concentration  and  growth  rates,  which  explains  why  Sr-­‐concentration   can   covary   with   growth   rates   lower   than   the   previously  suggested   0.5   mm/year   limit   by   Gabitov   and   Watson   (2006).   Fairchild   and  Treble  (2009)  suggest  that  the  good  correspondence  between  Sr-­‐concentrations  (and  Ba,  Na  concentrations)  and  growth  rate  is  related  to  growth  kinetics  rather  than   to   growth   rate.   Growth   kinetics   are   better   suited   to   refer   to   changes   in  crystal  morphology  and/or  growth  mechanisms,  usually  associated  with  growth  rate  changes.  Increased  kinetics  cause  augmentation  of  crystal  defects  and  leads  to  more  Sr2+  being   incorporated   in   the  calcite.  However,   growth  kinetics   is  not  the  only  factor  that  determines  the  Sr-­‐concentration  in  the  calcite.  The  presence  of   other   competing   elements   in   the   water   such   as   Ba   and   Na   can   lower   the  amount  of  Sr   that   is   introduced   in   the  crystal  system  since  these  elements  also  occupy   the   non-­‐lattice   sites.   To   determine   which   factor   causes   the   Sr-­‐concentration  variations,  Sr  needs  to  be  measured  together  with  the  ‘competing’  elements  Ba  and  Na.    2.4.2  The  Mg/Ca  and  Sr/Ca  ratio  in  the  precipitating  solution    Trace  elements   in   speleothems  can  originate   from  a  variety  of   sources   such  as  aeolian   particles,   dry   and   wet   atmospheric   deposition,   bedrock,   superficial  sediment  deposits  and  inorganic  soil  constituents,  and  elements  recycled  via  soil  biota  (Fairchild  and  Treble,  2009).      Atmosphere  is  only  a  subordinate  source  for  trace  elements.  However,  Sr-­‐isotope  signals  have  been  related  to  the  supply  of  aeolian  dust  in  speleothems  (Goede  et  al.,  1998;  Bar-­‐Matthews  et  al.,  1999;  Zhou  et  al.,  2009;  Hori  et  al.,  2013).  Sea-­‐salt  aerosols   in  wet  atmospheric  deposition  can  be  an   important   source   for  certain  trace   species   in   karst   waters,   sometimes   including   Mg   and   Sr   where   bedrock  supply   is   limited   (Bar-­‐Matthews   et   al.,   1999;   Fairchild   et   al.,   2000).   In   arid  environments   a   large   input   of   soil   derived   Mg   from   aeolian   input   or  evaporational   salts   formed   during   the   dry   season   can   also   form   a   major   Mg  source  in  the  drip  water  (Rutlidge  et  al.,  2014).    Despite   exceptions,   the   primary   source   of   calcium  Mg   and   Sr   is   the   carbonate  bedrock   (calcite,   dolomite)   and   overlying   soil,   including   bedrock   fragments   in  the  soil  (Fairchild  and  Treble,  2009).  However,  exceptions  occur  such  as  in  NW  Scotland   where   the   high   Sr-­‐concentrations   in   calcite   speleothems   arise   from  localized   veins   associated   with   a   thin   igneous   sill   in   the   overlying   Sr-­‐poor  

Page 50: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 40  

dolomite   (Roberts   et   al.,   1998).    When   clay  minerals   are   present,   Sr   is   usually  supplied  from  both  carbonates  and  silicates.      In  the  case  of  a  calcite  and  dolomite  composed  epikarst,  Mg/Ca  and  Sr/Ca  ratios  will  arise  from  their  dissolution.  The  slow  dissolution  of  dolomite  (main  source  of  Mg)  will  cause  the  water  Mg/Ca  ratio  to  be  smaller  than  the  bulk  ratio  in  the  bedrock.  Raised  Mg/Ca  ratios  compared  with  bulk  carbonate  were  first  thought  to  be  linked  to  incongruent  dissolution  (Fairchild  et  al.,  2000).  A  higher  residence  time  of  the  water  in  the  epikarst  due  to  slow  water  flow  or  drier  conditions  were  suggested   to   lead   to   higher   Mg/Ca   ratios   by   enhanced   dolomite   dissolution  (Roberts  et  al.,  1998;  Fairchild  et  al.,  2000;  Tooth  and  Fairchild,  2003;  Musgrove  and  Banner,   2004).  However,   this  mechanism   requires   dissolution   of   dolomite  into  solutions  already  saturated  for  calcite.  The  process  of  measurably  enhancing  Mg  contents  is  very  slow  and  is  likely  to  require  long  residence  times  of  months  to   years.   Fairchild   et   al.   (2006)   suggested   a  more   straightforward  hydrological  routing  model   for  which   there   is   strong   evidence   in  Mesozoic   aquifers.   In   this  model,  the  residence  time  effect  is  related  to  an  enhanced  contribution  of  Mg  rich  waters   from   low-­‐permeability  parts   of   the   aquifer   at   low   flow   (Fairchild   et   al.,  2006).   Enhanced   Mg   and/or   Sr   enrichment   during   annual   low   flows   can   be  related  to  higher  contents  of  dolomite  (for  Mg)  or  aragonite  (for  Sr)  associated  with  clay  (Fairchild  et  al.,  2006).    High  Mg/Ca   and   Sr/Ca   ratios   can   arise   from   preferential   leaching   form   newly  created  calcite  surfaces  (McGillen  and  Fairchild,  2005).  Selective  leaching  is  used  for   the   preferential   removal   of   Mg   and/or   Sr   from   an   homogenous   bedrock  phase  (Tremaine  and  Froelich,  2013).  However,  selective  leaching  of  calcite  does  not   occur   in   nature   in   older   limestone   beds   (Palmer   and   Edmond,   1992)  indicating   that   ‘recent’   processes   must   create   new   leaching   surfaces.   Winter  freezing  creating  fresh  surfaces  can  generate  such  new  leaching  surfaces  and  is  probably  the  case  for  the  variations  in  the  Ernesto  Cave  (Fairchild  et  al.,  2000).  The  process  of  selective  leaching  can  also  arise  where  storage  occurs  during  the  dry   season   as   sulphate   or   chloride   salts,   where   there   is   fresh   supply   of   clay  minerals  with  other  exchangeable  ions,  or  where  other  more  soluble  phases  such  as  aragonite  are  present  (Fairchild  et  al.,  2006).  Rutlidge  et  al.  (2014)  illustrated  the  importance  of  such  salt  depositions  in  the  soil  during  the  dry  season  in  arid  environments.  They  pointed  out  that  the  Mg-­‐input  by  soil  salts  can  be  as  large  as  Mg-­‐input  by  bedrock  dissolution.      The  process  mostly  used  to  interpret  increased  Mg/Ca  and  Sr/Ca  values  is  Prior  Calcite   Precipitation   or   PCP   (see   2.3.6).   This   process   was   first   suggested   by  Holland  et  al.  (1964)  and  further  illustrated  in  many  later  studies  to  be  the  cause  of  increased  Mg/Ca  and  Sr/Ca  ratios  (Tooth  and  Fairchild,  2003;  McDonald  et  al.,  2004;  Fairchild  et  al.,  2006;  Karmann  et  al.,  2007;  Tremaine  and  Froelich,  2013;  

Page 51: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  41  

Van   Rampelbergh   et   al.,   2013;   Rutlidge   et   al.,   2014).   The   effect   of   calcite  precipitation  is  to  remove  cations  from  the  water  in  the  proportion  in  which  they  are   incorporated   into   calcite   (Holland   et   al   1964).   PCP   causes   trace   elements  with  partition  coefficients  <1,  such  as  those  for  Mg,  Sr  and  Ba  to  become  enriched  in  the  solution  compared  to  calcite.  Trace  elements  with  partition  coefficients  >1,  such   as   is   the   case   for   Zn,   will   become   depleted   in   the   solution   compared   to  calcite  (Fairchild  and  Treble,  2009).      As   discussed   in   2.3.6,   PCP  most   often   causes   increased   drip  water  δ13C   values  under   long  residence  times  of   the  water   in   the  epikarst.  The  main  evidence   for  PCP  is  the  co-­‐variation  of  all  three  Mg/Ca,  Sr/Ca  and  δ13C  in  the  record  (McMillan  et   al.,   2005;   Johnson   et   al.,   2006).   However,   in   monsoon   regions   (such   as  Socotra),   δ18O   signals   reflecting   the   amount   effect   can   also   co-­‐vary   with  increased  Mg/Ca  and  Sr/Ca  ratios  (and  δ13C)  (Cruz  et  al.,  2007;  Van  Rampelbergh  et  al.,  2013).  During  periods  of  increased  rainfall,  the  δ18O  value  of  the  drip  water  lowers  and  more  water  is  pushed  through  the  epikarst  reducing  the  occurrence  of  PCP.  Consequently  lower  Mg/Ca  and  Sr/Ca  (and  δ13C)  correlating  with  lower  δ18O   values   indicate   periods   of   higher   water   excess.   Tremaine   et   al.   (2013)  indicated   that   co-­‐varying  Mg/Ca   and   Sr/Ca   ratios   are   directly   linked   to  water  excess  (precipitation  minus  evapotranspiration);  higher  net  precipitation  causes  PCP   to  decrease.  The  Mg/Ca  and  Sr/Ca  ratios   in  combination  with  δ18O  signals  can   also   be   used   to   determine   changed   rainfall   sources   (Cruz   et   al.,   2007;  Tremaine  and  Froelich,  2013).  A  change  in  rainfall  source  (source  effect)  induces  a  change  in  δ18O  signal  but  no  changes  in  the  Mg/Ca  and  Sr/Ca  ratios.  Large  shifts  in  δ18O  values  not  visible   in   trace  elemental  ratios  are  consequently  suggesting  source  effects  on  rainfall  δ18O  values  (Tremaine  and  Froelich,  2013).      However,   given   the   potential   multitude   of   factors   affecting   Mg   and   Sr  concentrations   including   aeolian   input,   selective   leaching,   hydrological   routing  causing   mixing   with   other   fluids   or   bedrock   with   different   Mg   and   Sr  concentration  and  prior  calcite  precipitation,  caution  should  be  exercised  when  interpreting  Mg/Ca  and  Sr/Ca  ratios  in  terms  of  climate.  This  is  particularly  the  case  when  no  long-­‐term  drip  water  studies  have  been  carried  out  at  the  cave  site  or  when  the  results  of  such  studies  may  not  be  relevant  due  to  strongly  different  climate  regimes  between  the  present  day  and  past   times  (Fairchild  and  Treble,  2009).      Best  speleothem  samples  for  using  trace  elemental  ratios  to  reconstruct  climate  parameters  are  stalagmites  growing  under  seepage  flow  systems  (Fairchild  and  Treble,   2009)   where     mixing   of   water   with   different   ages   is   not   expected   to  overshadow   the   signal   of   effective   rainfall   and   PCP.   By   contrast,   drip   water  through   crystalline,   fractured   limestones   are   likely   to   be   dominated   by  

Page 52: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 42  

hydrological   routing   (Fairchild   and   Treble,   2009).   Epikarst   thickness  may   also  play  a  role  for  the  rate  at  which  climate  signals  are  brought  to  the  cave  by  trace  elements.  This  effect  is  similar  to  the  effects  on  stable  isotopes  (last  paragraph  of  2.3.1).  Recent  studies  (Galy  et  al.,  2002;  Buhl  et  al.,  2007)  have  indicated  that  Mg-­‐isotopes   (δ26Mg)   could   be   a   tool   for   understanding   whether   source   effects,  hydrologic   routing   or   residence   time   are   dominating   at   a   particular   site.   Sr-­‐isotopes   (87Sr/86Sr)   can   be   used   to   determine   the   source   of   Sr   not   originating  from   the  epikarst.  Bar-­‐Matthews  et   al.   (1999)   interpreted  higher   87Sr/86Sr   in   a  speleothem   record   from   Soreq   cave   (Israel)   to   reflect   increased   input   of   sea  spray   droplets   and   dust   particles   through   the   last   glacial   period.   More   recent  studies  have  also  indicated  the  importance  of  87Sr/86Sr  ratios  in  speleothems  for  paleoclimate  reconstructions  (Zhou  et  al.,  2009;  Hori  et  al.,  2013).    2.5  Climate  reconstructions  from  speleothems      2.5.1  Quantitative  paleotemperature  estimates    An   early   goal   of   speleothems   was   to   reconstruct   absolute   changes   in   mean  annual   temperature  based  on  the   temperature  dependency  of   the  water-­‐calcite  oxygen   fractionation   factor   (Hendy   and  Wilson,   1968;   Thompson   et   al.,   1974).  However,   few   reliable   estimates   have   been   published   because   of   the  considerable   complexity   of   δ18O   in   the   atmosphere,   hydrosphere,   and   cave  environment. To   estimate   quantitatively   past   temperatures   from   speleothem  δ18O   data,   the   δ18O   value   of   the   drip   water   and   the   calcite   at   time   of   calcite  formation   are   required.   Since   only   the   δ18O   of   the   calcite   can   be   measured,  estimating   the   calcite   temperature   formation   remains   difficult.   Different  techniques   have   been   and   are   still   being   developed   to   reconstruct   the   paleo-­‐dripwater  δ18O  signal  either  by  measuring  old  groundwater  with  the  same  age  or  better  by  measuring  the  δ18O  value  of  water  entrapped  in  fluid  inclusions  in  the  speleothem  calcite.    The  δ18O  value  of   the  drip  water  can  be  determined  by  groundwater  of  known  age   and   isotopic   composition   from   near   the   studied   cave.   For   example,   a  stalagmite   from   South   Africa   and   the   δ18O   values   from   a   dated   water   aquifer  were   used   to   determine   the   relative   temperature   difference   between   the   late  glacial   and   today   (Talma   and   Vogel,   1992).   Fluid   inclusions   are   considered   a  better   technique   to  estimate  paleotemperatures   (Matthews  et  al.,  2000;  Dennis  et   al.,   2001;   Zhang   et   al.,   2008).   Although   fluid   inclusions   seemed   promising,  difficulties  associated  with  the  measuring  technique  impede  their  application.  A  first  problem  is  that  the  δ18O  of  the  water  in  the  fluid  inclusion  may  be  expected  to   have   equilibrated   with   the   δ18O   value   of   the   calcite.   However,   δD   values,  cannot   be   affected   by   this   process   and   are   used   to   estimate   the   original   δ18O  

Page 53: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  43  

values   of   the   water   using   the   meteoric   water   line.   Analytical   difficulties   in  measuring   the   δD   values   mainly   due   to   fractionation   have   been   reported  (Matthews  et  al.,  2000).  Also,  the  meteoric  water  line  of  today,  used  to  calculate  the   δ18O,   may   be   different   from   the   one   in   the   past,   which   may   result   in  significant  errors  in  deduced  paleotemperatures.  Different  studies  have  focused  on  optimizing   the  measuring   techniques   (Vonhof  et   al.,   2006;  Verheyden  et   al.,  2008a;  De  Cisneros  et  al.,  2011).  Recent  developments  allow  to  measure  water  isotopes   by   cavity   ring   down   spectroscopy,  which   leads   to   better   δ18O   and   δD  measurements  in  fluid  inclusions  (Arienzo  et  al.,  2013)  and  seems  promising  for  future  work  on  the  speleothems.      Estimation  of  past  drip  water  δ18O  with  better  confidence  has  been  done  using  noble  gases   in   fluid   inclusion  (Kluge  et  al.,  2008;  Scheidegger  et  al.,  2010).  The  solubility  of  noble  gases   in  water  depends  on  the  temperature  of   the  water.  By  measuring   noble   gases   content   in   speleothem   fluid   inclusions,   the  paleotemperature  of  the  calcite  formation  can  be  determined.  However,  a  major  problem  with   this   technique   is   the  often  high  contribution  of  noble  gases   from  air  inclusions  that  mask  the  temperature  information  present  in  the  noble  gases  dissolved  in  water-­‐filled  inclusions  (Kluge  et  al.,  2008).      The  clumped  isotope  technique  seems  the  most  promising  in  speleothem  based  paleotemperature  reconstructions  (Eiler,  2007).   It   is  based  on   the   temperature  dependent  preference  for  heavy  nuclides  to  bond  to  each  other,  rather  than  to  a  light   isotope.  For   carbonates,   the   clumped   isotope  paleo-­‐thermometer   is  based  on   the   abundance   of   13C-­‐18O   bonds   in   the   carbonate   lattice   relative   to   that  expected  at  a  random  distribution  of  isotopes  among  all  isotopologues.  The  CO2-­‐mass  47,  which  is  dominantly  13C18O16O,   is  used  as  parameter  to  determine  the  abundance   of   the   13C-­‐18O   bonds.   By   using   a   calibration   curve   (Kim   and  O'Neil,  1997),  the  measured  Delta(47)  provides  a  paleo-­‐temperature  estimation  for  the  formation  of   calcite.  However,   results   form   clumped   isotope  measurements   on  speleothems   have   shown   that   kinetic   effects   during   calcite   formation   (mainly  driven  by  CO2  degassing)  cause  large  offsets  in  the  Delta(47)  (Affek  et  al.,  2008;  Daeron  et  al.,  2011).    Affek  et  al.   (2008)   found  that  a  modern  speleothem  from  Soreq   Cave   (Israel)   yielded   formation   temperatures   8°C   too   high,   which   they  attributed  to  kinetic  effects.  However,  by  assuming  that  this  offset  is  constant  for  older  samples  from  the  same  cave,  they  arrived  at  a  set  of  results  that  are  largely  consistent   with   previous   independent   assessments.   Although   this   is   a   good  results   for  a   first-­‐of-­‐a-­‐kind  study,  a  much  more  stringent   test   for   the  degree  of  kinetic  modification  is  needed  for  this  method  to  become  routine.        

Page 54: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 44  

2.5.2  Semi-­‐empirical  climate  relationships      Because   of   the   difficulties   associated   with   quantitative   paleotemperature  reconstructions  from  speleothem  δ18O  signals,  alternative  approaches  have  tried  to   estimate   paleotemperature   variations   in   a   semi-­‐empirical   way.   Such  paleotemperature  estimations  are  based  on  the  combined  relationship  between  (i)  the  temperature  and  the  δ18O  value  of  rainwater  (dδ18Op/dT)  (δ18Op  being  the  δ18O  of  the  meteoric  precipitation)  and  (ii)  the  effect  of  cave  temperature  on  the  equilibrium   fractionation   associated   with   calcite   precipitation   (dδ18Oct/dT)  (δ18Oc  being  the  δ18O  of  the  speleothem  calcite).  The  dδ18Op/dT  can  vary  between  +0.17   and   +0.9  ‰/1°C,   depending   on   the   geographical   location   (Dansgaard,  1964;   Rozanski   et   al.,   1992;   Mook,   2000;   Schmidt   et   al.,   2007).   Schmidt   et   al.  (2007)   suggested   a   dependence   of   +0.3   ‰/1°C   for   central   Europe.   The  dδ18Oct/dT  has  to  be  estimated  for  the  studied  speleothem  and  can  vary  between  -­‐0.18‰/°C   to   -­‐0.25‰/°C   (Lachniet,   2009).   Depending   on   the   value   of   both  factors  the  net  effect  can  be  a  positive  or  a  negative  link  between  the  speleothem  δ18O  and  the  external  temperature.  Lauritzen  an  Lundberg  (1999)  calculated  the  dδ18Op/dT   of   a   Norwegian   speleothem   using   independently   derived   estimated  mean   annual   temperatures   of   today   and   of   the   Little   Ice  Age   (1300-­‐1800  AD).  Another  successful  temperature  reconstruction  has  been  carried  out  by  Mangini  et   al.   (2005)  using   a   speleothem   form   the  Austrian  Alps.  They   suggested   a  net  effect  of  -­‐0.22‰/°C  between  the  speleothem  δ18O  and  the  external  temperature.  However,  such  estimations  remain  difficult  and  fragile  since  the  dδ18Op/dT  may  vary   over   time.   Also,   other   effects   (such   as   source   effect,   amount   effect   or  disequilibrium   effects,   discussed   in   section   2.3)   can   cause   variations   in   the  speleothem  δ18O  that  are  not  related  to  temperature.    A   similar   approach   can   be   carried   out   to   reconstruct   changes   in   precipitation  intensity  from  speleothem  δ18O  values  that  reflect  the  amount  effect,  and  that  are  often   growing   in   tropical   or   sub-­‐tropical   environments.   To   do   this,   the  precipitation  dependency  of  the  δ18O  (dδ18Op/dP)  needs  to  be  known  (dP  being  the   variation   in   amount   of   precipitation).   However,   only   a   few   quantitative  applications   of   a   dδ18Op/dP   relation   to   speleothem   δ18O   variations   have   been  published   such   as   for   the   eastern   Mediterranean   region   (Bar-­‐Matthews   et   al.,  2003)  or  the  Peruvian  Amazonia  (van  Breukelen  et  al.,  2008).  Such  calibrations  have   been   rarely   exploited   for   the   very   simple   reason   that   the   dδ18Op/dP  relationship  may  not  have  been  constant  over  time  (Lachniet,  2009).  Further,  the  smoothing  effect  in  an  aquifer  due  to  mixing  of  infiltrated  waters  of  varying  ages  would   produce   a   dampened   δ18O   signal   in   a   speleothem,   and   would  underestimate  true  rainfall  variation.        

Page 55: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  45  

2.5.3  Quantification  of  the  isotope  effects  in  the  meteoric  water  cycle.    To  date,   little  work  has   explored   the   spatial   variations   caused  by   the   so-­‐called  isotope  ‘effects’  on  the  δ18O  of  meteoric  waters  (such  as  latitude  effects,  altitude  effect,   continentality  or   source  effect)   as   they  are  preserved   in   the   speleothem  δ18O.  As  speleothem  records  are  becoming  more  widely  used,  such  attempts  are  now  feasible.  For  example,  an  altitude  effect  of  -­‐2  ‰/km  in  soda-­‐straw  stalactite  δ18O   (covering   the   last   100   years)   was   reconstructed   from   New   Zealand  (Williams  et  al.,  1999),  from  stalagmites  from  China  (Kong  et  al.,  2005)  and  in  the  European  Alps  (McDermott  et  al.,  1999).  A  latitude  effect  of  0.3‰/°latitude  was  found   by   Williams   et   al.   (1999)   for   the   late   Holocene   that   is   larger   than   the  0.27‰/°latitude  for  modern  precipitation.  In  southern  Brazil,  a  δ18O  latitudinal  gradient   of   0.81‰/°latitude   was   calculated   during   the   Holocene   near   the  Atlantic   coast   (Cruz   et   al.,   2006).   A   recent   study   of   McDermott   et   al.   (2011)  provided  a   first  evolution  of   the  spatial  and  temporal    gradients   in  δ18O  during  the  Holocene  in  Europe  (Fig.  8).    

 Figure  8.  Map  showing  locations  of  cave  sites  used  to  investigate  the gradients  in  δ18O  of  European  Holocene  speleothems.  Contours  are  based  on  δ18O  values  for  recently  deposited  (zero-­‐age)  calcite  at  these  cave  sites.  The  altitudinal  effect  on  speleothem   δ18O   in   the   region   surrounding   the   Alps   is   clearly   visible   (after  McDermott  et  al.  2011).    

Page 56: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 46  

The   results   indicate   two   major   trends.   Trend   1   (Fig.   8)   shows   decreasing  speleothem   δ18O   values   from   the   oceanic   western   European   margins   (SW-­‐Ireland)   to   continental   Europe   reflecting   a   continental   effect.   Trend   2   (Fig.   8)  indicates   decreased   speleothem   δ18O   values   from   Scotland   northward   to  Scandinavia  reflecting  a  latitudinal  effect.  During  the  Early  Holocene  speleothem  δ18O   values   have   show   that   trend   1   displayed   a   steeper   gradient   compared   to  today.   The   decreased   δ18O   values   of   speleothems   from   the   Alps   indicate   an  altitudinal  effect.    2.5.4  Speleothems  as  tools  to  reconstruct  continental  climates    Important  insights  in  the  evolution  of  past  climates  have  been  retrieved  from  ice  cores   and   ocean   drilling   cores.   Ice   cores   are   limited   to   the   Polar   Regions   and  provide  no  information  on  climate  variations  closer  to  the  equator.  Ocean  cores  are  geographically  well  distributed,  but  they  are  often  homogenized  and  thus  not  able  to  reflect  more  regional  climate  variations,  they  do  not  allow  absolute  dating  nor   reconstructing   recent   climate   variations   at   high   resolution,   such   as   for  example   during   the   last   1000   years.   Although   both   very   useful,   ice   cores   and  ocean   cores   do   not   allow   reconstructing   past   climate   variations   in   temperate,  tropical   and   subtropical   continental   regions.   Furthermore,   reconstructing  continental  climates  is  largely  challenging  due  to  the  numerous  regional  or  even  local   effects.   Climate   variations   on   the   continents   form   a   large   puzzle   with  different   pieces   that   are   connected   and   interact.   To   reconstruct   such   climate  variations,  well-­‐spread  and  sensitive  archives  are  necessary.  Speleothems  fulfill  these   conditions   necessary   to   reconstruct   the   continental   climate   successfully.  Furthermore  they  allow  absolute  dating  of  the  time  series  up  to  the  most  recent  500  000  years.      The  best-­‐known   long-­‐term  speleothem  δ18O  time  series  comes   from  subaqeous  vein-­‐filling   calcite   at   Devils   Hole,   a   tectonic   fracture   in   the   Basin   and   Range  region  of  southern  Nevada,  continuously  covering  the   last  500  ka  (Winograd  et  al.,   1992)   (Fig.   9).   Surprisingly,   the   timing   of   the   glacial   to   interglacial   cycles  suggested   by   the   speleothem,   differed   significantly   from   that   predicted   by  Milankovitch   solar   insolation   variations   (Winograd   et   al.,   1992).   In   particular,  most  attention  has  focused  on  the  differences  in  timing  of  Termination  II,  which  is   predated   by   some   12   ka   compared   to   the   timing   in   ice   core   or   deep-­‐sea  records.   Although   the   scientific   debate   on   the   significance   and   forcing   of   the  Devils  Hole  record  continues  (Herbert  et  al.,  2002;  Winograd,  2002;  Yang  et  al.,  2005;  Winograd   et   al.,   2006),   the   time   series   has   proven   to   capture   long-­‐term  and  well-­‐dated  paleoclimatic  variations.    

Page 57: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  47  

Large  amplitude  North  Atlantic  climate  variations  over  the  last  200  kyr  such  as  the   rapid   warming   associated   with   the   Bølling   Allerød   (B-­‐A),   the   cooling  associated  with   the   Younger   Dryas   (YD),  millennial   variability   associated  with  Dansgaard-­‐Oescher   (D/O)  events   (1   to  2  kyr  cyclicity)   (Dansgaard  et  al.,  1993)  and   iceberg  discharge  events  known  as  Heinrich  Events  (Bond  et  al.,  1992)  are  recorded   in   speleothems.  Comparing   speleothem   time   series  with   ice   core   and  deep-­‐sea  records  provides  absolute  ages  on  these  events  and  indicate  how  these  events   are   manifested   on   the   continents   in   the   different   climate   regions.  Stalagmites   from   Hulu   Cave   (China),   covering   the   last   160   kyr,   located   in   the  Asian  Monsoon  region  were  the  first  to  demonstrate  a  millennia-­‐scale  monsoon  response   associated   with   the   D/O   events   (Wang   et   al.,   2001)(Fig.   9).   The  teleconnection  between  the  North  Atlantic  and  the  Asian  Monsoon   implies   that  warmer   periods   in   Greenland   correspond  with   lower   speleothem   δ18O   values,  which   relates   to   both   the   amount   effect   and   the   seasonal   variations   in  winter  versus  summer  rainfall.  A  similar  effect  on  the  monsoon  rainfall  is  also  recorded  in   the  δ18O  values  of   stalagmites   collected   from  Moomi   cave  on  Socotra   Island,  Yemen   (Burns   et   al.,   2003;   Shakun  et   al.,   2007)   confirming   the  warm/wet   and  cold/dry   teleconnection   between   the   North   Atlantic   and   the   Asian   Monsoon  region.  The  impact  of  D/O  events  and  the  Bølling-­‐Allerød  is  also  recorded  in  the  δ18O  signals  of   speleothems   from  the  European  Alps   (Spötl  and  Mangini,  2002;  Spötl  et  al.,  2002)  and  Lebanon  (Bar-­‐Matthews  et  al.,  1997;  Kaufman  et  al.,  1998;  1999;  2000;  Ayalon  et  al.,  2002;  2003)  and  in  both  the  δ18O  and  δ13C  signals  of  western  European  speleothems  (Genty  et  al.,  2003)  and  Turkey  (Fleitmann  et  al.,  2009).   The   study   of   Genty   et   al.   (2003)   offers   perhaps   the   best   available  chronological  control  on  the  timing  and  occurrence  of  such  events,  and  provides  an   improved   chronological   framework   for   the   GRIP   and   GISP   ice-­‐cores  (McDermott,   2004).   Particularly   impressive   is   the   well-­‐dated   composite   δ18O  record   covering   the   past   185   ka   based   on   21   speleothems   from   Soreq   Cave  (Israel),   which   allowed   to   establish   a   link   between   the   oceanic   realm   and  continental  climate  in  the  Levant  region.      (Bar-­‐Matthews  et  al.,  1997;  Kaufman  et  al.,  1998;  1999;  2000;  Ayalon  et  al.,  2002;  2003)  (Fig.  9).  The  Soreq  δ18O  record  appears   to   reflect   dominantly   2   effects:   (i)   changes   in   the   δ18O   of   the   oceanic  vapor  source  and  (ii)  the  amount  effect.    Increased  δ18O  values,  are  interpreted  to  reflect   cooling   forced   by  Heinrich   events   and   during   the   last   glacial  maximum  (Bar-­‐Matthews  et  al.,  1999),  although  some  of   the  Heinrich  events   in   the  North  Atlantic   do   not   appear   to   have   caused   δ18O   events   in   the   Soreq   Cave  speleothems.  The  timing  of  low  δ18O  values  coincided  with  regional  wet  periods  over   the   past   180   ka   during  which   the   distinctive   organic-­‐rich   sapropels  were  deposited   in   the   Mediterranean   Sea   (Bar-­‐Matthews   et   al.,   2000;   Ayalon   et   al.,  2002;  Bar-­‐Matthews  et  al.,  2003).      

Page 58: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 48  

   Figure  9.  A  selection  of  speleothem  δ18O  and  ice  core  records  indicate  the  global  climate  teleconnections.  Speleothem  records  from  the  northern  hemisphere  (e.g.  Moomi  Cave,  Hulu  and  Dongge  Caves,  Devils  Hole,  and  Soreq  Cave)  have  a  strong  imprint   of   high-­‐latitude   climates   (Greenland   and   Antarctica),   whereas   low  latitude  records  also  contain  an  imprint  of  precessional  variations  (dashed  thin  lines),  e.g.  in  Brazil  (Botuvera  Cave),  and  China  (Hulu  and  Dongge  Caves).  These  climate   records   demonstrate   a   robust   response   of   atmospheric   circulation   to  ocean–atmosphere–cryosphere   reorganizations.   For   all   records,   up   is   warm  and/or   wet.   There   is   a   clear   antiphasing   between   northern   and   southern  hemisphere   monsoon   records   (Brazil   and   China)   that   reflects   hemispherically  anti-­‐symmetric  precessional-­‐scale   insolation  variations.  The  raw  dataseries  are  shown  by   thin  grey   lines  with  a  5-­‐pt   running  average  (thick  black   line),  except  for   the   Devils   Hole   record,   which   is   the   raw   data,   and   the   older   Moomi   Cave  record  which   is  a  running  average.  See   text   for  references  (after  Lachniet  et  al.  2009).      

Page 59: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  49  

Speleothems  also  provided  evidence  that  monsoon  and  climate  dynamics  in  low-­‐latitude   regions   are   partly   controlled   by   solar   insolation   on   precessional   time  scales.   Records   from   the   Indian   Ocean   and   South   American   monsoon   regions  support  the  connection  between  stronger  summer  insolation  and  a  more  intense  monsoon  (Fig.  9).  In  southern  Oman,  the  intensity  of  the  monsoon  appears  to  be  strongly  linked  to  summer  insolation  over  the  Holocene  period  (Neff  et  al.,  2001;  Fleitmann   et   al.,   2003),   as   do   changes   in   the   strength   of   the   Asian   Monsoon  (Dykoski   et   al.,   2005)   where   decreased   monsoon   rainfall   is   linked   to   lower  northern   hemisphere   summer   insolation.   In   the   southern   South   America,  subtropical   stalagmites   show   a   pronounced   Holocene   increase   in   monsoon  intensity   that   coincides   with   an   increase   in   southern   hemisphere   summer  insolation   (Cruz   et   al.,   2005;   Wang   et   al.,   2006;   van   Breukelen   et   al.,   2008).  Brazilian  stalagmite  δ18O  show  a  strong  inverse  relation  with  the  Asian  Monsoon  intensity  (Fig.  9),  which  is  interpreted  to  reflect  an  interhemispheric  antiphasing  of  tropical  rainfall  attributed  to  variations  in  the  north-­‐south  position  of  the  ITCZ  (Wang  et  al.,  2006).      Shorter-­‐scaled  events  such  as  the  8.2  kyr  event,  which  caused  global  cooling  due  to  the  drainage  of  the  glacial  Lake  Agassiz  in  North  America  are  also  recorded  in  speleothems.   European   speleothems   registered   this   period   as   generally   cold  (Verheyden  et  al.,  2014),  while  monsoonal  speleothems  recorded  a  decrease   in  monsoon  rainfall  (Fleitmann  et  al.,  2003;  Dykoski  et  al.,  2005).  The  8.2  kyr  event  is  well  documented  in  a  laminated  speleothem  from  Monsoonal  East  Asia  (Liu  et  al.,  2013)  and  indicates  strongly  decreased  monsoonal  precipitation  despite  the  relatively  warm  conditions  that  caused  the  event.  Through  layer  counting  Liu  et  al.  (2013)  indicated  that  the  8.2  kyr  event  lasted  for  150  year  with  peak  drought  over   a   period   of   68   years.   They   suggest   that   the   Northern   Hemisphere-­‐Monsoonal  Asia  cold/dry  and  warm/wet  teleconnection  reacts  rapidly  to  climate  variations   and   are   mostly   driven   by   an   atmospheric   teleconnection.   Cold  conditions   increase   the   snow   cover   in   the   Northern   Hemisphere   causing  stronger  reflection  and  thus  lower  heating  of  the  northern  hemisphere  landmass,  which  shifts  the  ITCZ  southward  and  causes  drier  conditions  in  Monsoonal  Asia  (Liu  et  al.,  2013).      “For  paleoclimate,  the  last  two  decades  have  been  the  age  of  the  ice  core.  The  next  two  may   be   the   age   of   the   speleothem”,   Gideon   Henderson,   Science,   2006.   The  numerous   published   speleothem   studies   in   the   last   decade   confirm   the  statement  made  by  G.  Henderson.  Speleothems  are  valuable  archives  and  allow  reconstructing   climate   variations   on   the   continents   at   high   resolutions.   Most  important,   the   absolute   ages   that   can   be   measured   on   speleothems   allow  estimating   the   timing   and   duration   of   events   such   as   recently   done   for   the  duration  of  the  8.2  kyr  event.    

Page 60: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 50  

 References    Affek,  H.  P.,  Bar-­‐Matthews,  M.,  Ayalon,  A.,  Matthews,  A.,  and  Eiler,  J.  M.:  Glacial/interglacial  temperature  variations  in  Soreq  cave  speleothems  as  recorded  by  'clumped  isotope'  thermometry,  Geochimica  Et  Cosmochimica  Acta,  72,  5351-­‐5360,  2008.  

Arienzo,  M.  M.,  Swart,  P.  K.,  and  Vonhof,  H.  B.:  Measurement  of  delta  O-­‐18  and  delta  H-­‐2  values  of  fluid  inclusion  water  in  speleothems  using  cavity  ring-­‐down  spectroscopy  compared  with  isotope  ratio  mass  spectrometry,  Rapid  Communications  in  Mass  Spectrometry,  27,  2616-­‐2624,  2013.  

Ayalon,  A.,  Bar-­‐Matthews,  M.,  and  Kaufman,  A.:  Climatic  conditions  in  the  Eastern  Mediterranean  region  during  glacial  marine  isotopic  stage  6,  Geochimica  Et  Cosmochimica  Acta,  66,  A39-­‐A39,  2002.  

Bajo,  P.,  Drysdale,  R.,  Woodhead,  J.,  Hellstrom,  J.,  and  Zanchetta,  G.:  High-­‐resolution  U-­‐Pb  dating  of  an  Early  Pleistocene  stalagmite  from  Corchia  Cave  (central  Italy),  Quaternary  Geochronology,  14,  5-­‐17,  2012.  

Baker,  A.,  Genty,  D.,  Dreybrodt,  W.,  Barnes,  W.  L.,  Mockler,  N.  J.,  and  Grapes,  J.:  Testing  theoretically  predicted  stalagmite  growth  rate  with  Recent  annually  laminated  samples:  Implications  for  past  stalagmite  deposition,  Geochimica  Et  Cosmochimica  Acta,  62,  393-­‐404,  1998.  

Baker,  A.,  Ito,  E.,  Smart,  P.  L.,  and  McEwan,  R.  F.:  Elevated  and  variable  values  of  C-­‐13  in  speleothems  in  a  British  cave  system,  Chemical  Geology,  136,  263-­‐270,  1997.  

Baldini,  J.  U.  L.,  McDermott,  F.,  and  Fairchild,  I.  J.:  Spatial  variability  in  cave  drip  water  hydrochemistry:  Implications  for  stalagmite  paleoclimate  records,  Chemical  Geology,  235,  390-­‐404,  2006.  

Baldini,  J.  U.  L.,  McDermott,  F.,  Hoffmann,  D.  L.,  Richards,  D.  A.,  and  Clipson,  N.:  Very  high-­‐frequency  and  seasonal  cave  atmosphere  P-­‐CO2  variability:  Implications  for  stalagmite  growth  and  oxygen  isotope-­‐based  paleoclimate  records,  Earth  and  Planetary  Science  Letters,  272,  118-­‐129,  2008.  

Bar-­‐Matthews,  M.,  Ayalon,  A.,  Gilmour,  M.,  Matthews,  A.,  and  Hawkesworth,  C.  J.:  Sea-­‐land  oxygen  isotopic  relationships  from  planktonic  foraminifera  and  speleothems  in  the  Eastern  Mediterranean  region  and  their  implication  for  paleorainfall  during  interglacial  intervals,  Geochimica  Et  Cosmochimica  Acta,  67,  3181-­‐3199,  2003.  

Bar-­‐Matthews,  M.,  Ayalon,  A.,  and  Kaufman,  A.:  Late  quaternary  paleoclimate  in  the  eastern  Mediterranean  region  from  stable  isotope  analysis  of  speleothems  at  Soreq  Cave,  Israel,  Quaternary  Research,  47,  155-­‐168,  1997.  

Page 61: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  51  

Bar-­‐Matthews,  M.,  Ayalon,  A.,  and  Kaufman,  A.:  Timing  and  hydrological  conditions  of  Sapropel  events  in  the  Eastern  Mediterranean,  as  evident  from  speleothems,  Soreq  cave,  Israel,  Chemical  Geology,  169,  145-­‐156,  2000.  

Bar-­‐Matthews,  M.,  Ayalon,  A.,  Kaufman,  A.,  and  Wasserburg,  G.  J.:  The  Eastern  Mediterranean  paleoclimate  as  a  reflection  of  regional  events:  Soreq  cave,  Israel,  Earth  and  Planetary  Science  Letters,  166,  85-­‐95,  1999.  

BarMatthews,  M.,  Ayalon,  A.,  Matthews,  A.,  Sass,  E.,  and  Halicz,  L.:  Carbon  and  oxygen  isotope  study  of  the  active  water-­‐carbonate  system  in  a  karstic  Mediterranean  cave:  Implications  for  paleoclimate  research  in  semiarid  regions,  Geochimica  Et  Cosmochimica  Acta,  60,  337-­‐347,  1996.  

Beck,  W.  C.,  Grossman,  E.  L.,  and  Morse,  J.  W.:  Experimental  studies  of  oxygen  isotope  fractionation  in  the  carbonic  acid  system  at  15  degrees,  25  degrees,  and  40  degrees  C,  Geochimica  Et  Cosmochimica  Acta,  69,  3493-­‐3503,  2005.  

Bond,  G.,  Heinrich,  H.,  Broecker,  W.,  Labeyrie,  L.,  McManus,  J.,  Andrews,  J.,  Huon,  S.,  Jantschik,  R.,  Clasen,  S.,  Simet,  C.,  Tedesco,  K.,  Klas,  M.,  Bonani,  G.,  and  Ivy,  S.:  Evidence  for  massive  discharge  of  icebergs  into  the  North-­‐Atlantic  ocean  during  the  Last  Glacial  period  Nature,  360,  245-­‐249,  1992.  

Bonniver,  I.:  Etude  Hyrogeologique  et  dimmensionnement  par  modelisation  du  "systeme-­‐tracage"  du  reseau  karstique  the  Han-­‐sur-­‐Lesse  (Massif  de  Boine,  Belgique),  2011.  Geologie,  FUNDP  Namur,  Namur,  p  93  to  97  pp.,  2011.  

Borsato,  A.,  Frisia,  S.,  Fairchild,  I.  J.,  Somogyi,  A.,  and  Susini,  J.:  Trace  element  distribution  in  annual  stalagmite  laminae  mapped  by  micrometer-­‐resolution  X-­‐ray  fluorescence:  Implications  for  incorporation  of  environmentally  significant  species,  Geochimica  Et  Cosmochimica  Acta,  71,  1494-­‐1512,  2007.  

Breitenbach,  S.  F.  M.,  Rehfeld,  K.,  Goswami,  B.,  Baldini,  J.  U.  L.,  Ridley,  H.  E.,  Kennett,  D.  J.,  Prufer,  K.  M.,  Aquino,  V.  V.,  Asmerom,  Y.,  Polyak,  V.  J.,  Cheng,  H.,  Kurths,  J.,  and  Marwan,  N.:  COnstructing  Proxy  Records  from  Age  models  (COPRA),  Climate  of  the  Past,  8,  1765-­‐1779,  2012.  

Broecker,  W.  S.,  Olson,  E.  A.,  and  Orp,  P.  C.:  Radiacarbon  measurements  and  annual  rings  in  cave  formations,  Nature,  185,  93-­‐94,  1960.  

Buhl,  D.,  Immenhauser,  A.,  Smeulders,  G.,  Kabiri,  L.,  and  Richter,  D.  K.:  Time  series  delta  Mg-­‐26  analysis  in  speleothem  calcite:  Kinetic  versus  equilibrium  fractionation,  comparison  with  other  proxies  and  implications  for  palaeoclimate  research,  Chemical  Geology,  244,  715-­‐729,  2007.  

Burns,  S.  J.,  Fleitmann,  D.,  Matter,  A.,  Kramers,  J.,  and  Al-­‐Subbary,  A.  A.:  Indian  Ocean  climate  and  an  absolute  chronology  over  Dansgaard/Oeschger  events  9  to  13,  Science,  301,  1365-­‐1367,  2003.  

Chacko,  T.  and  Deines,  P.:  Theoretical  calculation  of  oxygen  isotope  fractionation  factors  in  carbonate  systems,  Geochimica  Et  Cosmochimica  Acta,  72,  3642-­‐3660,  2008.  

Page 62: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 52  

Cheng,  H.,  Edwards,  R.  L.,  Broecker,  W.  S.,  Denton,  G.  H.,  Kong,  X.,  Wang,  Y.,  Zhang,  R.,  and  Wang,  X.:  Ice  Age  Terminations,  Science,  326,  248-­‐252,  2009a.  

Cheng,  H.,  Edwards,  R.  L.,  Hoff,  J.,  Gallup,  C.  D.,  Richards,  D.  A.,  and  Asmerom,  Y.:  The  half-­‐lives  of  uranium-­‐234  and  thorium-­‐230,  Chemical  Geology,  169,  17-­‐33,  2000.  

Cheng,  H.,  Fleitmann,  D.,  Edwards,  R.  L.,  Wang,  X.  F.,  Cruz,  F.  W.,  Auler,  A.  S.,  Mangini,  A.,  Wang,  Y.  J.,  Kong,  X.  G.,  Burns,  S.  J.,  and  Matter,  A.:  Timing  and  structure  of  the  8.2  kyr  BP  event  inferred  from  delta  O-­‐18  records  of  stalagmites  from  China,  Oman,  and  Brazil,  Geology,  37,  1007-­‐1010,  2009b.  

Cobb,  K.  M.,  Adkins,  J.  F.,  Partin,  J.  W.,  and  Clark,  B.:  Regional-­‐scale  climate  influences  on  temporal  variations  of  rainwater  and  cave  dripwater  oxygen  isotopes  in  northern  Borneo,  Earth  and  Planetary  Science  Letters,  263,  207-­‐220,  2007.  

Cruz,  F.  W.,  Burns,  S.  J.,  Jercinovic,  M.,  Karmann,  I.,  Sharp,  W.  D.,  and  Vuille,  M.:  Evidence  of  rainfall  variations  in  Southern  Brazil  from  trace  element  ratios  (Mg/Ca  and  Sr/Ca)  in  a  Late  Pleistocene  stalagmite,  Geochimica  Et  Cosmochimica  Acta,  71,  2250-­‐2263,  2007.  

Cruz,  F.  W.,  Burns,  S.  J.,  Karmann,  I.,  Sharp,  W.  D.,  and  Vuille,  M.:  Reconstruction  of  regional  atmospheric  circulation  features  during  the  late  Pleistocene  in  subtropical  Brazil  from  oxygen  isotope  composition  of  speleothems,  Earth  and  Planetary  Science  Letters,  248,  495-­‐507,  2006.  

Cruz,  F.  W.,  Burns,  S.  J.,  Karmann,  I.,  Sharp,  W.  D.,  Vuille,  M.,  Cardoso,  A.  O.,  Ferrari,  J.  A.,  Dias,  P.  L.  S.,  and  Viana,  O.:  Insolation-­‐driven  changes  in  atmospheric  circulation  over  the  past  116,000  years  in  subtropical  Brazil,  Nature,  434,  63-­‐66,  2005.  

Cuntz,  M.:  CARBON  CYCLE  A  dent  in  carbon's  gold  standard,  Nature,  477,  547-­‐548,  2011.  

Daeron,  M.,  Guo,  W.,  Eiler,  J.,  Genty,  D.,  Blamart,  D.,  Boch,  R.,  Drysdale,  R.,  Maire,  R.,  Wainer,  K.,  and  Zanchetta,  G.:  (CO)-­‐C-­‐13-­‐O-­‐18  clumping  in  speleothems:  Observations  from  natural  caves  and  precipitation  experiments,  Geochimica  Et  Cosmochimica  Acta,  75,  3303-­‐3317,  2011.  

Dansgaard,  W.:  Stable  Isotopes  In  Precipitation,  Tellus,  16,  436-­‐468,  1964.  

Dansgaard,  W.,  Johnsen,  S.  J.,  Clausen,  H.  B.,  Dahljensen,  D.,  Gundestrup,  N.  S.,  Hammer,  C.  U.,  Hvidberg,  C.  S.,  Steffensen,  J.  P.,  Sveinbjornsdottir,  A.  E.,  Jouzel,  J.,  and  Bond,  G.:  Evidence  for  general  instability  of  past  climate  from  a  250-­‐Kyr  Ice-­‐Core  record  Nature,  364,  218-­‐220,  1993.  

De  Cisneros,  C.  J.,  Caballero,  E.,  Vera,  J.  A.,  and  Andreo,  B.:  An  optimized  thermal  extraction  system  for  preparation  of  water  from  fluid  inclusions  in  speleothems,  Geologica  Acta,  9,  149-­‐158,  2011.  

Page 63: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  53  

Deininger,  M.,  Fohlmeister,  J.,  Scholz,  D.,  and  Mangini,  A.:  Isotope  disequilibrium  effects:  The  influence  of  evaporation  and  ventilation  effects  on  the  carbon  and  oxygen  isotope  composition  of  speleothems  -­‐  A  model  approach,  Geochimica  Et  Cosmochimica  Acta,  96,  57-­‐79,  2012.  

Demeny,  A.,  Kele,  S.,  and  Siklosy,  Z.:  Empirical  equations  for  the  temperature  dependence  of  calcite-­‐water  oxygen  isotope  fractionation  from  10  to  70  degrees  C,  Rapid  Communications  in  Mass  Spectrometry,  24,  3521-­‐3526,  2010.  

Dennis,  P.  F.,  Rowe,  P.  J.,  and  Atkinson,  T.  C.:  The  recovery  and  isotopic  measurement  of  water  from  fluid  inclusions  in  speleothems,  Geochimica  Et  Cosmochimica  Acta,  65,  871-­‐884,  2001.  

Dorale,  J.  A.,  Edwards,  R.  L.,  Ito,  E.,  and  Gonzalez,  L.  A.:  Climate  and  vegetation  history  of  the  midcontinent  from  75  to  25  ka:  A  speleothem  record  from  Crevice  Cave,  Missouri,  USA,  Science,  282,  1871-­‐1874,  1998.  

Dreybrodt,  W.:  Comments  on  processes  contributing  to  the  isotope  composition  of  13C  and  18O  in  calcite  depostied  in  speleothems,  Acta  Carsologica,  40,  2011.  

Dreybrodt,  W.:  Deposition  of  calcite  from  thin  water  films  of  natural  calcareous  solutions  and  the  growth  of  speleothems  Chemical  Geology,  29,  89-­‐105,  1980.  

Dreybrodt,  W.:  Evolution  of  the  isotopic  composition  of  carbon  in  a  calcite  precipitating  H2O-­‐CO2-­‐CaCO3  solution  and  the  related  isotopic  composition  of  calcite  in  stalagmites,  Geochimica  Et  Cosmochimica  Acta,  72,  4712-­‐4724,  2008.  

Dreybrodt,  W.:  Processes  in  karts  systems.  Physics,  Chemistry  and  Geology,  Belrin/Heidelberg/  New  York,  1988.  

Dreybrodt,  W.  and  Scholz,  D.:  Climatic  dependence  of  stable  carbon  and  oxygen  isotope  signals  recorded  in  speleothems:  From  soil  water  to  speleothem  calcite,  Geochimica  Et  Cosmochimica  Acta,  75,  734-­‐752,  2011.  

Dulinski,  M.  and  Rozanski,  K.:  Formation  of  12C/13C  isotope  ratios  in  spsleothems:  a  semi-­‐dynamic  model,  Radiocarbon,  32,  7-­‐16,  1990.  

Dykoski,  C.  A.,  Edwards,  R.  L.,  Cheng,  H.,  Yuan,  D.  X.,  Cai,  Y.  J.,  Zhang,  M.  L.,  Lin,  Y.  S.,  Qing,  J.  M.,  An,  Z.  S.,  and  Revenaugh,  J.:  A  high-­‐resolution,  absolute-­‐dated  Holocene  and  deglacial  Asian  monsoon  record  from  Dongge  Cave,  China,  Earth  and  Planetary  Science  Letters,  233,  71-­‐86,  2005.  

Edwards,  R.  L.,  Chen,  J.  H.,  and  Wasserburg,  G.  J.:  238U-­‐234U-­‐230Th-­‐232Th  systematics  and  the  precise  measurement  of  time  over  the  past  500  000  years,  Earth  and  Planetary  Science  Letters,  81,  175-­‐192,  1987.  

Eiler,  J.  M.:  "Clumped-­‐isotope"  geochemistry  -­‐  The  study  of  naturally-­‐occurring,  multiply-­‐substituted  isotopologues,  Earth  and  Planetary  Science  Letters,  262,  309-­‐327,  2007.  

Page 64: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 54  

Emrich,  K.,  Ehhalt,  D.  H.,  and  Volgel,  J.  C.:  Carbon  isotope  fractionation  during  the  precipitation  of  calciumcarbonates,  Earth  and  Planetary  Science  Letters,  8,  363-­‐371,  1970.  

Fairchild,  I.  J.  and  Baker,  A.:  Speleothem  Science:  From  process  to  past  environment,  Blackwell  Publishing  ltd.,  2012.  

Fairchild,  I.  J.,  Baker,  A.,  Borsato,  A.,  Frisia,  S.,  Hinton,  R.  W.,  McDermott,  F.,  and  Tooth,  A.  F.:  Annual  to  sub-­‐annual  resolution  of  multiple  trace-­‐element  trends  in  speleothems,  Journal  of  the  Geological  Society,  158,  831-­‐841,  2001.  

Fairchild,  I.  J.,  Borsato,  A.,  Tooth,  A.  F.,  Frisia,  S.,  Hawkesworth,  C.  J.,  Huang,  Y.  M.,  McDermott,  F.,  and  Spiro,  B.:  Controls  on  trace  element  (Sr-­‐Mg)  compositions  of  carbonate  cave  waters:  implications  for  speleothem  climatic  records,  Chemical  Geology,  166,  255-­‐269,  2000.  

Fairchild,  I.  J.,  Smith,  C.  L.,  Baker,  A.,  Fuller,  L.,  Spötl,  C.,  Mattey,  D.,  McDermott,  F.,  and  Eimp:  Modification  and  preservation  of  environmental  signals  in  speleothems,  Earth-­‐Science  Reviews,  75,  105-­‐153,  2006.  

Fairchild,  I.  J.  and  Treble,  P.  C.:  Trace  elements  in  speleothems  as  recorders  of  environmental  change,  Quaternary  Science  Reviews,  28,  449-­‐468,  2009.  

Fleitmann,  D.,  Burns,  S.  J.,  Mudelsee,  M.,  Neff,  U.,  Kramers,  J.,  Mangini,  A.,  and  Matter,  A.:  Holocene  forcing  of  the  Indian  monsoon  recorded  in  a  stalagmite  from  Southern  Oman,  Science,  300,  1737-­‐1739,  2003.  

Fleitmann,  D.,  Burns,  S.  J.,  Neff,  U.,  Mudelsee,  M.,  Mangini,  A.,  and  Matter,  A.:  Palaeoclimatic  interpretation  of  high-­‐resolution  oxygen  isotope  profiles  derived  from  annually  laminated  speleothems  from  Southern  Oman,  Quaternary  Science  Reviews,  23,  935-­‐945,  2004.  

Fleitmann,  D.,  Cheng,  H.,  Badertscher,  S.,  Edwards,  R.  L.,  Mudelsee,  M.,  Gokturk,  O.  M.,  Fankhauser,  A.,  Pickering,  R.,  Raible,  C.  C.,  Matter,  A.,  Kramers,  J.,  and  Tuysuz,  O.:  Timing  and  climatic  impact  of  Greenland  interstadials  recorded  in  stalagmites  from  northern  Turkey,  Geophysical  Research  Letters,  36,  2009.  

Fohlmeister,  J.,  Schroder-­‐Ritzrau,  A.,  Scholz,  D.,  Spötl,  C.,  Riechelmann,  D.  F.  C.,  Mudelsee,  M.,  Wackerbarth,  A.,  Gerdes,  A.,  Riechelmann,  S.,  Immenhauser,  A.,  Richter,  D.  K.,  and  Mangini,  A.:  Bunker  Cave  stalagmites:  an  archive  for  central  European  Holocene  climate  variability,  Climate  of  the  Past,  8,  1751-­‐1764,  2012.  

Friedman,  I.  and  O'Neil,  J.  R.:  Compilation  of  stable  isotope  fractionation  factors  of  geochemical  interest,  Geolog.  Surv.  Prof.  Paper  440-­‐KK,  1977.  117,  1977.  

Frisia,  S.,  Borsato,  A.,  Fairchild,  I.  J.,  and  McDermott,  F.:  Calcite  fabrics,  growth  mechanisms,  and  environments  of  formation  in  speleothems  from  the  Italian  Alps  and  southwestern  Ireland,  Journal  of  Sedimentary  Research,  70,  1183-­‐1196,  2000.  

Page 65: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  55  

Frisia,  S.,  Borsato,  A.,  Preto,  N.,  and  McDermott,  F.:  Late  Holocene  annual  growth  in  three  Alpine  stalagmites  records  the  influence  of  solar  activity  and  the  North  Atlantic  Oscillation  on  winter  climate,  Earth  and  Planetary  Science  Letters,  216,  411-­‐424,  2003.  

Fuller,  L.,  Baker,  A.,  Fairchild,  I.  J.,  Spötl,  C.,  Marca-­‐Bell,  A.,  Rowe,  P.,  and  Dennis,  P.  F.:  Isotope  hydrology  of  dripwaters  in  a  Scottish  cave  and  implications  for  stalagmite  palaeoclimate  research,  Hydrology  and  Earth  System  Sciences,  12,  1065-­‐1074,  2008.  

Gabitov,  R.  I.  and  Watson,  E.  B.:  Partitioning  of  strontium  between  calcite  and  fluid,  Geochemistry  Geophysics  Geosystems,  7,  2006.  

Galy,  A.,  Bar-­‐Matthews,  M.,  Halicz,  L.,  and  O'Nions,  R.  K.:  Mg  isotopic  composition  of  carbonate:  insight  from  speleothem  formation,  Earth  and  Planetary  Science  Letters,  201,  105-­‐115,  2002.  

Gascoyne,  M.:  Paleioclimatic  determination  from  cave  calcite  deposits  Quaternary  Science  Reviews,  11,  609-­‐632,  1992.  

Gascoyne,  M.:  Trace  element  partition  coefficients  in  the  calcite  water  system  and  their  paleoclimatic  significance  in  cave  studies  Journal  of  Hydrology,  61,  213-­‐222,  1983.  

Gat,  J.  R.:  Oxygen  and  hydrogen  isotopes  in  the  hydrologic  cycle,  Annual  Review  of  Earth  and  Planetary  Sciences,  24,  225-­‐262,  1996.  

Genty,  D.,  Baker,  A.,  Massault,  M.,  Proctor,  C.,  Gilmour,  M.,  Pons-­‐Branchu,  E.,  and  Hamelin,  B.:  Dead  carbon  in  stalagmites:  Carbonate  bedrock  paleodissolution  vs.  ageing  of  soil  organic  matter.  Implications  for  C-­‐13  variations  in  speleothems,  Geochimica  Et  Cosmochimica  Acta,  65,  3443-­‐3457,  2001a.  

Genty,  D.,  Baker,  A.,  and  Vokal,  B.:  Intra-­‐  and  inter-­‐annual  growth  rate  of  modern  stalagmites,  Chemical  Geology,  176,  191-­‐212,  2001b.  

Genty,  D.,  Blamart,  D.,  Ouahdi,  R.,  Gilmour,  M.,  Baker,  A.,  Jouzel,  J.,  and  Van-­‐Exter,  S.:  Precise  dating  of  Dansgaard-­‐Oeschger  climate  oscillations  in  western  Europe  from  stalagmite  data,  Nature,  421,  833-­‐837,  2003.  

Genty,  D.  and  Deflandre,  G.:  Drip  flow  variations  under  a  stalactite  of  the  Pere  Noel  cave  (Belgium).  Evidence  of  seasonal  variations  and  air  pressure  constraints,  Journal  of  Hydrology,  211,  208-­‐232,  1998.  

Genty,  D.,  Labuhn,  I.,  Hoffmann,  G.,  Danis,  P.  A.,  Mestre,  O.,  Bourges,  F.,  Wainer,  K.,  Massault,  M.,  Van  Exter,  S.,  Regnier,  E.,  Orengo,  P.,  Falourd,  S.,  and  Minster,  B.:  Rainfall  and  cave  water  isotopic  relationships  in  two  South-­‐France  sites,  Geochimica  Et  Cosmochimica  Acta,  131,  323-­‐343,  2014.  

Genty,  D.  and  Massault,  M.:  Bomb  C-­‐14  recorded  in  laminated  speleothems:  Calculation  of  dead  carbon  proportion,  Radiocarbon,  39,  33-­‐48,  1997.  

Page 66: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 56  

Genty,  D.  and  Quinif,  Y.:  Annually  laminated  sequences  in  the  internal  structure  of  some  Belgian  stalagmites  -­‐  Importance  for  paleoclimatology,  Journal  of  Sedimentary  Research,  66,  275-­‐288,  1996.  

Goede,  A.,  McCulloch,  M.,  McDermott,  F.,  and  Hawkesworth,  C.:  Aeolian  contribution  to  strontium  and  strontium  isotope  variations  in  a  Tasmanian  speleothem,  Chemical  Geology,  149,  37-­‐50,  1998.  

Hansen,  M.,  Dreybrodt,  W.,  and  Scholz,  D.:  Chemical  evolution  of  dissolved  inorganic  carbon  species  flowing  in  thin  water  films  and  its  implications  for  (rapid)  degassing  of  CO2  during  speleothem  growth,  Geochimica  Et  Cosmochimica  Acta,  107,  242-­‐251,  2013.  

Hartland,  A.,  Fairchild,  I.  J.,  Lead,  J.  R.,  and  Baker,  A.:  Fluorescent  properties  of  organic  carbon  in  cave  dripwaters:  Effects  of  filtration,  temperature  and  pH,  Science  of  the  Total  Environment,  408,  5940-­‐5950,  2010.  

Henderson,  G.  M.:  Climate  -­‐  Caving  in  to  new  chronologies,  Science,  313,  620-­‐622,  2006.  

Hendy,  C.  H.:  Isotopic  geochemistry  of  speleothems:  1.  Calculations  of  effects  on  different  modes  of  formation  on  isotopic  composition  of  speleothems  and  their  applicability  as  paleoclimatic  indicators,  Geochimica  Et  Cosmochimica  Acta,  35,  801-­‐824,  1971.  

Hendy,  C.  H.  and  Wilson,  A.  T.:  Palaeoclimatic  data  from  speleothems  Nature,  219,  48-­‐&,  1968.  

Herbert,  T.  D.,  Schuffert,  J.  D.,  Andreasen,  D.,  Heusser,  L.,  Lyle,  M.,  Mix,  A.,  Ravelo,  A.  C.,  Stott,  L.  D.,  and  Heguera,  J.  C.:  The  California  Current,  devils  hole,  and  Pleistocene  climate  -­‐  Response,  Science,  296,  2002.  

Hercman,  H.  and  Pawlak,  J.:  MOD-­‐AGE:  An  age-­‐depth  model  construction  algorithm,  Quaternary  Geochronology,  12,  1-­‐10,  2012.  

Hodge,  E.  J.,  Richards,  D.  A.,  Smart,  P.  L.,  Gines,  A.,  and  Mattey,  D.  P.:  Sub-­‐millennial  climate  shifts  in  the  western  Mediterranean  during  the  last  glacial  period  recorded  in  a  speleothem  from  Mallorca,  Spain,  Journal  of  Quaternary  Science,  23,  713-­‐718,  2008.  

Hoefs,  J.:  Stable  isotopes  geochemistry  Berlin,  1997.  

Holland,  H.  D.,  Kirsipu,  T.  V.,  Huebner,  J.  S.,  and  Oxburgh,  U.  M.:  On  some  aspects  of  the  chemical  evolution  of  cave  waters  Journal  of  Geology,  72,  36-­‐67,  1964.  

Hori,  M.,  Ishikawa,  T.,  Nagaishi,  K.,  Lin,  K.,  Wang,  B.  S.,  You,  C.  F.,  Shen,  C.  C.,  and  Kano,  A.:  Prior  calcite  precipitation  and  source  mixing  process  influence  Sr/Ca,  Ba/Ca  and  Sr-­‐87/Sr-­‐86  of  a  stalagmite  developed  in  southwestern  Japan  during  18.0  4.5  ka,  Chemical  Geology,  347,  190-­‐198,  2013.  

Page 67: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  57  

Horita,  J.  and  Clayton,  R.  N.:  Comment  on  the  studies  of  oxygen  isotope  fractionation  between  calcium  carbonates  and  water  at  low  temperatures  by  Zhou  and  Zheng  (2003;2005),  Geochimica  Et  Cosmochimica  Acta,  71,  3131-­‐3135,  2007.  

Huang,  H.  M.,  Fairchild,  I.  J.,  Borsato,  A.,  Frisia,  S.,  Cassidy,  N.  J.,  McDermott,  F.,  and  Hawkesworth,  C.  J.:  Seasonal  variations  in  Sr,  Mg  and  P  in  modern  speleothems  (Grotta  di  Ernesto,  Italy),  Chemical  Geology,  175,  429-­‐448,  2001.  

Jimenez-­‐Lopez,  C.,  Caballero,  E.,  Huertas,  F.  J.,  and  Romanek,  C.  S.:  Chemical,  mineralogical  and  isotope  behavior,  and  phase  transformation  during  the  precipitation  of  calcium  carbonate  minerals  from  intermediate  ionic  solution  at  25  degrees  C,  Geochimica  Et  Cosmochimica  Acta,  65,  3219-­‐3231,  2001.  

Karmann,  I.,  Cruz,  F.  W.,  Viana,  O.,  and  Burns,  S.  J.:  Climate  influence  on  geochemistry  parameters  of  waters  from  Santana-­‐Perolas  cave  system,  Brazil,  Chemical  Geology,  244,  232-­‐247,  2007.  

Katz,  A.:  Interaction  of  Mg  with  Calcite  during  crystal-­‐growth  at  25-­‐90  degrees  C  and  1  atmophere,  Geochimica  Et  Cosmochimica  Acta,  37,  1563-­‐&,  1973.  

Katz,  A.,  Starinsk.A,  Sass,  E.,  and  Holland,  H.  D.:  Strontium  behaviour  in  aragonite-­‐calcite  transformation-­‐  experimental  study  at  40-­‐98  degrees  C  Geochimica  Et  Cosmochimica  Acta,  36,  481-­‐&,  1972.  

Kaufman,  A.,  Wasserburg,  G.  J.,  Porcelli,  D.,  Bar-­‐Matthews,  M.,  Ayalon,  A.,  and  Halicz,  L.:  U-­‐Th  isotope  systematics  from  the  Soreq  cave,  Israel  and  climatic  correlations,  Earth  and  Planetary  Science  Letters,  156,  141-­‐155,  1998.  

Kim,  S.-­‐T.,  Mucci,  A.,  and  Taylor,  B.  E.:  Phosphoric  acid  fractionation  factors  for  calcite  and  aragonite  between  25  and  75  degrees  C:  Revisited,  Chemical  Geology,  246,  135-­‐146,  2007.  

Kim,  S.  T.  and  O'Neil,  J.  R.:  Equilibrium  and  nonequilibrium  oxygen  isotope  effects  in  synthetic  carbonates,  Geochimica  Et  Cosmochimica  Acta,  61,  3461-­‐3475,  1997.  

Kluge,  T.,  Marx,  T.,  Scholz,  D.,  Niggemann,  S.,  Mangini,  A.,  and  Aeschbach-­‐Hertig,  W.:  A  new  tool  for  palaeoclimate  reconstruction:  Noble  gas  temperatures  from  fluid  inclusions  in  speleothems,  Earth  and  Planetary  Science  Letters,  269,  407-­‐414,  2008.  

Kluge,  T.,  Riechelmann,  D.  F.  C.,  Wieser,  M.,  Spötl,  C.,  Sultenfuss,  J.,  Schroder-­‐Ritzrau,  A.,  Niggemann,  S.,  and  Aeschbach-­‐Hertig,  W.:  Dating  cave  drip  water  by  tritium,  Journal  of  Hydrology,  394,  396-­‐406,  2010.  

Kong,  X.  G.,  Wang,  Y.  J.,  Edwards,  L.  R.,  Cheng,  H.,  and  Wang,  X.  F.:  The  altitude  effect  on  oxygen  isotope  composition  of  stalagmite,  Abstracts  with  programs,  Geological  Society  of  America,  37,  13,  2005.  

Lachniet,  M.  S.:  Climatic  and  environmental  controls  on  speleothem  oxygen-­‐isotope  values,  Quaternary  Science  Reviews,  28,  412-­‐432,  2009.  

Page 68: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 58  

Lachniet,  M.  S.  and  Patterson,  W.  P.:  Use  of  correlation  and  stepwise  regression  to  evaluate  physical  controls  on  the  stable  isotope  values  of  Panamanian  rain  and  surface  waters,  Journal  of  Hydrology,  324,  115-­‐140,  2006.  

Lauritzen,  S.  E.  and  Lundberg,  J.:  Calibration  of  the  speleothem  delta  function:  an  absolute  temperature  record  for  the  Holocene  in  northern  Norway,  Holocene,  9,  659-­‐669,  1999.  

Liu,  Y.  H.,  Henderson,  G.  M.,  Hu,  C.  Y.,  Mason,  A.  J.,  Charnley,  N.,  Johnson,  K.  R.,  and  Xie,  S.  C.:  Links  between  the  East  Asian  monsoon  and  North  Atlantic  climate  during  the  8,200  year  event,  Nature  Geoscience,  6,  117-­‐120,  2013.  

Mangini,  A.,  Spötl,  C.,  and  Verdes,  P.:  Reconstruction  of  temperature  in  the  Central  Alps  during  the  past  2000  yr  from  a  delta  O-­‐18  stalagmite  record,  Earth  and  Planetary  Science  Letters,  235,  741-­‐751,  2005.  

Mattey,  D.,  Lowry,  D.,  Duffet,  J.,  Fisher,  R.,  Hodge,  E.,  and  Frisia,  S.:  A  53  year  seasonally  resolved  oxygen  and  carbon  isotope  record  from  a  modem  Gibraltar  speleothem:  Reconstructed  drip  water  and  relationship  to  local  precipitation,  Earth  and  Planetary  Science  Letters,  269,  80-­‐95,  2008.  

Matthews,  A.,  Ayalon,  A.,  and  Bar-­‐Matthews,  M.:  D/H  ratios  of  fluid  inclusions  of  Soreq  cave  (Israel)  speleothems  as  a  guide  to  the  Eastern  Mediterranean  Meteoric  Line  relationships  in  the  last  120  ky,  Chemical  Geology,  166,  183-­‐191,  2000.  

McDermott,  F.:  Palaeo-­‐climate  reconstruction  from  stable  isotope  variations  in  speleothems:  a  review,  Quaternary  Science  Reviews,  23,  901-­‐918,  2004.  

McDermott,  F.,  Atkinson,  T.  C.,  Fairchild,  I.  J.,  Baldini,  L.  M.,  and  Mattey,  D.  P.:  A  first  evaluation  of  the  spatial  gradients  in  delta  O-­‐18  recorded  by  European  Holocene  speleothems,  Glob.  Planet.  Change,  79,  275-­‐287,  2011.  

McDermott,  F.,  Frisia,  S.,  Huang,  Y.  M.,  Longinelli,  A.,  Spiro,  B.,  Heaton,  T.  H.  E.,  Hawkesworth,  C.  J.,  Borsato,  A.,  Keppens,  E.,  Fairchild,  I.  J.,  van  der  Borg,  K.,  Verheyden,  S.,  and  Selmo,  E.:  Holocene  climate  variability  in  Europe:  Evidence  from  delta  O-­‐18,  textural  and  extension-­‐rate  variations  in  three  speleothems,  Quaternary  Science  Reviews,  18,  1021-­‐1038,  1999.  

McDermott,  F.,  Schwarcz,  H.  P.,  and  Rowe,  P.  J.:  Isotopes  in  speleothems,  In  Isotopes  in  paleoenvironmental  research  (ed.  M.J.  Leng),  2005.  185-­‐226,  2005.  

McDonald,  J.,  Drysdale,  R.,  and  Hill,  D.:  The  2002-­‐2003  El  Nino  recorded  in  Australian  cave  drip  waters:  Implications  for  reconstructing  rainfall  histories  using  stalagmites,  Geophysical  Research  Letters,  31,  2004.  

McDonald,  J.,  Drysdale,  R.,  Hill,  D.,  Chisari,  R.,  and  Wong,  H.:  The  hydrochemical  response  of  cave  drip  waters  to  sub-­‐annual  and  inter-­‐annual  climate  variability,  Wombeyan  Caves,  SE  Australia,  Chemical  Geology,  244,  605-­‐623,  2007.  

Page 69: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  59  

McGillen,  M.  R.  and  Fairchild,  I.  J.:  An  experimental  study  of  incongruent  dissolution  of  CaCO3  under  analogue  glacial  conditions,  Journal  of  Glaciology,  51,  383-­‐390,  2005.  

McIntire,  W.  L.:  Trace  element  partition  coeficients-­‐  a  review  of  theory  and  applications  to  Geology,  Geochimica  Et  Cosmochimica  Acta,  27,  1209-­‐1264,  1963.  

Mickler,  P.  J.,  Stern,  L.  A.,  and  Banner,  J.  L.:  Large  kinetic  isotope  effects  in  modern  speleothems,  Geological  Society  of  America  Bulletin,  118,  65-­‐81,  2006.  

Miorandi,  R.,  Borsato,  A.,  Frisia,  S.,  Fairchild,  I.  J.,  and  Richter,  D.  K.:  Epikarst  hydrology  and  implications  for  stalagmite  capture  of  climate  changes  at  Grotta  di  Ernesto  (NE  Italy):  results  from  long-­‐term  monitoring,  Hydrological  Processes,  24,  3101-­‐3114,  2010.  

Mook,  W.  G.:  Volume  1:Introduction-­‐theory,  methods,  review.  In:,  Environmental  Isotopes  in  the  Hydrological  Cycle.  Priciples  and  Applications.,  2000.  2000.  

Mühlinghaus,  C.,  Scholz,  D.,  and  Mangini,  A.:  Modelling  fractionation  of  stable  isotopes  in  stalagmites,  Geochimica  Et  Cosmochimica  Acta,  doi:  10.1016/j.gca.2009.09.010,  2009.  7275-­‐7289,  2009.  

Mühlinghaus,  C.,  Scholz,  D.,  and  Mangini,  A.:  Modelling  stalagmite  growth  and  delta  C-­‐13  as  a  function  of  drip  interval  and  temperature,  Geochimica  Et  Cosmochimica  Acta,  71,  2780-­‐2790,  2007.  

Musgrove,  M.  and  Banner,  J.  L.:  Controls  on  the  spatial  and  temporal  variability  of  vadose  dripwater  geochemistry:  Edwards  Aquifer,  central  Texas,  Geochimica  Et  Cosmochimica  Acta,  68,  1007-­‐1020,  2004.  

Neff,  U.,  Burns,  S.  J.,  Mangini,  A.,  Mudelsee,  M.,  Fleitmann,  D.,  and  Matter,  A.:  Strong  coherence  between  solar  variability  and  the  monsoon  in  Oman  between  9  and  6  kyr  ago,  Nature,  411,  290-­‐293,  2001.  

O'Neil,  J.  R.,  Clayton,  R.  N.,  and  Mayeda,  T.  K.:  Oxygen  isotope  fractionation  in  divalent  metal  carbonates  Journal  of  Chemical  Physics,  51,  5547-­‐&,  1969.  

Orland,  I.  J.,  Burstyn,  Y.,  Bar-­‐Matthews,  M.,  Kozdon,  R.,  Ayalon,  A.,  Matthews,  A.,  and  Valley,  J.  W.:  Seasonal  climate  signals  (1990-­‐2008)  in  a  modern  Soreq  Cave  stalagmite  as  revealed  by  high-­‐resolution  geochemical  analysis,  Chemical  Geology,  363,  322-­‐333,  2014.  

Palmer,  M.  R.  and  Edmond,  J.  M.:  Controls  over  Strontium  isotope  composition  of  river  water  Geochimica  Et  Cosmochimica  Acta,  56,  2099-­‐2111,  1992.  

Piccini,  L.,  Zanchetta,  G.,  Drysdale,  R.  N.,  Hellstrom,  J.,  Isola,  I.,  Fallick,  A.  E.,  Leone,  G.,  Doveri,  M.,  Mussi,  M.,  Mantelli,  F.,  Molli,  G.,  Lotti,  L.,  Roncioni,  A.,  Regattieri,  E.,  Meccheri,  M.,  and  Vaselli,  L.:  The  environmental  features  of  the  Monte  Corchia  cave  system  (Apuan  Alps,  central  Italy)  and  their  effects  on  speleothem  growth,  International  Journal  of  Speleology,  37,  153-­‐172,  2008.  

Page 70: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 60  

Pingitore,  N.  E.  and  Eastman,  M.  P.:  The  co-­‐precipitation  of  Sr2+  with  calcite  at  25  degrees  C  and  1  atm,  Geochimica  Et  Cosmochimica  Acta,  50,  2195-­‐2203,  1986.  

Polyak,  V.,  Hill,  C.,  and  Asmerom,  Y.:  Age  and  evolution  of  the  Grand  Canyon  revealed  by  U-­‐Pb  dating  of  water  table-­‐type  speleothems,  Science,  319,  1377-­‐1380,  2008.  

Proctor,  C.  J.,  Baker,  A.,  Barnes,  W.  L.,  and  Gilmour,  R.  A.:  A  thousand  year  speleothem  proxy  record  of  North  Atlantic  climate  from  Scotland,  Climate  Dynamics,  16,  815-­‐820,  2000.  

Rayleigh,  J.  W.  S.:  Theoritical  considerations  respecting  the  separation  of  gasses  by  diffusion  and  similar  processes,  Philos  Magasine,  12,  493p,  1896.  

Richards,  D.  A.,  Bottrell,  S.  H.,  Cliff,  R.  A.,  Strohle,  K.,  and  Rowe,  P.  J.:  U-­‐Pb  dating  of  a  speleothem  of  Quaternary  age,  Geochimica  Et  Cosmochimica  Acta,  62,  3683-­‐3688,  1998.  

Riechelmann,  D.  F.  C.,  Deininger,  M.,  Scholz,  D.,  Riechelmann,  S.,  Schroeder-­‐Ritzrau,  A.,  Spoetl,  C.,  Richter,  D.  K.,  Mangini,  A.,  and  Immenhauser,  A.:  Disequilibrium  carbon  and  oxygen  isotope  fractionation  in  recent  cave  calcite:  Comparison  of  cave  precipitates  and  model  data,  Geochimica  Et  Cosmochimica  Acta,  103,  232-­‐244,  2013.  

Riechelmann,  D.  F.  C.,  Schroeder-­‐Ritzrau,  A.,  Scholz,  D.,  Fohlmeister,  J.,  Spoetl,  C.,  Richter,  D.  K.,  and  Mangini,  A.:  Monitoring  Bunker  Cave  (NW  Germany):  A  prerequisite  to  interpret  geochemical  proxy  data  of  speleothems  from  this  site,  Journal  of  Hydrology,  409,  682-­‐695,  2011.  

Roberts,  M.  S.,  Smart,  P.  L.,  and  Baker,  A.:  Annual  trace  element  variations  in  a  Holocene  speleothem,  Earth  and  Planetary  Science  Letters,  154,  237-­‐246,  1998.  

Rozanski,  K.,  Araguasaraguas,  L.,  and  Gonfiantini,  R.:  Relationship  between  long-­‐term  trends  of  18O  isotope  composition  of  precipitation  and  climate  Science,  258,  981-­‐985,  1992.  

Rutlidge,  H.,  Baker,  A.,  Marjo,  C.  E.,  Andersen,  M.  S.,  Graham,  P.  W.,  Cuthbert,  M.  O.,  Rau,  G.  C.,  Roshan,  H.,  Markowska,  M.,  Mariethoz,  G.,  and  Jex,  C.  N.:  Dripwater  organic  matter  and  trace  element  geochemistry  in  a  semi-­‐arid  karst  environment:  Implications  for  speleothem  paleoclimatology,  Geochimica  Et  Cosmochimica  Acta,  135,  217-­‐230,  2014.  

Salomons,  W.  and  Mook,  W.  G.:  Isotope  Geochemistry  in  carbonates  in  the  weathering  zone,  1986.  

Scheidegger,  Y.,  Baur,  H.,  Brennwald,  M.  S.,  Fleitmann,  D.,  Wieler,  R.,  and  Kipfer,  R.:  Accurate  analysis  of  noble  gas  concentrations  in  small  water  samples  and  its  application  to  fluid  inclusions  in  stalagmites,  Chemical  Geology,  272,  31-­‐39,  2010.  

Page 71: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  61  

Schmidt,  G.  A.,  LeGrande,  A.  N.,  and  Hoffmann,  G.:  Water  isotope  expressions  of  intrinsic  and  forced  variability  in  a  coupled  ocean-­‐atmosphere  model,  Journal  of  Geophysical  Research-­‐Atmospheres,  112,  2007.  

Scholz,  D.  and  Hoffmann,  D.  L.:  StalAge  -­‐  An  algorithm  designed  for  construction  of  speleothem  age  models,  Quaternary  Geochronology,  6,  369-­‐382,  2011.  

Scholz,  D.,  Hoffmann,  D.  L.,  Hellstrom,  J.,  and  Ramsey,  C.  B.:  A  comparison  of  different  methods  for  speleothem  age  modelling,  Quaternary  Geochronology,  14,  94-­‐104,  2012.  

Scholz,  D.,  Muehlinghaus,  C.,  and  Mangini,  A.:  Modelling  delta  C-­‐13  and  delta  O-­‐18  in  the  solution  layer  on  stalagmite  surfaces,  Geochimica  Et  Cosmochimica  Acta,  73,  2592-­‐2602,  2009.  

Shakun,  J.  D.,  Burns,  S.  J.,  Fleitmann,  D.,  Kramers,  J.,  Matter,  A.,  and  Al-­‐Subary,  A.:  A  high-­‐resolution,  absolute-­‐dated  deglacial  speleothem  record  of  Indian  Ocean  climate  from  Socotra  Island,  Yemen,  Earth  and  Planetary  Science  Letters,  259,  442-­‐456,  2007.  

Smart,  P.  L.  and  Friedrich,  H.:  Water  movement  and  storage  in  the  unsaturated  zone  of  amaturely  karstified  aquifer,  Mendip  Hills,  England,  Proceedings  of  the  Conference  on  Environ-­‐mental  Problems  in  Karst  Terrains  and  Teir  Solution,  Bowling  Green,  Kentucky,National  Water  Well  Association,  1987.  57-­‐87,  1987.  

Smith,  C.  L.,  Fairchild,  I.  J.,  Spötl,  C.,  Frisia,  S.,  Borsato,  A.,  Moreton,  S.  G.,  and  Wynn,  P.  M.:  Chronology  building  using  objective  identification  of  annual  signals  in  trace  element  profiles  of  stalagmites,  Quaternary  Geochronology,  4,  11-­‐21,  2009.  

Spötl,  C.  and  Mangini,  A.:  Stalagmite  from  the  Austrian  Alps  reveals  Dansgaard-­‐Oeschger  events  during  isotope  stage  3:  Implications  for  the  absolute  chronology  of  Greenland  ice  cores,  Earth  and  Planetary  Science  Letters,  203,  507-­‐518,  2002.  

Spötl,  C.,  Unterwurzacher,  M.,  Mangini,  A.,  and  Longstaffe,  F.  J.:  Carbonate  speleothems  in  the  dry,  inneralpine  Vinschgau  Valley,  northernmost  Italy:  Witnesses  of  changes  in  climate  and  hydrology  since  the  last  glacial  maximum,  Journal  of  Sedimentary  Research,  72,  793-­‐808,  2002.  

Stern,  L.,  Banner,  J.  L.,  Cowan,  B.,  Copeland,  E.,  Mickler,  P.  J.,  Guifoyle,  A.,  James,  E.,  Musgrove,  M.,  and  Mack,  L.:  Trace  element  variations  in  speleothem  calcite:  influence  of  non-­‐environmental  factors,  Geol.  Soc.  Am.  Abstr.  Prog.,  193,  435,  2005.  

Suess,  H.:  Radiocarbon  concentration  in  modern  wood,  Science,  122,  415-­‐417,  1955.  

Talma,  A.  S.  and  Vogel,  J.  C.:  Late  Quaternary  paleotemperatures  derived  from  a  speleothem  from  Cango  Caves,  Cape  province,  South  Africa  Quaternary  Research,  37,  203-­‐213,  1992.  

Page 72: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 62  

Tarutani,  T.,  Clayton,  R.  N.,  and  Mayeda,  T.  K.:  Effect  of  polymorphism  and  magnesium  substitution  on  oxygen  isotope  fractionnation  between  calcium  carbonate  and  water,  Geochimica  Et  Cosmochimica  Acta,  33,  987-­‐&,  1969.  

Thompson,  P.,  Schwarcz,  H.  P.,  and  Ford,  D.  C.:  Continental  Pleistocene  climatic  variations  from  speleothen  age  and  isotopic  data  Science,  184,  893-­‐895,  1974.  

Tooth,  A.  F.  and  Fairchild,  I.  J.:  Soil  and  karst  aquifer  hydrological  controls  on  the  geochemical  evolution  of  speleothem-­‐forming  drip  waters,  Crag  Cave,  southwest  Ireland,  Journal  of  Hydrology,  273,  51-­‐68,  2003.  

Treble,  P.,  Shelley,  J.  M.  G.,  and  Chappell,  J.:  Comparison  of  high  resolution  sub-­‐annual  records  of  trace  elements  in  a  modern  (1911-­‐1992)  speleothem  with  instrumental  climate  data  from  southwest  Australia,  Earth  and  Planetary  Science  Letters,  216,  141-­‐153,  2003.  

Treble,  P.  C.,  Chappell,  J.,  Gagan,  M.  K.,  McKeegan,  K.  D.,  and  Harrison,  T.  M.:  In  situ  measurement  of  seasonal  delta  O-­‐18  variations  and  analysis  of  isotopic  trends  in  a  modem  speleothem  from  southwest  Australia,  Earth  and  Planetary  Science  Letters,  233,  17-­‐32,  2005a.  

Treble,  P.  C.,  Chappell,  J.,  and  Shelley,  J.  M.  G.:  Complex  speleothem  growth  processes  revealed  by  trace  element  mapping  and  scanning  electron  microscopy  of  annual  layers,  Geochimica  Et  Cosmochimica  Acta,  69,  4855-­‐4863,  2005b.  

Tremaine,  D.  M.  and  Froelich,  P.  N.:  Speleothem  trace  element  signatures:  A  hydrologic  geochemical  study  of  modern  cave  dripwaters  and  farmed  calcite,  Geochimica  Et  Cosmochimica  Acta,  121,  522-­‐545,  2013.  

Tremaine,  D.  M.,  Froelich,  P.  N.,  and  Wang,  Y.:  Speleothem  calcite  farmed  in  situ:  Modern  calibration  of  delta  O-­‐18  and  delta  C-­‐13  paleoclimate  proxies  in  a  continuously-­‐monitored  natural  cave  system,  Geochimica  Et  Cosmochimica  Acta,  75,  4929-­‐4950,  2011.  

Vaks,  A.,  Woodhead,  J.,  Bar-­‐Matthews,  M.,  Ayalon,  A.,  Cliff,  R.  A.,  Zilberman,  T.,  Matthews,  A.,  and  Frumkin,  A.:  Pliocene-­‐Pleistocene  climate  of  the  northern  margin  of  Saharan-­‐Arabian  Desert  recorded  in  speleothems  from  the  Negev  Desert,  Israel,  Earth  and  Planetary  Science  Letters,  368,  88-­‐100,  2013.  

van  Breukelen,  M.  R.,  Vonhof,  H.  B.,  Hellstrom,  J.  C.,  Wester,  W.  C.  G.,  and  Kroon,  D.:  Fossil  dripwater  in  stalagmites  reveals  Holocene  temperature  and  rainfall  variation  in  Amazonia,  Earth  and  Planetary  Science  Letters,  275,  54-­‐60,  2008.  

Van  Rampelbergh,  M.,  Fleitmann,  D.,  Verheyden,  S.,  Cheng,  H.,  Edwards,  L.,  De  Geest,  P.,  De  Vleeschouwer,  D.,  Burns,  S.  J.,  Matter,  A.,  Claeys,  P.,  and  Keppens,  E.:  Mid-­‐  to  late  Holocene  Indian  Ocean  Monsoon  variability  recorded  in  four  speleothems  from  Socotra  Island,  Yemen,  Quaternary  Science  Reviews,  65,  129-­‐142,  2013.  

Van  Rampelbergh,  M.,  Verheyden,  S.,  Allan,  M.,  Quinif,  Y.,  Keppens,  E.,  Cheng,  H.,  Edwards,  L.  R.,  and  Claeys,  P.:  A  500-­‐year  seasonally  resolved  δ18O  and  δ13C  and  

Page 73: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate  

  63  

layer  thickness  record  from  a  speleothem  of  the  Han-­‐sur-­‐Lesse  cave,  Belgium,  Climate  of  the  Past,  in  review.  in  review.  

Van  Rampelbergh,  M.,  Verheyden,  S.,  Allan,  M.,  Quinif,  Y.,  Keppens,  E.,  and  Claeys,  P.:  Seasonal  variations  recorded  in  cave  monitoring  results  and  a  10-­‐year  monthly  resolved  speleothem  δ18O  and  δ13C  record  from  the  Han-­‐sur-­‐Lesse  cave,  Belgium.,  Climate  of  the  Past,  10,  1-­‐15,  2014.  

Verheyden,  S.,  Genty,  D.,  Cattani,  O.,  and  van  Breukelen,  M.  R.:  Water  release  patterns  of  heated  speleothem  calcite  and  hydrogen  isotope  composition  of  fluid  inclusions,  Chemical  Geology,  247,  266-­‐281,  2008a.  

Verheyden,  S.,  Genty,  D.,  Deflandre,  G.,  Quinif,  Y.,  and  Keppens,  E.:  Monitoring  climatological,  hydrological  and  geochemical  parameters  in  the  Pere  Noel  cave  (Belgium):  implication  for  the  interpretation  of  speleothem  isotopic  and  geochemical  time-­‐series,  International  Journal  of  Speleology,  37,  221-­‐234,  2008b.  

Verheyden,  S.,  Keppens,  E.,  Quinif,  Y.,  Cheng,  H.  J.,  and  Edwards,  L.  R.:  Late-­‐glacial  and  Holocene  climate  reconstruction  as  inferred  from  a  stalagmite  -­‐  Grotte  du  Pere  Noel,  Han-­‐sur-­‐Lesse,  Belgium,  Geologica  Belgica,  17,  83-­‐89,  2014.  

Vonhof,  H.  B.,  van  Breukelen,  M.  R.,  Postma,  O.,  Rowe,  P.  J.,  Atkinson,  T.  C.,  and  Kroon,  D.:  A  continuous-­‐flow  crushing  device  for  on-­‐line  delta(2)H  analysis  of  fluid  inclusion  water  in  speleothems,  Rapid  Communications  in  Mass  Spectrometry,  20,  2553-­‐2558,  2006.  

Walker,  J.,  Cliff,  R.  A.,  and  Latham,  A.  G.:  U-­‐Pb  isotopic  age  of  the  StW  573  hominid  from  Sterkfontein,  South  Africa,  Science,  314,  1592-­‐1594,  2006.  

Wang,  X.  F.,  Auler,  A.  S.,  Edwards,  R.  L.,  Cheng,  H.,  Ito,  E.,  and  Solheid,  M.:  Interhemispheric  anti-­‐phasing  of  rainfall  during  the  last  glacial  period,  Quaternary  Science  Reviews,  25,  3391-­‐3403,  2006.  

Wang,  Y.  J.,  Cheng,  H.,  Edwards,  R.  L.,  An,  Z.  S.,  Wu,  J.  Y.,  Shen,  C.  C.,  and  Dorale,  J.  A.:  A  high-­‐resolution  absolute-­‐dated  Late  Pleistocene  monsoon  record  from  Hulu  Cave,  China,  Science,  294,  2345-­‐2348,  2001.  

Williams,  P.:  Th  erole  of  the  epikarst  in  karst  and  cave  hydrology:  a  review,  International  Journal  of  Speleology,  37,  1-­‐10,  2008.  

Williams,  P.  W.,  Marshall,  A.,  Ford,  D.  C.,  and  Jenkinson,  A.  V.:  Palaeoclimatic  interpretation  of  stable  isotope  data  from  holocene  speleothems  of  the  Waitomo  district,  North  Island,  New  Zealand,  Holocene,  9,  649-­‐657,  1999.  

Winograd,  I.  J.:  Evidence  from  uranium-­‐series-­‐dated  speleothems  for  the  timing  of  the  penultimate  deglaciation  of  northwestern  Europe,  Quaternary  Research,  58,  60-­‐61,  2002.  

Page 74: Maïté Van Rampelbergh

Chapter  2:  Speleothems  and  climate    

 64  

Winograd,  I.  J.,  Coplen,  T.  B.,  Landwehr,  J.  M.,  Riggs,  A.  C.,  Ludwig,  K.  R.,  Szabo,  B.  J.,  Kolesar,  P.  T.,  and  Revesz,  K.  M.:  Continuous  500,000-­‐year  climate  record  from  vein  calcite  in  Devils-­‐Hole,  Nevada,  Science,  258,  255-­‐260,  1992.  

Winograd,  I.  J.,  Landwehr,  J.  M.,  Coplen,  T.  B.,  Sharp,  W.  D.,  Riggs,  A.  C.,  Ludwig,  K.  R.,  and  Kolesar,  P.  T.:  Devils  Hole,  Nevada,  delta  O-­‐18  record  extended  to  the  mid-­‐Holocene,  Quaternary  Research,  66,  202-­‐212,  2006.  

Woodhead,  J.,  Hellstrom,  J.,  Maas,  R.,  Drysdale,  R.,  Zanchetta,  G.,  Devine,  P.,  and  Taylor,  E.:  U-­‐Pb  geochronology  of  speleothems  by  MC-­‐ICPMS,  Quaternary  Geochronology,  1,  208-­‐221,  2006.  

Yang,  W.  B.,  Lowenstein,  T.  K.,  Krouse,  H.  R.,  Spencer,  R.  J.,  and  Ku,  T.  L.:  A  200,000-­‐year  delta  O-­‐18  record  of  closed-­‐basin  lacustrine  calcite,  Death  Valley,  California,  Chemical  Geology,  216,  99-­‐111,  2005.  

Zeebe,  R.  E.,  Wolf-­‐Gladrow,  D.  A.,  and  Jansen,  H.:  On  the  time  required  to  establish  chemical  and  isotopic  equilibrium  in  the  carbon  dioxide  system  in  seawater,  Marine  Chemistry,  65,  135-­‐153,  1999.  

Zhang,  R.,  Schwarcz,  H.  P.,  Ford,  D.  C.,  Schroeder,  F.  S.,  and  Beddows,  P.  A.:  An  absolute  paleotemperature  record  from  10  to  6  Ka  inferred  from  fluid  inclusion  D/H  ratios  of  a  stalagmite  from  Vancouver  Island,  British  Columbia,  Canada,  Geochimica  Et  Cosmochimica  Acta,  72,  1014-­‐1026,  2008.  

Zhou,  H.  Y.,  Feng,  Y.  X.,  Zhao,  J.  X.,  Shen,  C.  C.,  You,  C.  F.,  and  Lin,  Y.:  Deglacial  variations  of  Sr  and  Sr-­‐87/Sr-­‐86  ratio  recorded  by  a  stalagmite  from  Central  China  and  their  association  with  past  climate  and  environment,  Chemical  Geology,  268,  233-­‐247,  2009.    

Page 75: Maïté Van Rampelbergh

Chapter  3:  Speleothems  from  Socotra  

  65  

Chapter  3        

Socotran   speleothems   reveal   monsoon  changes  during  the  Mid-­‐  to  Late-­‐Holocene      This  chapter  elaborates  on  the  current  knowledge  on  the  behavior  of  the  Indian  Ocean  Monsoon  (IOM)  during  the  Mid-­‐  to  Late-­‐Holocene  for  the  northern  Indian  Ocean.   Numerous   speleothem   studies   already   provide   information   on   the   IOM  variations   and   its   warm/wet   and   cold/dry   teleconnection   with   the   Northern  Atlantic  climate  (Burns  et  al.,  1998;  Burns  et  al.,  2001;  Neff  et  al.,  2001;  Burns  et  al.,  2003;  Fleitmann  et  al.,  2004;  Fleitmann  et  al.,  2007;  Shakun  et  al.,  2007).  The  North  Atlantic  and  the  IOM  are  linked  through  the  boreal  summer  position  of  the  Intertropical   Convergence   Zone   (ITCZ)   (Fig.   1).   During  warmer   North   Atlantic  periods,  stronger  northern  hemisphere  insolation  causes  the  summer  position  of  the   ITCZ   to   move   further   up   north,   which   causes   the   precipitation   amounts  brought  by  the  southwest  IOM  to  increase.  Since  the  Mid-­‐Holocene,  the  summer  position   of   the   ITCZ   is   gradually   moving   south   due   to   the   diminishing   boreal  summer   insolation   (Wanner  et   al.,   2006).  This   southern   retreat  of   the   summer  ITCZ   causes   a   decrease   in   summer   IOM   precipitation   since   8   ka.   However,   no  records  describing  the  evolution  of  the  other  IOM  subsystem,  being  the  northeast  winter  monsoon,  have  been  established  yet.            

Page 76: Maïté Van Rampelbergh

Chapter  3:  Speleothems  from  Socotra  

 66  

 Figure  1.  The  present-­‐day  location  of  the  Intertropical  Convergence  zone  (ITCZ).  During   boreal   summer,   increased   northern   hemisphere   insolation   causes   the  intensification   of   low   pressure   cells,   which   pull   the   ITCZ   to   the   north   until   it  reaches   its   northernmost   position   in   July.   During   boreal   winter,   the   pressure  gradients   reverse   and   the   ITCZ  moves   south   until   it   reaches   its   southernmost  position  in  December.  Socotra  Island  lies  in  the  upper  part  of  the  oscillation  belt  (red  star)  (adapted  from  www.meteoweb.eu).    The   Island   of   Socotra,   located   in   the   northern   Indian   Ocean   (Fig.1),   has   the  valuable  advantage  to  separate  the  rains  from  both  subsystems  of  the  IOM  due  to  the  ‘barrier’  action  of  the  Haggeher  Mountains.  Northeast  winter  monsoon  rains  mainly  affect  the  northern  and  eastern  side  of  the  island  between  September  and  December,  while  southwestern  summer  monsoon  rains  cause  wetter  conditions  between   May   and   June   only   on   the   southwestern   part   of   the   island.   Four  stalagmites   from   the   eastern   side   of   Socotra   cover   the   last   6   000   years   and  indicated   a   different   evolution   of   the   northeast   monsoon   compared   to   the  southwest   monsoon.   While,   the   southwest   monsoon   displays   a   gradual  precipitation   decrease,   the   northeast  monsoon  precipitation   intensity  weakens  between   6.0   and   3.8   ka,   and   remains   constant   afterwards   with   two  superimposed  drier  periods,  between  0  and  0.6  ka  and  from  2.2  to  3.8  ka.  No  link  can   be   established   between   the   winter   IOM   variations   and   the   Greenland   ice  cores.  More  high-­‐resolution  records  from  this  region  are  required  to  understand  the  exact  forcing  behind  the  northeast  monsoon  in  this  area.              

Page 77: Maïté Van Rampelbergh

Chapter  3:  Speleothems  from  Socotra  

  67  

REFERENCES    Burns,  S.  J.,  Fleitmann,  D.,  Matter,  A.,  Kramers,  J.,  and  Al-­‐Subbary,  A.  A.:  Indian  Ocean  climate  and  an  absolute  chronology  over  Dansgaard/Oeschger  events  9  to  13,  Science,  301,  1365-­‐1367,  2003.  

Burns,  S.  J.,  Fleitmann,  D.,  Matter,  A.,  Neff,  U.,  and  Mangini,  A.:  Speleothem  evidence  from  Oman  for  continental  pluvial  events  during  interglacial  periods,  Geology,  29,  623-­‐626,  2001.  

Burns,  S.  J.,  Matter,  A.,  Frank,  N.,  and  Mangini,  A.:  Speleothem-­‐based  paleoclimate  record  from  northern  Oman,  Geology,  26,  499-­‐502,  1998.  

Fleitmann,  D.,  Burns,  S.  J.,  Mangini,  A.,  Mudelsee,  M.,  Kramers,  J.,  Villa,  I.,  Neff,  U.,  Al-­‐Subbary,  A.  A.,  Buettner,  A.,  Hippler,  D.,  and  Matter,  A.:  Holocene  ITCZ  and  Indian  monsoon  dynamics  recorded  in  stalagmites  from  Oman  and  Yemen  (Socotra),  Quaternary  Science  Reviews,  26,  170-­‐188,  2007.  

Fleitmann,  D.,  Burns,  S.  J.,  Neff,  U.,  Mudelsee,  M.,  Mangini,  A.,  and  Matter,  A.:  Palaeoclimatic  interpretation  of  high-­‐resolution  oxygen  isotope  profiles  derived  from  annually  laminated  speleothems  from  Southern  Oman,  Quaternary  Science  Reviews,  23,  935-­‐945,  2004.  

Neff,  U.,  Burns,  S.  J.,  Mangini,  A.,  Mudelsee,  M.,  Fleitmann,  D.,  and  Matter,  A.:  Strong  coherence  between  solar  variability  and  the  monsoon  in  Oman  between  9  and  6  kyr  ago,  Nature,  411,  290-­‐293,  2001.  

Shakun,  J.  D.,  Burns,  S.  J.,  Fleitmann,  D.,  Kramers,  J.,  Matter,  A.,  and  Al-­‐Subary,  A.:  A  high-­‐resolution,  absolute-­‐dated  deglacial  speleothem  record  of  Indian  Ocean  climate  from  Socotra  Island,  Yemen,  Earth  and  Planetary  Science  Letters,  259,  442-­‐456,  2007.                                  

Page 78: Maïté Van Rampelbergh

Chapter  3:  Speleothems  from  Socotra  

 68  

                     

Page 79: Maïté Van Rampelbergh

Mid- to late Holocene Indian Ocean Monsoon variability recorded infour speleothems from Socotra Island, Yemen

Maïté Van Rampelbergh a,*, Dominik Fleitmann b,c, Sophie Verheyden a,d, Hai Cheng e,f,Lawrence Edwards f, Peter De Geest g, David De Vleeschouwer a, Stephen J. Burns h, Albert Matter b,Philippe Claeys a, Eddy Keppens a

a Earth System Sciences Department, Vrije Universiteit Brussel (VUB), Pleinlaan 2, B-1050 Brussels, Belgiumb Institute of Geosciences, University of Bern, Baltzerstrasse 1-3, CH-3012 Bern, SwitzerlandcOeschger Centre for Climate Change Research, University of Bern, Zähringerstrasse 25, CH-3012 Bern, SwitzerlanddGeological Survey of Belgium, Royal Belgian Institute of Natural Sciences, Jennerstraat 13, B-1000 Brussels, Belgiume Institute of Global Environmental Change, Xi’an Jiaotong University, Xi’an 710049, ChinafDepartment of Geological Sciences, University of Minnesota, 100 Union Street SE, Minneapolis, MN 55455, USAgBijlokestraat 57, B-9070 Destelbergen, BelgiumhDepartment of Geosciences, University of Massachusetts, Morrill Science Center 233, Amherst, MA 01003, USA

a r t i c l e i n f o

Article history:Received 10 May 2012Received in revised form11 January 2013Accepted 17 January 2013Available online

Keywords:SpeleothemsIndian Ocean MonsoonSocotraPaleoclimateStable isotopesTrace elements

a b s t r a c t

Four stalagmites covering the last 7.0 ka were sampled on Socotra, an island in the northern Indian Oceanto investigate the evolution of the northeast Indian Ocean Monsoon (IOM) since the mid Holocene. OnSocotra, rain is delivered at the start of the southwest IOM in MayeJune and at the start of the northeastIOM from September to December. The Haggeher Mountains act as a barrier forcing precipitationbrought by the northeast winds to fall preferentially on the eastern side of the island, where the studiedcaves are located. d18O and d13C and Mg/Ca and Sr/Ca signals in the stalagmites reflect precipitationamounts brought by the northeast winds. For stalagmite STM6, this amount effect is amplified by kineticeffects during calcite deposition. Combined interpretation of the stalagmites’ signals suggest a weaken-ing of the northeast precipitation between 6.0 and 3.8 ka. After 3.8 ka precipitation intensities remainconstant with two superimposed drier periods, between 0 and 0.6 ka and from 2.2 to 3.8 ka. No link canbe established with Greenland ice cores and with the summer IOM variability.

In contrast to the stable northeast rainy season suggested by the records in this study, speleothemrecords from western Socotra indicate a wettening of the southwest rainy season on Socotra after 4.4 ka.The local wettening of western Socotra could relate to a more southerly path (more over the IndianOcean) taken by the southwest winds. Stalagmite STM5, sampled at the fringe between both rain areasdisplays intermediate d18O values. After 6.2 ka, similar precipitation changes are seen between easternSocotra and northern Oman indicating that both regions are affected similarly by the monsoon. Differentpalaeoclimatologic records from the Arabian Peninsula currently located outside the ITCZ migrationpathway display an abrupt drying around 6 ka due to their disconnection from the southwest rain in-fluence. Records that are nowadays still receiving rain by the southwest winds, suggest a more gradualdrying reflecting the weakening of the southwest monsoon.

! 2013 Elsevier Ltd. All rights reserved.

1. Introduction

The seasonal migration of the Intertropical Convergence Zone(ITCZ) and coupled monsoon systems influences more than half of

the world population. For the countries around the Arabian Sea, atthe northern limit of the ITCZ-pathway, a good understanding ofthe Indian Ocean Monsoon (IOM) and its summer and wintersubsystems is of major importance, especially considering thepredicted further drying (Fleitmann et al., 2007; Kropelin et al.,2008).

Around thenorthern IndianOcean, informationonchanges in theIOM wind direction and strength during the late-Pleistocene and

* Corresponding author. Tel.: !32 2 6293397; fax: !32 2 6293391.E-mail address: [email protected] (M. Van Rampelbergh).

Contents lists available at SciVerse ScienceDirect

Quaternary Science Reviews

journal homepage: www.elsevier .com/locate/quascirev

0277-3791/$ e see front matter ! 2013 Elsevier Ltd. All rights reserved.http://dx.doi.org/10.1016/j.quascirev.2013.01.016

Quaternary Science Reviews 65 (2013) 129e142

Maite Van Rampelbergh
69
Maite Van Rampelbergh
Maite Van Rampelbergh
Maite Van Rampelbergh
Chapter 3: Spelothems from Socotra
Page 80: Maïté Van Rampelbergh

Holocene have mostly been extracted from Indian Ocean sedimentcores (Sirocko et al., 1993; Gupta et al., 2003; Ivanochko et al., 2005)and fromdunedeposits in the southof theArabianPeninsula (Radieset al., 2005; Parker et al., 2006; Lezine et al., 2010). However, in thesearchives, radiocarbon ages with relative high uncertainties impedethe precise determination of the timing and duration of IOM varia-tions. Tree rings, allowing higher resolution dates, have proveduseful around the Indian Ocean, but they are limited to the last 1000years (Cook et al., 2010). The lacking information on hydrologicalpatterns can be retrieved from speleothems (Neff et al., 2001;Wanget al., 2001; Fleitmann et al., 2003a; Shakun et al., 2007). Speleo-thems have a relatively fast and often continuous growth over longtime periods allowing the elaboration of long-term records at highresolution (McDermott, 2004; Fairchild et al., 2006).Most important,speleothems can be precisely dated both by counting annual layers(if present) and/or by using the U/Th-dating method, which makesthem powerful archives to date important climatic, historical orcultural eventsprecisely (Wanget al., 2001;Henderson, 2006; Zhanget al., 2008; Cheng et al., 2009a).

Most of available Holocene palaeoclimatological records forthe area reflect variations in the summer IOM subsystem becausethe highest amounts of precipitation are associated with thesummer IOM, also known as the Asian summer monsoon. Con-sequently, the precipitation signals brought by the winter ornortheast IOM subsystem are often overwritten by the southwestsignal.

In this study, we present four new high-resolution stalagmiterecords sampled on the eastern side of Socotra Island (Yemen)(Fig. 1). Rainfall on the eastern side of the island mainly consists ofnortheast winter monsoon precipitation due to orographic effects(Scholte and De Geest, 2010). Eastern Socotra constitutes thereforean ideal location to study the changes in the northeast or winterIOM subsystem (source, directions, amounts). Different studiesaround the Arabian Sea have already shown that speleothem d18Ovalues reflect changes in palaeo-precipitation intensities(Fleitmann et al., 2003b, 2004a, 2007; Shakun et al., 2007). In thisstudy, several proxies (d18O, d13C, Mg/Ca and Sr/Ca ratios) aremeasured on selected stalagmites to complete the current under-standing of the northeast IOM subsystem around the NorthernIndian Ocean during the mid- to late Holocene (7 kae0 ka). A betterinsight in the mid- to late Holocene palaeoclimatic and environ-mental evolution of Socotra constitutes an important step to un-derstand regional climate dynamics.

2. Regional setting

2.1. Site location and present climate

Socotra Island lies in the northwestern part of the Indian Ocean,between the horn of Africa and the Arabic Peninsula (Fig. 1). Meanannual rainfall and temperature measured by a network of 11meteorological stations from 2002 to 2006, are 216mm and 28.9 "Crespectively, referring to a semi-arid tropical climate (Scholte andDe Geest, 2010) (see also Fig. 1). The present-day climate onSocotra Island is governed by the seasonal migration of the ITCZand episodic passages of tropical cyclones as well as oceaneatmosphere interactions, such as the Indian Ocean Dipole (IOD)and the El Nino-Southern Oscillation (ENSO) (Cheung et al., 2006).From the western Pacific to the Indian Ocean, the ITCZ differs fromits traditional notion as a narrowwell-defined cloud band. The ITCZabove the Indian Ocean is broader in latitude with significantlymore horizontal spatial variation (Waliser and Gautier, 1993).Because of its atypical configuration, discussion can arise on theexact use of the term ITCZ in region around the northern IndianOcean. In this study, the ITCZ forms part of the ascending branch ofthe Hadley cell and is defined as the convergence zone betweensouthwest and northeast winds where precipitation is fairly high.This definition is similar to the ITCZ definition used in Fleitmannet al. (2003a, 2007).

In boreal summer, the low-pressure cell centred above thesouthern foothills of the Tibetan plateau pulls the ITCZ north until itreaches its northernmost position in July. During boreal autumn,the pressure gradients reverse and the ITCZ retreats southwarduntil it reaches its southernmost position in January (Fleitmannet al., 2004b). This annual migration of the ITCZ creates two windpatterns on Socotra, known as the southwest summer and thenortheast winter monsoon period interrupted by two short tran-sition periods whenwinds from all directions ensue. The southwestmonsoon season begins in early May (Fig. 2a) with a stablesouthwesterly and moisture loaden wind (course w230") comingfrom the Indian Ocean. In July, the ITCZ reaches its northernmostposition, wind speed increases and the airflow takes a morewesterly path (w240") across the semi-desert of Somalia. Strongand dry southwest monsoonal winds blow over the island. InSeptember at the end of the summer monsoon season, wind di-rection swings back to w230", wind speed decreases and windsfilled with moisture from the Indian Ocean deliver precipitation

Fig. 1. Location of the island of Socotra in the northern Indian Ocean. Black dots represent the location of the caves. The numbers represent the location of the 11 meteorologicalstations studied in Scholte and De Geest (2010). The watershed is represented by the dotted line.

M. Van Rampelbergh et al. / Quaternary Science Reviews 65 (2013) 129e142130

Maite Van Rampelbergh
Maite Van Rampelbergh
70
Maite Van Rampelbergh
Maite Van Rampelbergh
Maite Van Rampelbergh
Maite Van Rampelbergh
Chapter 3: Spelothems from Socotra
Page 81: Maïté Van Rampelbergh

(Culek et al., 2006). After an autumn transition period starting inOctober, the northeast monsoon starts in the first half of November(Fig. 2b). The northeast monsoon onset, however, is not as suddenas that of the southwest summer monsoon due to the lower

pressure gradient between the high-pressure cell above the Tibetanplateau and the high-pressure cell near Madagascar. During thisperiod, rainfall reaches a maximum averaging of 120 mm (or 42% ofthe mean annual rainfall) as winds transport moisture originating

Fig. 2. a) The black star indicates the location of Socotra. Wind patterns and ITCZ location for the northern Indian Ocean during (1) northeast winter monsoon in January, (2)southwest rainy season in May, (3) southwest summer monsoon in July and (4) northeast rainy season in November (adapted after Fleitmann et al., 2004a). b) Socotra’s monsoonand precipitation intensities for every month derived from NASA’s Tropical Rainfall Measuring Mission. Purple bars indicate percentages of the mean annual rainfall of 100.3 mm.The black and dotted lines indicate the percentage of monsoon strength for both monsoon periods. Grey bars indicate the transition periods.

M. Van Rampelbergh et al. / Quaternary Science Reviews 65 (2013) 129e142 131

Maite Van Rampelbergh
Maite Van Rampelbergh
71
Maite Van Rampelbergh
Maite Van Rampelbergh
Chapter 3: Spelothems from Socotra
Maite Van Rampelbergh
Page 82: Maïté Van Rampelbergh

from the warm Arabian Sea towards Socotra Island (Scholte and DeGeest, 2010). In early February, the ITCZ starts tomigrate back northand the northeast monsoon weakens. The spring transition occursat the end of March with winds blowing from both northwesternand southwestern directions. At the beginning of April, southwestwinds become dominant again, cumulus clouds form and bring rain(Culek et al., 2006). In summary, Socotra experiences two distinctrainy seasons (typically from April to June and from November toDecember) interrupted by two dry seasons (typically from Januaryto March and June to October) (Fig. 2b). Since the ITCZ is defined asthe convergence zone between northeast and southwest winds, thetwo Socotran rainy seasons are associated with the passage of theITCZ over the island.

Based on detailed analyses of cloud cover satellite images forSocotra Island (Fig. 1), the northern and southern plateaus experi-ence different precipitation regimes (Scholte and De Geest, 2010).The 400e600 m high limestone cliffs at the northern and southerncoast and elevated plateaus around the Haggeher Mountains causeorographic uplift and induce two distinct rain areas on both sides ofthemountain range. The northern and eastern parts receive most oftheir rain during the northeast rainy season in November (Fig. 1)whereas the western and southern coasts receive almost equalamounts of rain during both rainy seasons (Scholte and De Geest,2010). Sporadically, during strong southwest spring rainy years,the northern regions, which are normally located in the rainshadow of the Haggeher Mountains, can be influenced by south-west rains (Culek et al., 2006).

Based on the geology and geomorphology, Socotra Island issubdivided into three zones: the Quaternary alluvial coastal andinland plains, the PalaeoceneeEocene reef-limestone plateaus andthe Precambrian granitic Haggeher Mountains (Cheung et al.,2006). The limestone plateaus, covering approximately half ofSocotra’s surface, are strongly karstified and harbour numerouslarge cave systems (Cheung et al., 2006). The main karst areas arethe Momi karst plateau in the east, the Diksam/Sibehon karst pla-teau in the centre and the Ma’alah karst plateau in the north-western part of the island (Fig. 1).

Three stalagmites were sampled in Hoq Cave (12"35011.900N;54"21015.4400E, elevation 335 masl), located on the Momi karstplateau, 5 km from the northeast coast (Fig. 1). The entrance, fullyfacing the seaside, is around 45 m wide and 30 m high. The firststalagmite (specimen Hq1) was sampled in 2000, 200 m from theentrance in a chamber that often experienced strong ventilationand varying humidity, as observed during fieldwork. Two coevalstalagmites, STM1 and STM6, were sampled 1 m next to each otherin 2003 and 2006 respectively, approximately 2 km from the onlyknown entrance, where ventilation is minimal. At this site, cave airtemperature remained constant throughout the year at 25 # 0.5 "C(continuously monitored between January and December 2003).Relative humidity measured during six visits between 2003 and2006 was always higher than 98%.

Stalagmite STM5 was collected from Casecas Cave(12"33020.0200Ne54"18033.3400E, elevation 542 masl) in January2004, 6 km southwest from Hoq Cave (Fig. 1), also on the Momikarst, in an upper gallery where ventilation is expected to beminimal. Temperature and humidity measured during samplingwere 29 # 0.5 "C and above 95%.

2.2. Palaeoclimate of the region

Changes in palaeorainfall in the areas located at the northernfringe of the Indian Ocean Monsoon domain are generallyexplained by variations in the strength of the IOM and the asso-ciated boreal summer position of the ITCZ (Fleitmann et al., 2003a,2007). Various Late Pleistocene palaeoclimatic studies using

different archives from northeast Africa (Gasse, 2000), southernArabia (Burns et al., 1998, 2001; Preusser et al., 2002; Fleitmannet al., 2003a, 2003b, 2004a; Parker et al., 2006; Fleitmann et al.,2007; Fuchs and Buerkert, 2008; Lezine et al., 2010), the ArabianSea (Gupta et al., 2003) and China (Dykoski et al., 2005; Wang et al.,2005), demonstrate that changes in the position of the ITCZ aregoing along with variations in IOM rainfall intensity. During theGlacial to Lateglacial, the ITCZwas considered to be located south ofthe Arabian Peninsula and IOM rainfall did not reach southernArabia as indicated by the absence of speleothem growth (Burnset al., 2001; Fleitmann et al., 2007), lack of lacustrine deposits(Lezine et al., 2010) and homogenous sedimentation rates of aeo-lian sediments (Fuchs and Buerkert, 2008). A northward shift in themean latitudinal position of the summer ITCZ and orbitally forcedintensification of the IOM during the early Holocene led to theonset of a humid period. This Holocene wet optimum is welldocumented in northeast Africa and southern Arabia and charac-terized by widespread formation of lakes (Gasse, 2000) andenhanced speleothems deposition (high effective moisture) (Burnset al., 1998, 2001; Fleitmann et al., 2003a, 2007). The termination ofthe Holocene wet optimum is associated with a southward dis-placement of the ITCZ to its present-day position along the coast ofsouthern Arabia (Fleitmann et al., 2007). Discrepancies in thetiming and duration of this humid period remain. Speleothemsfrom Oman and Yemen suggest an onset at 10.5 ka whereas datafrom Yemen (Lezine et al., 2007) suggest an onset at 12 ka or in theUAE an even later onset at 8.5 ka (Parker et al., 2006). However,most of the studies from southern Arabia confirm a period ofmaximal rainfall at 8 ka. According to sediment records fromnorthern Oman (Fuchs and Buerkert, 2008), speleothems fromChina (Dykoski et al., 2005; Wang et al., 2005), Oman (Fleitmannet al., 2003a, 2007) and the findings of Lézine et al. (2010) inYemen, a gradual long term decrease in precipitation starts shortafter 8 ka. In contrast, Burns et al. (2001) identified a reduction inrainfall at ca 6 ka in northern Oman as well as in the UAE whereParker et al. (2006) placed the Holocene wet period termination at6 ka. These partially contrary results from various studies might bedue to differences in the sensitivity of the investigated archives andto a more complex pattern of atmospheric conditions at a localscale, as will also be shown further in this study.

Contrary tothe long termdecrease inprecipitationon theArabianPeninsula, a speleothem from western Socotra (Dimarshim Cave;Fig. 1) shows increasing humid conditions since 4.4 ka (Fleitmannet al., 2007). This anticorrelation between western Socotra and theArabian Peninsula is explained by Fleitmann et al. (2007) as a resultof a progressing southward displacement in the mean latitudinalsummer position of the ITCZ, implying a decrease in precipitation inthe areas located at the northern fringe of the IOM, but an increase inareas closer to the equator (Fleitmann et al., 2007).

3. Materials and methods

For analyses, a slab of 1 cm was cut from the middle of eachstalagmite parallel to its growth axis using a diamond saw. Theslabs were polished using carbide powder and finished with Al2O3-powder. Twenty-six (13 for STM1, 7 for STM6 and 6 for STM6) U-series age determinations were carried out at the University ofMinnesota (USA), using the procedures for uranium and thoriumchemical separation and purification described in Edwards et al.(1987) and Cheng et al. (2000, 2009a, 2009b). Age determinationson Hq1 were carried out at the University of Bern. Details on ana-lytical methods are given in Fleitmann et al. (2007). Age models areestablished using the StalAge algorithm (Scholz and Hoffmann,2011) and are reported with their uncertainty. All ages areexpressed in a B2011.

M. Van Rampelbergh et al. / Quaternary Science Reviews 65 (2013) 129e142132

Maite Van Rampelbergh
72
Maite Van Rampelbergh
Chapter 3: Spelothems from Socotra
Page 83: Maïté Van Rampelbergh

Samples for d13C and d18O measurements in STM1, STM5 andSTM6 were drilled along the stalagmites central axis with a Mer-chantec Micromill. Ethanol was used to clean the speleothem sur-face and drill bit prior to sampling. Between every sampling, thedrill bit and sampling surface were blown clean with compressedair. Sample resolution in STM1 was 500 mm for the upper 32 mm,from 48 to 51 mm and from 177 to 248 mm. The remaining partswere sampled at a 1 mm resolution. Stalagmites STM5 and STM6were sampled every 500 mm. Stable isotope sampling in the Hq1stalagmite was carried out in the upper 305mm every 600 mm. OneHendy test was carried out on STM1 and three on STM5 (Hendy,1971) by drilling samples along an individual growth layer. Toobtain samples of modern precipitated calcite, 6 glass slabs wereplaced in Hoq Cave for approximately one year between January2003 and May 2005; 2 slabs rested on top of STM6 and 4 at the endof the cave. Stable isotope analyses on STM1, STM5, STM6 and onfresh calcite from the glass slabs were carried out at the VrijeUniversiteit Brussel with a Kiel-III-device coupled on a ThermoDelta plus XL. Analytical uncertainties (1s) were $0.06& for d13Cand $0.08& for d18O.

For isotopic analyses on stalagmite Hq1 approximately 5 mg ofpowder was drilled from the sample and analysed with an on-line,automated, carbonate preparation system linked to a VG Prism IIisotope ratio mass spectrometer at the University of Bern. Repro-ducibility of standardmaterials is 0.08& (1s). All isotopic values arereported in per mille (&) relative to Vienna Pee Dee Belemnite(VPDB).

Samples for elemental concentration determination of Ca, Mgand Sr were taken every 5 mm along the growth axis of STM1 andSTM6 using a carbide dental drill (1 mm diameter). A total of 111samples (ca 15 mg) for STM1 were analysed by Atomic AbsorptionSpectrometry at the Vrije Universiteit Brussel with analytical un-certainties (2s) less than 5%. 32 samples (ca 5 mg) for STM6 weremeasured on an Element 2 HR-ICP-MS at the Royal Museum forCentral Africa (Brussels, Belgium) with analytical uncertainties (2s)less than 5%.

Three seepagewater samples from Casecas Cave and 21 samplesfrom Hoq were collected for d18O measurements. The water sam-ples were prepared using the CO2/H2O-equilibration methoddescribed by Epstein and Mayeda (1953). Measurements wereperformed on a Finnigan Delta E mass spectrometer at the VrijeUniversiteit Brussel. All values are reported in per mill (&) relativeto Standard Mean Ocean Water (SMOW). Analytical uncertainties(2s) were less than 0.10&.

4. Results

4.1. Hoq Cave

The stable isotopic compositions of calcite deposited on glassslabs in the deepest parts of the cave (near STM6) display largevariations between %2.15& and %4.22& for d18O and between%3.18& and %7.86& for d13C (Table 1). No link can be establishedbetween the measured isotopic values and the slabs location in thecave. A similar large range in d18O values is also observed in the 21seepage waters ranging between %1.36& and %4.26& (Fig. 3).

Stalagmite Hq1, sampled only 200 m from the wide entrance, isa 381 mm tall stalagmite that was still actively dripping whencollected in 2000. Nine U-series dates carried out on the upper290mm (Table 2a) indicate continuous growth at an average rate of32 mm/yr. According the results given by StalAge, large age un-certainties occur after 6.9 ka (Fig. 4). Because the timing of events islargely insecure after 6.9 ka, no climatic interpretations are basedon that part. The d18O values vary between %3.06& and %0.70&around an average of%1.69& (Fig. 5). The d13C values vary between

%8.69& and 2.41& around an average of %0.50& and correlatesignificantlywith the d18O values (r& 0.83 and p& 3.5758' 10%129).Two periods of very positive isotopic values occur from 0.7 to 1.5 kaand from 3.9 until 6.9 ka.

Stalagmites STM1 (length 520 mm, sampled January 2003) andSTM6 (length 160 mm, sampled January 2006) were activelydripping when sampled at the end of the cave. Calcite fabric in bothstalagmites varies between dark compact and white porous parts,with lamination only visible in the white porous parts. In STM1, at10 cm from the base, a shift in growth axis occurs (Fig. 4). Based on13 U-series dates (Table 2b) and the StalAge age model STM1 grewconstantly since 5.6 ka at a rate of around 87 mm/yr. In STM6, 7 U/Th-ages (Table 2b) show that the stalagmite grew constantly from4.5 ka (Fig. 4) with an average growth rate of 34 mm/yr, which issimilar to the growth rate of stalagmite Hq1 located about 2 kmaway, and three times slower than stalagmite STM1 only 1 m away.

The STM1 d18O values vary between %4.24 and %0.84&, arounda mean of %2.97& (Fig. 5). Similar d18O values are found in STM6varying between %5.06& and %0.60& (average at %2.27&). Thed13C values of STM1 vary between%12.11& and%3.64&, average at%8.41&. For STM6, the d13C values are slightly more positive,varying between %10.58& and %3.54& around an average at%5.88&. A good correlation can be established between the d18Oand d13C profiles in STM1 (r & 0.70 and p & 1.7988 ' 10%93) and inSTM6 (r & 0.55 and p & 8.6530 ' 10%26). Compared to stalagmiteHq1, sampled close to the cave entrance, STM1 and STM6 displaygenerally more negative d18O and d13C values. From the start of therecords until 3.8 ka, the STM1 and STM6 isotopic records display

Table 1Location of the 6 glass slabs in the cave, period of calcite deposition and d18O andd13C values of the fresh present-day calcite.

Fresh calcite from glass slabs

Location Period d13C(% VPDB)

d18O(& VPDB)

Start Stop

STM6 January 2003 January 2004 %7.7 %2.5STM6 November 2004 January 2006 %4.5 %2.1END October 2003 November 2004 %7.2 %3.4END October 2003 November 2004 %3.2 %2.2END November 2004 May 2005 %7.9 %4.2END November 2004 May 2005 %7.3 %4.1

Fig. 3. d18O composition of the 21 Hoq Cave (black dots) and 3 Casecas Cave (whitedots) seepage waters. Most negative values occur in November, when Socotra receivesmost of its rain, suggesting that the drip water d18O values are influenced by the‘amount effect’.

M. Van Rampelbergh et al. / Quaternary Science Reviews 65 (2013) 129e142 133

Maite Van Rampelbergh
Maite Van Rampelbergh
73
Maite Van Rampelbergh
Maite Van Rampelbergh
Maite Van Rampelbergh
Chapter 3: Spelothems from Socotra
Maite Van Rampelbergh
Page 84: Maïté Van Rampelbergh

decreasing values with similar centennial andmillennial variations.After 3.8 ka, the d18O signal of both stalagmites varies around%2.5& without significant trend. Between 0 and 0.6 ka and from2.2 to 3.8 ka the STM6 record shifts to more positive values.

The Mg/Ca ratios ('103) of STM1 range from 7.03 to 13.83, withan average of 9.40 (Fig. 5). The STM1 Sr/Ca ratios ('103) vary be-tween 0.20 and 0.34 averaging 0.27 and correlate successfully withtheMg/Ca ratios (r& 0.65 and p& 6.5197'10%15) (Fig. 5). For STM6,similar conclusions can be established; the Mg/Ca ('103) valuesvary between 7.74 and 13.03 and average at 10.69. The Sr/Ca ('103)values vary between 0.21 and 0.31 and average at 0.26 and correlatewell with the Mg/Ca ('103) profile (r & 0.72, p & 6.6556 ' 10%6).

For all three stalagmites, calcite colour, growth rate and isotopicsignals appear to correlate, with darker compact calcite (indicative

of slowgrowth) coincidingwithmore positive d18O and d13C values.Lighter calcite is formed during faster growth periods and is char-acterized by more negative d18O and d13C values. In the coevalstalagmites STM1 and STM6, higher trace elemental concentrationscorrespond to darker calcite, slower growth rate and more positiveisotopic values. Due to the low-resolution trace elemental mea-surements in STM6, the covariationwith the d18O and d13C values isless clear.

4.2. Casecas Cave

Three seepage water samples were collected from Casecas Caveduring the dry winter monsoon season in January 2003 and 2004.Their d18O values vary between%1.97& and%3.01& (Fig. 3) and areslightly more negative than the January d18O values of Hoq Cave.STM5 was continuously deposited between January 2004 (date ofsampling) and 0.7 ka as is shown by the six 230Th ages (Table 2b).

Table 2aU/Th measurements for stalagmite Hq1 (University of Bern). All ages were converted to before 2011

Samplenumber

Depth(mm)

c (U) (ppb) c (Th) (ppb) 234U/238U 230Th/232Th 230Th/234U Age (ka)a Age (ka B2011)

Hq1-1 13.1 20,321.1 # 52.51 0.3160 # 0.0061 1.0000 # 0.0006 349.6907 # 7.4303 0.0018 # 0.0000 0.198 # 0.02 0.207 # 0.02Hq1-2 38.9 22,029.6 # 58.77 0.4610 # 0.0222 0.9931 # 0.0007 808.1328 # 39.1271 0.0056 # 0.0000 0.610 # 0.02 0.619 # 0.02Hq1-3 67.1 10,921.3 # 28.55 0.3595 # 0.0027 1.0136 # 0.0009 1341.2558 # 14.8675 0.0143 # 0.0001 1.572 # 0.03 1.581 # 0.03Hq1-4 110.3 22,383.2 # 58.84 2.4010 # 0.0133 1.0166 # 0.0009 602.1049 # 6.4128 0.0210 # 0.0002 2.319 # 0.03 2.328 # 0.03Hq1-5 157.1 13,513.6 # 35.45 6.5158 # 0.0354 1.0284 # 0.0009 195.7649 # 1.4870 0.0304 # 0.0002 3.357 # 0.03 3.366 # 0.03Hq1-6 191.9 8725.9 # 22.46 6.0375 # 0.0308 1.0007 # 0.0009 167.6249 # 1.2385 0.0384 # 0.0002 4.227 # 0.05 4.236 # 0.05Hq1-7 220.l 8048.1 # 21.26 6.1969 # 0.0445 1.0123 # 0.0010 196.0517 # 1.7976 0.0494 # 0.0003 5.524 # 0.06 5.533 # 0.06Hq1-8 250.1 12,223.9 # 31.18 4.5306 # 0.0248 1.0234 # 0.0006 514.2779 # 4.0313 0.0615 # 0.0004 6.927 # 0.06 6.936 # 0.06Hq1-9 290.3 18,368.8 # 47.14 1.0082 # 0.0060 1.0216 # 0.0007 3933.7280 # 34.9869 0.0686 # 0.0005 7.760 # 0.09 7.769 # 0.09

a Ages relative to AD 2002.

Fig. 4. Age versus depth plots and average growth rate of the 4 studied stalagmites.Black line represent the StalAge (Scholz and Hoffmann, 2011) age models. Grey linesindicate the uncertainties as modelled by StalAge.

Fig. 5. d18O (black line) and d13C (grey line) values in & VPDB for the studied sta-lagmites plotted against age in ka before 2011. Dots with error bars mark the U/Th ageswith their uncertainty. For STM1 and STM6, the Mg/Ca ' 103 record is indicated inblack and the Sr/Ca ' 103 record is indicated in grey.

M. Van Rampelbergh et al. / Quaternary Science Reviews 65 (2013) 129e142134

Maite Van Rampelbergh
74
Maite Van Rampelbergh
Chapter 3: Spelothems from Socotra
Maite Van Rampelbergh
Page 85: Maïté Van Rampelbergh

STM5’s growth rate averages 123 mm/yr making this stalagmite thefastest growing speleothem of our four samples (Fig. 4). The large U/Th-error bars, especially in U/Th-samples STM5-1 and STM5-3, aredue to the high amounts of detrital 232Th, leading to large un-certainties in the age model (Fig. 4). Therefore, this record will onlybe used to discuss multi-millennial variations. STM 5’s d18O values,averaging%3.24&, range from%4.24& to%1.41& (Fig. 5). As for theseepage waters, these d18O values in STM5 are more negativecompared to the d18O values found in the Hoq Cave stalagmites. Thed13C values of STM5 range from %7.40& to %2.33& and average at%5.26&. They are more positive compared to the d13C values of theHoq Cave speleothems. A good correlation can be established be-tween the d18O and the d13C signal (r& 0.76 and p& 4.9724'10%34).

5. Discussion

5.1. Low d18O and d13C values indicate wetter conditions

Provided that calcite formed under conditions of isotopic equi-librium, the d18O of speleothem calcite is governed by the waterecalcite fractionation factor (0.25& decrease per 1 "C increase),and by the d18O of the cave-seepage water that in turn is deter-mined by the d18O of rainwater (Lachniet, 2009). In tropical andsubtropical areas, such as Socotra, the “amount effect”, describingthe inverse relationship between the amount of precipitation andits oxygen isotopic composition, is mainly responsible for changesin rainwater d18O (Dansgaard, 1964; Rozanski et al., 1992) andconsequently for the d18O composition of cave seepage waters. InSocotra, changes in the d18O of stalagmites deposited in or close to

isotopic equilibrium with their seepage waters, reflect fluctuationsin the amount of precipitation. In this study, the presence of the“amount effect” is clearly demonstrated by more negative d18Ovalues of the seepagewaters in November (Fig. 3), whenmost of therain falls on the island. The rest of the year, rainwater d18O valuesdisplay less negative values, independent of source of the rainindicating the absence of a strong ‘source effect’. Other studiesusing Socotran speleothems also interpreted the changes in d18Ocomposition as reflecting the “amount effect” with more negatived18O values occurring during wetter conditions (Burns et al., 2003;Fleitmann et al., 2007; Shakun et al., 2007).

The d13C of speleothems deposited in equilibrium is mainlydetermined by the isotopic composition of soil-CO2, which reachesthe cavewith the seepagewater and normally constitutes themajorcarbon source (Genty et al., 2001). Variations in the d13C compo-sition of soil-CO2 are mainly controlled by changes in the type ofvegetation cover in terms of C3/C4/CAM-plants above the cave(Smith and Epstein, 1970; Frumkin et al., 2000). If no major vege-tation changes occurred over the studied period, as is most likelythe case in this study, variations in the soil-CO2 d13C are primarilyrelated to the intensity of soil activity with heavier d13C valuesduring drier periods (Genty et al., 2003).

If the stalagmites were not deposited in full isotopic equilibrium,additional intra-cave mechanisms may have a distinct influence ond18O and d13C calcite values. A first way to investigate the equilib-rium conditions of speleothem calcite can be done by a Hendy-testwhere several samples are drilled along a single growth layer acrossthe stalagmite. If in one layer (1) a simultaneous enrichment in d18Oand d13C occurs away from the growth axis and (2) a good

Table 2bU/Th measurements for stalagmites STM1, STM5 and STM6 (University of Minnesota). All ages were converted to before 2011.

Samplenumber

Depth(mm)

238U (ppb) 232Th (ppt) 230/Th232Th(atomic ' 10%6)

d234Ua

(measured)

230/Th238U(activity)

230Th age (y)(uncorrected)

230Th age (y)b

(corrected)d234UInitial

c

(corrected)

230Th age(ka B2011)d

(corrected)

STM1 Hoq CaveSTM1-1 6 213.6 # 0.4 42 # 22 18 # 30 99.8 # 2.9 0.00041 # 0.00036 41 # 36 36 # 36 99.8 # 2.9 0.044 # 0.036STM1-2 50 161.4 # 0.3 86 # 24 140 # 44 92.6 # 2.9 0.00451 # 0.00059 450 # 60 440 # 60 92.7 # 2.9 0.448 # 0.060STM1-3 80 289.3 # 0.6 322 # 24 152 # 13 94.1 # 2.7 0.01025 # 0.00047 1027 # 47 998 # 49 94.4 # 2.7 1.006 # 0.049STM1-4 166 202.7 # 0.4 35 # 13 1590 # 570 100.2 # 2.1 0.01678 # 0.00054 1678 # 54 1673 # 54 100.7 # 2.1 1.681 # 0.054STM1-5 201 307.7 # 0.5 204 # 13 454 # 31 98.8 # 1.8 0.01821 # 0.00037 1824 # 38 1807 # 39 99.3 # 1.8 1.815 # 0.039STM1-6 219 192.0 # 0.4 38 # 17 1816 # 800 91.4 # 2.8 0.02164 # 0.00068 2186 # 70 2180 # 70 92.0 # 2.9 2.188 # 0.070STM1-7 302 185.3 # 0.3 58 # 13 1632 # 378 93.9 # 2.1 0.03074 # 0.00058 3112 # 60 3104 # 60 94.7 # 2.1 3.112 # 0.060STM1-8 352 197.1 # 0.3 33 # 14 3877 # 1590 98.2 # 2.2 0.03946 # 0.00070 3993 # 70 3990 # 72 99.3 # 2.2 3.998 # 0.072STM1-9 386 239.8 # 0.4 119 # 20 1420 # 240 106.6 # 2.6 0.04259 # 0.00081 4283 # 83 4270 # 84 107.9 # 2.6 4.278 # 0.084STM1-10 422 227.0 # 0.4 132 # 16 1342 # 168 100.5 # 2.5 0.04723 # 0.00080 4787 # 83 4772 # 84 101.8 # 2.5 4.780 # 0.084STM1-11 428 197.9 # 0.3 806 # 16 210 # 6 101.1 # 2.2 0.05189 # 0.00091 5268 # 95 5160 # 109 102.5 # 2.2 5.168 # 0.109STM1-12 487 180.3 # 0.4 197 # 21 816 # 88 91.3 # 3.1 0.0540 # 0.0011 5540 # 120 5510 # 120 92.7 # 3.1 5.518 # 0.120STM1-13 548 148.8 # 0.3 253 # 19 558 # 45 94.9 # 3.0 0.0568 # 0.0014 5816 # 150 5770 # 150 96.4 # 3.1 5.778 # 0.150STM6 Hoq CaveSTM6-1 17 122.6 # 0.1 46 # 1 376 # 19 105.6 # 1.9 0.0086 # 0.0004 848 # 38 838 # 38 106 # 2 0.838 # 0.038STM6-2 42 275.8 # 0.3 29 # 1 2140 # 74 100.1 # 1.7 0.0138 # 0.0002 1372 # 18 1369 # 18 100 # 2 1.369 # 0.018STM6-3 50 201.2 # 0.4 324 # 13 184 # 8 97.4 # 2.3 0.0180 # 0.0003 1803 # 28 1761 # 36 97.9 # 2.3 1.766 # 0.036STM6-4 82 111.3 # 0.1 46 # 1 1082 # 31 94.4 # 1.6 0.0272 # 0.0004 2747 # 41 2736 # 41 95 # 12 2.736 # 0.041STM6-5 100 225.4 # 0.4 249 # 12 527 # 26 93.7 # 2.2 0.03528 # 0.00036 3579 # 38 3549 # 40 94.6 # 2.2 3.554 # 0.038STM6-6 166 185.4 # 0.5 506 # 17 281 # 10 94.2 # 3.0 0.0465 # 0.0005 4737 # 53 4664 # 64 95.4 # 3.0 4.669 # 0.053STM5 Casecas CaveSTM5-1 12 790 # 2 7237 # 30 5.5 # 0.3 173.7 # 1.6 0.00315 # 0.00018 293 # 16 66 # 115 173.7 # 1.6 0.073 # 0.115STM5-2 15 741 # 1 884 # 18 33 # 2 231.7 # 2.2 0.0024 # 0.0001 212 # 9 184 # 22 232 # 2 0.184 # 0.022STM5-3 31 871 # 2 7333 # 147 10 # 1 203.2 # 2.4 0.0053 # 0.0001 484 # 10 280 # 145 203 # 2 0.280 # 0.145STM5-4 52 689 # 1 3073 # 35 23.4 # 1.1 165.4 # 1.9 0.00631 # 0.00029 593 # 28 481 # 62 165.6 # 1.9 0.488 # 0.062STM5-5 64 595 # 1 579 # 12 104 # 3 185.9 # 2.4 0.0061 # 0.0001 566 # 9 542 # 19 186 # 2 0.542 # 0.019STM5-6 88 617 # 1 4490 # 28 24.6 # 0.7 70.7 # 1.6 0.01083 # 0.00032 1110 # 33 910 # 110 70.8 # 1.6 0.917 # 0.110

230Th dating results. The error is 2s. l230 & 9.1577 ' 10%6 y%1, l234 & 2.8263 ' 10%6 y%1, l238 & 1.55125 ' 10%10 y%1.Corrected 230Th ages assume the initial 230Th/232Th atomic ratio of 14.4 # 2.2 ' 10%6.Those are the values for a material at secular equilibrium, with the crustal 232Th/238U value of 3.8. The errors are arbitrarily assumed to be 50%.

a d234U & [234U/238U]activity ' 1000.b Ages relative to AD 2003 for STM1, AD 2006 for STM6 and AD 2004 for STM5. Samples STM6-1, STM6-2, STM6-4, STM5-2, STM5-3 and STM5-5 are relative to AD 2011.c d234Uinitial was calculated based on 230Th age (T), i.e., d234Uinitial & d234Umeasured ' el234'T.d Ages before AD 2011.

M. Van Rampelbergh et al. / Quaternary Science Reviews 65 (2013) 129e142 135

Maite Van Rampelbergh
75
Maite Van Rampelbergh
Chapter 3: Spelothems from Socotra
Maite Van Rampelbergh
Page 86: Maïté Van Rampelbergh

correlation between both stable isotopic signals can be established,calcite precipitation was affected by kinetic fractionation (Hendy,1971). One Hendy-test was carried out on STM1 and three onSTM5 (Fig. 6). The Hendy test of STM1 displays constant d18O andd13C signals within 20 mm of the stalagmites apex and an increasefurther away. The three Hendy tests in STM5 are very different fromeach otherwhich is partially due to the difficulty of samplingwithinthe very thin layers. STM5(75mm) fails theHendy test by displayingincreasing d18O and d13C values away from the apex. For the Hendytests STM5(9 mm) and STM5(90 mm), d18O and d13C values remainconstant within 10mmof the stalagmites apex. Only the d18O signalin STM5 (90 mm) on the right side of the apex displays decreasingvalues before displaying an increase after 30 mm from the apex.Although, STM1 and STM5 theoretically fail the Hendy-test, therather constant d18O and d13C values close to the stalagmites’ apexand the weak correlation suggest fragile equilibrium near the cen-tral axis. However, different studies have pointed out that even ifstalagmites pass the Hendy test, speleothem calcite could still havebeen deposited out of isotopic equilibrium (Dorale and Liu, 2009;Mühlinghaus et al., 2009). Better is to look at the enrichment of d13Cvalues along an individual growth layer as indicator of dis-equilibrium (Mühlinghaus et al., 2009). As discussed in the Hendytest, the d13C values remain rather constant within 20 mm of theapex for STM1 andwithin 10mm of the apex for STM5 and increasetowards the sides of the stalagmite suggesting fragile equilibriumconditions of the deposited calcite. The fragile equilibrium condi-tions andpresence of kinetic effects in the studied stalagmites is alsosuggested by a strong correlation between the d18O and the d13Csignals in each stalagmite, indicating that similar processes, mostlikelybykinetic effects, influencebothproxies. A third test providinginformation on the degree of isotopic equilibrium is to calculate theexpected d18O and d13C values based on the theoretical “equilib-rium” waterecalcite fractionation factors for C and O (Kim andO’Neil, 1997), the present-day cave temperature and the isotopiccompositions of seepage waters and to compare these results withthe measured d18O and d13C values. For Hoq Cave, the large range(2.89&) in d18O values measured for the seepage waters (Fig. 3)hampers any meaningful modelling of this kind. Moreover, freshcalcite deposited on glass slabs also displays large variations inisotopic composition (Table 1). Such large variations do suggest thatsite specific differences such as different groundwater flow-pathsand the water residence time in the epikarst (storage water versusevent water) can partially be responsible for these large variations.

Also strong locally varying intra-cave factors such as degree of hu-midity and in particular ventilation may be responsible for theselarge variations in isotopic composition of the seepage waters andthe present-day calcite within one cave. Also for Casecas Cave, nosignificant modelling of this kind is possible because the seepagewaters were sampled in January only, which is the dry season andwill consequently lead to too heavy calculated d18O values for theexpected speleothem calcite.

Taken the evidences together, stalagmites in Hoq Cave andCasecas Cave were deposited under fragile equilibrium conditionsand the isotopic signals may partially be influenced by kineticeffects.

To summarize, any above described mechanism can/will influ-ence the d18O and d13C signal of the studied stalagmites in the samedirection; higher d18O and d13C values will always occur duringdrier conditions. Similar conclusions have been established inprevious work on speleothems in Yemen and Oman (Burns et al.,2001; Fleitmann et al., 2003a, 2004a, 2007; Shakun et al., 2007).

5.2. High Mg/Ca and Sr/Ca ratios indicate drier conditions

Mg/Ca ratios and to a lesser extend Sr/Ca ratios can be used ashydrological proxies (Fairchild and Treble, 2009). In arid and semi-arid areas, prior calcite precipitation (PCP) is considered to bea main reason for variable Mg/Ca and Sr/Ca ratios in speleothems(Fairchild et al., 2000, 2006). When downward percolating waterencounters a zone with lower pCO2, degassing occurs and calcitecan precipitate. Consequently, Mg and Sr become enriched com-pared to Ca in the residual water. During drier periods PCP isenhanced as aerated zones increase in the aquifer and residencetime of the water becomes longer (Fairchild et al., 2000). The mainevidence for PCP is covarying Mg/Ca and Sr/Ca ratios (McMillanet al., 2005; Johnson et al., 2006). Because of calcite precipitationin the epikarst during PCP, d13C calcite values increase in tandemwith Mg/Ca and Sr/Ca ratios. The strong similarities between theMg/Ca and Sr/Ca profiles in stalagmites STM1 and STM6 suggestthat PCP is a primary control for trace elemental ratios in bothstalagmites (Fig. 5). This assumption is further validated by thesignificant correlation between d13C and Mg/Ca (r & 0.33,p & 5.8112 ' 10%4) ratios in STM1. Furthermore, d18O values alsoshow a significant correlation with Mg/Ca (r & 0.38,p & 6.3180 ' 10%5) and Sr/Ca (r & 0.42, p & 8.6029 ' 10%6) ratios instalagmite STM1. As mentioned before, the low resolution of the

Fig. 6. d18O (black line) and d13C (grey line) values in & VPDB for the Hendy tests carried out on stalagmites STM1 (b) and STM5 (a and d). Speleothem calcite of STM1 is depositednear equilibrium conditions within 20 mm from the stalagmites apex. Due to very thin layers in STM5, sampling was difficult explaining the less successful results. STM5(75 mm)fails the Hendy test while STM5 (9 mm) and STM5(90 mm), displays rather constant values within 10 mm of the stalagmites apex.

M. Van Rampelbergh et al. / Quaternary Science Reviews 65 (2013) 129e142136

Maite Van Rampelbergh
76
Maite Van Rampelbergh
Maite Van Rampelbergh
Chapter 3: Spelothems from Socotra
Maite Van Rampelbergh
Page 87: Maïté Van Rampelbergh

STM6 Mg/Ca and Sr/Ca time series hampers a meaningful correla-tion with the d-signals. Nevertheless lower Mg/Ca and Sr/Ca ratiosappear to correspond roughly to more negative isotope values.Based on our observations, we suggest that Mg/Ca and Sr/Ca ratiosare sensitive hydrological proxies with higher ratios during drierperiods when PCP is enhanced.

5.3. Kinetic effects as amplifier of the rainfall signal

As indicated by the different equilibrium tests and by the strongcorrelation between the trace elements and the stable isotopesignals, calcite deposition of the studied stalagmites is affected bykinetic fractionation. As discussed in Dreybrodt and Scholz (2011)and Dreybrodt (2011), the degree of kinetic isotopic enrichmentof the deposited calcite is not caused by rapid CO2 degassing butmainly depends on drip rates and calcite precipitation rates, whichin turn depends on the calcite supersaturation. Since the isotopicprofiles of stalagmites STM1 and STM6 sampled at the end of thecave differ strongly from the isotopic signals of stalagmite Hq1,sampled near the cave entrance, we expect that different kineticeffects affect calcite deposition. At the end of 2 km-long Hoq Cave,changes in the d18O and d13C signals of the coeval stalagmites STM1and STM6 are interpreted to be controlled by variations in effectivemoisture, with lower isotopic values indicating higher net precip-itation. However, despite their close proximity (w1 m), the isotopicprofiles of stalagmites STM1 and STM6 are not identical. The moststriking difference occurs from 0 to 0.6 ka and from 2.2 ka to 3.8 kawhen the STM6 isotopic profiles shift to positive values differing by1.5& for d18O and 3& for d13C with the STM1 isotopic record(Fig. 7). A similar isotopic variation range is also observed in freshcalcite deposited on glass slabs (Table 1), in this case 2& for d18Oand 4.5& for d13C. This suggests that despite the high humidity andreduced ventilation at the end of the cave, strong differences inisotopic composition occur between the different drip sites. Duringdry periods drip rates will decrease at the fast growing (STM1) and

slow growing (STM6) sites. However, calcite precipitation at theSTM6 site will be more strongly affected by kinetic effects, as itsgrowth rate is considerably lower compared to the one at the STM1site. The effect of drip rate on the degree of kinetic isotopicenrichment of the deposited calcite is confirmed by differentmodelling (Dreybrodt, 2011; Dreybrodt and Scholz, 2011; Deiningeret al., 2012) and laboratory experiments with synthetic carbonates(Polag et al., 2010; Day and Henderson, 2011). Stalagmite STM6 isthusmore sensitive to small reduction in precipitation and effectivemoisture. Because of its faster growth rate, STM1 keeps growingduring these slightly drier periods and calcite precipitation stilloccurs closer to isotopic equilibrium. Deep in the cave, the slowerthe growth rate, the stronger calcite deposition will be sensitive tosmall changes in kinetic effects and the more the rainfall signalsinduced by the amount effect will be amplified.

Compared to the coeval stalagmites STM1 and STM6, the iso-topic signals of Hq1 vary around much more positive values. Also,the isotopic records of Hq1 display no similarities onmillennial andcentennial scale with the records from the coeval stalagmites STM1and STM6 (Fig. 7). These observations suggest very strong isotopicdisequilibrium deposition of the Hq1 calcite that is further con-firmed by its very strong correlated d18O and d13C signals. Com-pared to the coeval stalagmites sampled at the end of Hoq Cave,stalagmite Hq1 was sampled only 200 m from the entrance whererelative humidity is low and ventilation effects are considerablystronger. Thewide Hoq Cave entrance is located on the face of a cliffwith a large opening towards the downhill seaside allowing strongair circulation in the first parts of the cave. The presence of aircirculation in chambers near the entrance of Hoq Cave is validatedby the preferential growth direction of helictitese a small variety ofstalactites that are twisted and contorted with no apparent regardfor gravity. The stronger the air circulation in the chambers near theentrance, the stronger the evaporation effects on the stalagmitesurface and the further out of equilibrium the calcite will precipi-tate. The relationship between enhanced evaporation and dis-equilibrium deposition of the speleothem calcite has quantitativelybeen confirmed by Deininger et al. (2012). They demonstrated thatloss of water on the solution layer due to evaporation increases theCa2! leading to higher precipitation rates and consequently largerkinetic fractionation effects. Higher isotopic values in stalagmiteHq1 therefore most probably reflect stronger ventilation near thecave entrance.

To summarize, kinetic effects affecting the d18O and d13C signalsin the Hoq Cave stalagmites are related to their location in the cave.For stalagmites STM1 and STM6, sampled deep in the cave whereventilation is minimal, the d18O and d13C signals are affected bychanges in precipitation and effective moisture. Furthermore, theslower the stalagmite’s growth rate, the stronger the amplificationof the precipitation signal by the kinetic effects. For stalagmite Hq1,sampled near the entrance, kinetic effects are related to enhancedventilation.

The Hq1 record provides information on the intensity of ven-tilation near the cave entrance. For reconstructing the easternSocotra climate variability, the best records are given by the coevalSTM1 and STM6 stalagmites. Stalagmite STM6 is even more sensi-tive to small climatic variations because kinetic effects amplify itsclimate signal.

5.4. Evolution of the IOM and its northeast subsystem since 7 ka

Combined interpretation of the stalagmites STM1 and STM6provides the best precipitation reconstruction for eastern Socotra.Both stalagmites reflect variations in precipitation brought by thenortheast winds. From 6.0 ka until 3.8 ka, the STM1 and STM6 re-cords suggest a gradual decrease in precipitation brought by the

Fig. 7. Superimposed d18O and d13C values, both in & VPDB, of the Hoq Cave stalag-mites STM1 (black) STM6 (grey) and Hq1 (dotted black) together with their U/Thpoints (large dots). Two drier northeast monsoon periods with lower cave ventilationoccur from 0 to 0.6 ka and from 2.2 until 3.8 ka (indicated in grey), when stalagmitesSTM1 and STM6 break-up and stalagmite Hq1 shifts to more negative values.

M. Van Rampelbergh et al. / Quaternary Science Reviews 65 (2013) 129e142 137

Maite Van Rampelbergh
77
Maite Van Rampelbergh
Chapter 3: Spelothems from Socotra
Maite Van Rampelbergh
Page 88: Maïté Van Rampelbergh

northeast monsoon (Fig. 7). After 3.8 ka, no long-term trend isvisible and precipitation intensities brought by the northeast windsvary around a constant value. As suggested by the more sensitiveSTM6, two drier periods occur between 0 and 0.6 ka and from 2.2 to3.8 ka. Superimposed shorter-term millennial and centennial var-iations are similar in both stalagmites but can differ up to 100 yearsdue to age uncertainties in the agemodels. The exact forcing behindthe variations in precipitation brought by northeast winds remainsunclear. Its characterization would require comparison with othernortheast precipitation records in the region. The first limitationthat hampers such comparison is that records affected by onlywinter monsoon precipitation are difficult to find in the area. Formost of the sites in Monsoonal Asia, the amounts of precipitationbrought by the southwest monsoon are higher compared to thosebrought by the northeast winds. Consequently, the northeast signalis often overwritten by the southwest signal. Furthermore, mostexisting records such as ocean cores or sedimentary records do notprovide the high resolution needed to compare centennial or mil-lennial scale variations.

The only record allowing comparison with our northeast pre-cipitation records is a d18O stalagmite record from Hoti Cave,northern Oman (Fleitmann et al., 2007). This speleothem shows theevolution of rain brought by northeast winds for northern Oman atmillennial and centennial scale covering the period between 6 kauntil 5.2 ka and from 2.5 ka until present (Fig. 8). As for the coevalstalagmites STM1 and STM6, the d18O values in the northern Oman

stalagmite vary around%2.3& and display no significant long-termtrends. The presence of an important hiatus in the Hoti Cave record,makes a long-term comparison with STM1 and STM6 difficult.However, similar millennial scale variations can be found betweenboth regions (see also dotted lines in Fig. 8), confirming thatnorthern Oman and eastern Socotra are affected by similar mon-soon dynamics since 6.2 ka. Indeed, after 6.2 ka the southwardretreating ITCZ shifted south of northern Oman, making winterprecipitation brought by northeast winds the only moisture source(Fleitmann et al., 2007).

Comparison of the records reflecting northeast rain variationswith the Greenland ice core records (Rasmussen et al., 2006;Vinther et al., 2006) is difficult. On a long term, the Greenland icecores records display a gradual decrease that cannot be found in ourrecords (Fig. 8). This suggests that the long-term decrease in highlatitude temperatures doesn’t influence the long-term evolution ofthe northeast monsoon precipitation. Also on a shorter time scale,millennial and centennial variations don’t correlate between ourrecords and the Greenland ice core records. Also no links existswith the Bond events (Bond et al., 1997).

In contrast to the evolution of the northeast monsoon, recordsreflecting variations in southwest-monsoon precipitation showsimilarities with Greenland ice core records. Southern Oman sta-lagmites (Fleitmann et al., 2003a, 2007) display a long termweakening of the monsoon and shorter-term variations that cor-relate with Greenland ice core variations (Fig. 8). Colder Northernhemisphere periods correspond to weaker and consequently driersouthwest monsoon periods. Since our records do not matchthose of the Greenland ice core, we expect no similarities withprecipitation variations of the southwest monsoon. Indeed, com-parison between different southwest records (Sirocko et al., 1993;Cullen et al., 2000; Fleitmann et al., 2003a; Gupta et al., 2003) andthe northeast records from this study display no similarities(Figs. 8 and 10).

A speleothem d18O record fromwestern Socotra is interpreted toreflect variations in the southwest summer monsoon since 4.4 ka(Fleitmann et al., 2007). Comparison with our records from theeastern side of the island, show that both sides of Socotra havea different long-term evolution (Fig. 9a). Whereas western Socotragradually evolves towards wetter conditions (Fleitmann et al.,2007), eastern Socotra (this study) has a stable long-term precipi-tation trend since 3.8 ka. This emphasizes the important role of theHaggeher Mountains as a watershed creating two different pre-cipitation areas on the island. The presence of two precipitationareas on Socotra is confirmed by the intermediate d18O values ofstalagmite STM5 from Casecas Cave (Fig. 9a). The latter is located atthe fringe between the northeast rain area in the east and themixed southwest northeast rain area in the west (Fig. 1).

Despite the different long-term evolution of both sides ofSocotra Island, the shorter-term (millennial and centennial scale)variations between eastern and western Socotra display similarchanges. Due to age uncertainties, 200-year offsets occur betweenthe variations of western and eastern Socotra, causing an unsuc-cessful statistical correlation between the records. Such an offset isclearly visible for the positive peak around 2.6 ka where the STM6record lags approximately 200 years (thus within age un-certainties) behind the peak in the western Socotran record(Fig. 9a). The observation that shorter-termvariations are similar onboth sides of the island can be explained in two ways. The south-west rains also affects the eastern part of Socotra or the northeastrains also affect western Socotra. Nowadays, western Socotra re-ceives equal amounts of rain during the southwest and the north-east rainy season. For eastern Socotra, the rainfall amounts broughtby the northeast rains are three times the precipitation amountsbrought by the southwest rains (Scholte and De Geest, 2010).

Fig. 8. After 6 ka, the northern Oman record (Hoti Cave (3), Fleitmann et al., 2007)displays similar variations as the eastern Socotran records (STM1 and STM6 (2), thisstudy). Values vary around a similar average of %2.3& (dotted line) with similarmillennial scale variations. No similarities with the Greenland ice core record ((1),Rasmussen et al., 2006; Vinther et al., 2006) indicate that the northeast rains innorthern Oman and eastern Socotra are not sensitive to Northern Hemisphere tem-perature variations. No similarities can be seen between the northeast monsoon sig-nals from this study and a southwest monsoon precipitation signal ((4), Qunf Cave,Fleitmann et al., 2007) suggesting that both monsoons have different mechanisms.Values are reporten in & SMOW for (1) and in & VPDB for (2, 3 and 4).

M. Van Rampelbergh et al. / Quaternary Science Reviews 65 (2013) 129e142138

Maite Van Rampelbergh
78
Maite Van Rampelbergh
Chapter 3: Spelothems from Socotra
Maite Van Rampelbergh
Page 89: Maïté Van Rampelbergh

Therefore, we conclude that the most plausible hypothesisexplaining the similar millennial and centennial scale variationsbetween east and west Socotra is that the northeast rains reach thewestern side of Socotra. Consequently, the millennial and cen-tennial scale variations in the d18O record fromwestern Socotra arecreated by variations in the northeast rainy season.

The long-term trend of the western Socotra record displays anevolution towards wetter conditions that cannot be seen in easternSocotra (Fig. 9a). Since the northeast monsoon does not display anincrease in precipitation after 3.8 ka, the long-term trend onwestern Socotra is most likely linked to increased southwestmonsoonal rainfall. This is apparently surprising considering thatfor Monsoonal Asia, the southwest monsoon is weakening from thebeginning of the Holocene (Sirocko et al., 1993; Neff et al., 2001;Fleitmann et al., 2003a, 2004a, 2007; Gupta et al., 2003;Wang et al.,2005). Modern wind direction measurements carried out on thewestern side of Socotra indicate that when the southwest monsoonreaches its maximal intensity in July, the southwest winds gradu-ally change direction from the south to southwest (Culek et al.,2006). The southwest winds are first passing over dry Somaliabefore hitting Socotra. This means that during the summer mon-soon, thus when the ITCZ reaches its northernmost position, airreaching Socotra is drier than during other periods when air isdirectly coming from over the Indian Ocean. The more northern theITCZ, the more the southwest winds are forced into awesterly path,and the drier the southwest rainy season on Socotra. Based on theseobservations, the following hypothesis for the increasing wetteningof western Socotra can be established. After 8 ka, the mean lat-itudinal position of the summer ITCZmoved southward in response

to the decreasing boreal summer insolation (Fleitmann et al.,2003a). As a consequence, the southwest winds that were origi-nally coming from over dry Somalia are forced into a more south-erly path over the Indian Ocean (Fig. 9b). The resulting winds willtherefore contain more moisture and lead to a wetter summermonsoon and thus an increase of precipitation on the western sideof Socotra. This hypothesis suggests that the increasingly wettersouthwest summer monsoon over Socotra reflects a local effect andis therefore not representative for the whole summer monsoonregion. More research and comparison with currently other highresolution regional late Holocene precipitation reconstructionsfrom around Socotra are necessary to confirm this hypothesis.However, so far such records are lacking.

5.5. End of the Holocene wet period

The similar forcing behind the northern Oman and the easternSocotra records sheds new light on the timing and characteristic ofthe termination of the Holocene wet period in southern Arabia.During the early to middle Holocene (6e10.5 ka), southern Arabiaexperienced clearly wetter conditions than today (Burns et al.,2001). Different proxy records such as from speleothems, lakeand dune deposits and marine sediments from the Arabian Sea donot seem to agree on the timing and characteristics of the end of theHolocene wet period. Roughly, these records can be subdivided intwo groups. The ones that suggest a gradual decrease in precipi-tation since 8 ka (Fig. 10, records 8e10) and those advocating anabrupt decrease around 6 ka (Fig. 10, records 1e7). The first group(Fig. 10, records 8e10) suggesting a gradual decrease since 8 ka

Fig. 9. (a) Whereas western Socotra becomes wetter over the last 4.4 ka (black line, Dimarshim cave, Fleitmann et al., 2007), conditions remain stable on eastern socotra after 3.8 kaas indicated by STM1 (light grey) and STM6 (dark grey) from Hoq Cave (this study). STM5 from Casecas Cave (dotted grey, this study) is influenced by both rainy seasons and hasintermediate d18O values between east and west Socotra. Superimposed millennial variations are similar for all stalagmites but can display offsets of 200 years (within the ageuncertainty ranges). (b) Possible hypothesis: southwest winds hitting western Socotra gradually come from more over the Indian Ocean due to the southward displacement of thesummer ITCZ causing a wetter southwest rainy season.

M. Van Rampelbergh et al. / Quaternary Science Reviews 65 (2013) 129e142 139

Maite Van Rampelbergh
79
Maite Van Rampelbergh
Chapter 3: Spelothems from Socotra
Maite Van Rampelbergh
Page 90: Maïté Van Rampelbergh

includes the speleothems from southern Oman (Fleitmann et al.,2007) and sedimentary cores from the Arabian Sea (Sirocko et al.,1993; Gupta et al., 2003). To the second group (Fig. 10, records 1e7) belong speleothems from northern Oman (Burns et al., 2001;Fleitmann et al., 2007), sediment records from the UAE (Parkeret al., 2006) and the Wahiba sands in northern Oman (Radieset al., 2005), all of them place the termination of Holocene wetperiod at around 6 ka. All locations showing an abrupt end of theHolocene wet period at 6 ka are currently located outside the ITCZmigration pathway and receive their precipitation only once a yearduring the winter from the northeast winds. Around 6 ka the ITCZshifted south of the UAE and northern Oman, disconnecting theseareas from the southwest rains and explaining the clearly markedend of the wet period. Winter and spring precipitation brought by

northeast winds is the only moisture source for that region. Ar-chives still located within the ITCZ migration pathway such assouthern Oman, the Yemenite coast and the northern Indian Oceanshow a rather gradual decrease in precipitation since 8 ka. Theseareas experience a gradual decrease in precipitation due to thegradual southward retreat of the summer ITCZ. In this matter, thetwo apparently opposite views can be reconciled.

Two lake records from the Yemenite lowlands and the Yem-enite Highlands in the west (Lezine et al., 2010) seem to form anexception to this hypothesis by showing an abrupt end of Holo-cene wet period at 7.2 ka (Fig. 10, record 4) and 5 ka (Fig. 10, record5). A more gradual decrease would be expected for this area sincethese lakes receive no rain from northeast winds. These abruptshifts correspond perhaps to a different timing of the threshold

Fig. 10. Compilation of mid Holocene records from the southern Arabian Peninsula and the northern Indian Ocean. Location of all records is shown on the map on top. The Hoq CaveSTM6 record (7) correlates well with the northern Oman speleothem record (6) suggesting that over the last 6.0 ka northern Oman received rain only by northeast winds. Recordscurrently located out of the ITCZ influence (1e4 and 6) display an abrupt end of the mid Holocene wet period around 6.0 ka (indicated by the dotted line) due to their disconnectionof the influence of the southwest monsoonal rains. Records still located within the ITCZ migration pathway (8e10) display a gradual decrease in precipitation due to the southwestmonsoon weakening. Sediment records from northwestern Yemen (5) still received rain after 6.0 ka due to orographic effects explaining the longer wet conditions for that area. TheHolocene boreal summer insolation curve is established using Analyseries (Paillard et al., 1996). 1. Awafi, Parker et al., 2006; 2. Hajar Mountains, Fuchs and Buerkert, 2008; 3. WahibaSands, Radies et al., 2005; 4. Al-Hawa, Lezine et al., 2010; 5. Rada and Saada, Lezine et al., 2010; 6. N-Oman, Fleitmann et al., 2007; 7. E-Socotra, this study; 8. S-Oman, Fleitmannet al., 2007; 9. Arabian Sea, Gupta et al., 2003; 10. Arabian Sea, Sirocko et al., 1993.

M. Van Rampelbergh et al. / Quaternary Science Reviews 65 (2013) 129e142140

Maite Van Rampelbergh
80
Maite Van Rampelbergh
Chapter 3: Spelothems from Socotra
Maite Van Rampelbergh
Page 91: Maïté Van Rampelbergh

point for lake survival. At 7.2 ka BP rain brought by the southwardmigrating ITCZ was too weak to feed the Yemenite lowland lakeslocated in the rain shadow of the Yemenite Highlands, explainingthe abrupt drying. Due to orographic effects, lakes located in theYemen Highlands still received enough precipitation until thethreshold value was reached at 5 ka, when the amount of rainbrought by the southward migrating ITCZ also became insufficientto maintain the lake. A similar situation occurs with a playa-likesediment record from the Hajar Mountains (Northern Oman,Fuchs and Buerkert, 2008) (see also Fig. 10, record 2) that show anabrupt decrease in sedimentation rate at 8 ka while an abruptdecrease at 6 ka should be expected for this region. Althoughprecipitation decreased gradually until the abrupt shift at 6 ka innorthern Oman, the threshold point for these lakes was alreadyreached at 8 ka.

In summary, the end of the Holocene wet period at 8 ka insouthern Arabia is related to the continuous southward retreat ofthe summer ITCZ in response to the decreasing Holocene borealsummer insolation (Fleitmann et al., 2003a). Currently, areas stillreached by the summer ITCZ display a gradual end of the Holocenewet period starting around 8 ka. The areas that are not affectedanymore by the ITCZ and currently receive rain only once a yearonly from northeast winds display an abrupt end at 6.2 ka. Around6 ka, the summer ITCZ was located south of northern Oman andthus winter precipitation delivered by frontal depression systemsfrom the Mediterranean Sea became the dominant source ofmoisture. Climate in the western parts of Socotra evolved graduallytowards wetter conditions since 4.4 ka due to the trajectory of thesouthwest summer monsoon winds passing gradually more overthe Indian Ocean. However, since the currently available westernSocotra records do not cover the mid Holocene period, no robustconclusion can be established about the presence or absence ofa mid Holocene wet period on Socotra.

6. Conclusions

1. Strong differences in the isotopic composition of the seepagewater, modern calcite and between the isotopic profiles of theHoq Cave stalagmites demonstrate that kinetic effects are sitespecific and affect the isotopic composition of contempora-neously deposited stalagmites significantly. A detailed under-standing of the cave dynamics and waterecarbonateinteraction by cave monitoring, combined with a multi-proxyand multi-stalagmite approach is necessary to derive solidclimate reconstruction in semi-arid environments. At the endof Hoq Cave, kinetic effects are caused by variations in growthand drip rate whereas near the entrance, evaporation effectsare responsible for out of equilibrium deposition of the calcite.

2. Understanding the present local climate is necessary to cor-rectly interpret the obtained records. For Socotra, due to thewatershed action of the Haggeher Mountains, records obtainedfrom the southern and western parts reflect the long-termchanges of the southwest rainy season with superimposedsmaller scale variations of the northeast monsoon. Recordsobtained from the eastern and northeast parts documentchanges in northeast rains only.

3. The northeast winter monsoon displays a drying from 6.0 kauntil 3.8 ka and remains stable after 3.8 ka. Two superimposedweaker northeast monsoon periods occur between 0 and 0.6 kaand from 2.2 until 3.8 ka. No correlation can be establishedwith variations in the southwest monsoon and with theNorthern Hemisphere climatic variations. More high resolutionrecords are required to understand the exact forcing behind thenortheast monsoon for this area.

4. The long term wettening of the southwest monsoon on west-ern Socotra since 4.4 ka (Fleitmann et al., 2007) is a local effectthat relates to a changing wind path. A possible hypothesiscould be that in response to the weakening of the southwestmonsoon, the southwest winds are forced into a more south-erly path over the Indian Ocean consequently causing a wettersouthwest monsoon rainy season on western Socotra.

5. After 6.2 ka, similar precipitation signals can be found betweeneastern Socotra and northern Oman suggesting both regionsare similarly affected by the northeast winter monsoon fromthen on. Areas on the Arabian Peninsula such as northernOman currently receiving rain only once a year from northeastwinds display an abrupt end of the Holocene wet optimumaround 6 ka due to their disconnection from the southwestwinds. In contrast, records from the southern Arabian Pen-insula still located within the ITCZ migration pathway andreceiving rain during both monsoon seasons display a gradualdrying after the Holocene wet optimum due to the weakeningof the southwest monsoon after 8 ka.

Acknowledgements

We thank the Yemen Ministry of Water and Environment andthe Environment Protection Authority (EPA e Socotra Branch), themembers of the Socotra Karst Project and of the Friends of Socotragroup for their help during the fieldwork. Maïté Van Rampelberghalso want to thank Mr. R. Van Dierendonck for his interest andsupport, Kay Van Damme for sharing his knowledge on Socotra’svegetation cover and Dirk Van Dorpe for providing additional in-formation on the Hoq Cave dimensions. This work is support by theHercules Foundation to Philippe Claeys, and Research FoundationFlanders (FWO) through project G-0422-10 to Philippe Claeys.

References

Bond, G., Showers, W., Cheseby, M., Lotti, R., Almasi, P., deMenocal, P., Priore, P.,Cullen, H., Hajdas, I., Bonani, G., 1997. A pervasive millennial-scale cycle in NorthAtlantic Holocene and glacial climates. Science 278, 1257e1266.

Burns, S.J., Matter, A., Frank, N., Mangini, A., 1998. Speleothem-based paleoclimaterecord from northern Oman. Geology 26, 499e502.

Burns, S.J., Fleitmann, D., Matter, A., Neff, U., Mangini, A., 2001. Speleothem evidencefrom Oman for continental pluvial events during interglacial periods. Geology29, 623e626.

Burns, S.J., Fleitmann, D., Matter, A., Kramers, J., Al-Subbary, A.A., 2003. IndianOcean climate and an absolute chronology over Dansgaard/Oeschger events 9to 13. Science 301, 1365e1367.

Cheng, H., Edwards, R.L., Hoff, J., Gallup, C.D., Richards, D.A., Asmerom, Y., 2000. Thehalf-lives of uranium-234 and thorium-230. Chemical Geology 169, 17e33.

Cheng, H., Edwards, R.L., Broecker, W.S., Denton, G.H., Kong, X., Wang, Y., Zhang, R.,Wang, X., 2009a. Ice Age terminations. Science 326, 248e252.

Cheng, H., Fleitmann, D., Edwards, R.L., Wang, X.F., Cruz, F.W., Auler, A.S.,Mangini, A., Wang, Y.J., Kong, X.G., Burns, S.J., Matter, A., 2009b. Timing andstructure of the 8.2 kyr BP event inferred from delta O-18 records of stalagmitesfrom China, Oman, and Brazil. Geology 37, 1007e1010.

Cheung, C., De Vantier, L., Van Damme, K., 2006. Socotra, A Natural History of theIslands and Their People. Odyssey Books and Guides, Airphoto InternationalLtd., Hong Kong, pp. 408.

Cook, E.R., Anchukaitis, K.J., Buckley, B.M., D’Arrigo, R.D., Jacoby, G.C., Wright, W.E.,2010. Asian monsoon failure and megadrought during the last millennium.Science, 486e489.

Culek, M., Kral, K., Habrova, H., Aldolt, R., Pavlis, J., Madera, P., 2006. Socotra’s annualweather pattern. In: Cheung, C., De Vantier, L., Van Damme, K. (Eds.), Socotra, ANatural History of the Islands and Their People. Odyssey Books and Guides,Airphoto International Ltd., Hong Kong, pp. 42e46.

Cullen, H.M., deMenocal, P.B., Hemming, S., Hemming, G., Brown, F.H.,Guilderson, T., Sirocko, F., 2000. Climate change and the collapse of the Akka-dian empire: evidence from the deep sea. Geology 28, 379e382.

Dansgaard, W., 1964. Stable isotopes in precipitation. Tellus 16, 436e468.Day, C.C., Henderson, G.M., 2011. Oxygen isotopes in calcite grown under cave-

analogue conditions. Geochimica Et Cosmochimica Acta 75, 3956e3972.Deininger, M., Fohlmeister, J., Scholz, D., Mangini, A., 2012. Isotope disequilibrium

effects: the influence of evaporation and ventilation effects on the carbon andoxygen isotope composition of speleothem e a model approach. Geochimica EtCosmochimica Acta 96, 57e79.

M. Van Rampelbergh et al. / Quaternary Science Reviews 65 (2013) 129e142 141

Maite Van Rampelbergh
81
Maite Van Rampelbergh
Chapter 3: Spelothems from Socotra
Maite Van Rampelbergh
Page 92: Maïté Van Rampelbergh

Dorale, J.A., Liu, Z., 2009. Limitations of Hendy test criteria in judging the paleo-climatic suitability of speleothems and the need for replication. Journal of Caveand Karst Studies 71, 73e80.

Dreybrodt, W., Scholz, D., 2011. Climatic dependence of stable carbon and oxygenisotope signals recorded in speleothems: from soil water to speleothem calcite.Geochimica Et Cosmochimica Acta 75, 734e752.

Dreybrodt, W., 2011. Comments on processes contributing to the isotope compo-sition of 13C and 18O in calcite deposited in speleothems. Acta Carsologica 40.

Dykoski, C.A., Edwards, R.L., Cheng, H., Yuan, D.X., Cai, Y.J., Zhang, M.L., Lin, Y.S.,Qing, J.M., An, Z.S., Revenaugh, J., 2005. A high-resolution, absolute-dated Ho-locene and deglacial Asian monsoon record from Dongge Cave, China. Earth andPlanetary Science Letters 233, 71e86.

Edwards, R.L., Chen, J.H., Wasserburg, G.J., 1987. 238U-234U-230Th-232Th systematicsand the precise measurement of time over the past 500 000 years. Earth andPlanetary Science Letters 81, 175e192.

Epstein, S., Mayeda, 1953. Variations of the 18O/16O ratio in natural waters. Geo-chimica Et Cosmochimica Acta 4, 213e224.

Fairchild, I.J., Treble, P.C., 2009. Trace elements in speleothems as recorders ofenvironmental change. Quaternary Science Reviews 28, 449e468.

Fairchild, I.J., Borsato, A., Tooth, A.F., Frisia, S., Hawkesworth, C.J., Huang, Y.M.,McDermott, F., Spiro, B., 2000. Controls on trace element (Sr-Mg) compositionsof carbonate cave waters: implications for speleothem climatic records.Chemical Geology 166, 255e269.

Fairchild, I.J., Smith, C.L., Baker, A., Fuller, L., Spotl, C., Mattey, D., McDermott, F.,Eimp, 2006. Modification and preservation of environmental signals in spe-leothems. Earth-science Reviews 75, 105e153.

Fleitmann, D., Burns, S.J., Mudelsee, M., Neff, U., Kramers, J., Mangini, A., Matter, A.,2003a. Holocene forcing of the Indian monsoon recorded in a stalagmite fromSouthern Oman. Science 300, 1737e1739.

Fleitmann, D., Burns, S.J., Neff, U., Mangini, A., Matter, A., 2003b. Changing moisturesources over the last 330,000 years in Northern Oman from fluid-inclusionevidence in speleothems. Quaternary Research 60, 223e232.

Fleitmann, D., Burns, S.J., Neff, U., Mudelsee, M., Mangini, A., Matter, A., 2004a.Palaeoclimatic interpretation of high-resolution oxygen isotope profiles derivedfrom annually laminated speleothems from Southern Oman. Quaternary Sci-ence Reviews 23, 935e945.

Fleitmann, D., Matter, A., Burns, S.J., Al-Subbary, A., Al-Aowah, M.A., 2004b. Geologyand Quaternary climate history of Socotra. Fauna of Arabia 20, 27e44.

Fleitmann, D., Burns, S.J., Mangini, A., Mudelsee, M., Kramers, J., Villa, I., Neff, U., Al-Subbary, A.A., Buettner, A., Hippler, D., Matter, A., 2007. Holocene ITCZ andIndian monsoon dynamics recorded in stalagmites from Oman and Yemen(Socotra). Quaternary Science Reviews 26, 170e188.

Frumkin, A., Ford, D.C., Schwarcz, H.P., 2000. Paleoclimate and vegetation of the lastglacial cycles in Jerusalem from a speleothem record. Global BiogeochemicalCycles 14, 863e870.

Fuchs,M., Buerkert, A., 2008. A 20 ka sediment record from theHajarMountain rangein N-Oman, and its implication for detecting arid-humid periods on the south-eastern Arabian Peninsula. Earth and Planetary Science Letters 265, 546e558.

Gasse, F., 2000. Hydrological changes in the African tropics since the Last GlacialMaximum. Quaternary Science Reviews 19, 189e211.

Genty, D., Baker, A., Massault, M., Proctor, C., Gilmour, M., Pons-Branchu, E.,Hamelin, B., 2001. Dead carbon in stalagmites: carbonate bedrock paleo-dissolution vs. ageing of soil organic matter. Implications for C-13 variations inspeleothems. Geochimica Et Cosmochimica Acta 65, 3443e3457.

Genty, D., Blamart, D., Ouahdi, R., Gilmour, M., Baker, A., Jouzel, J., Van-Exter, S.,2003. Precise dating of Dansgaard-Oeschger climate oscillations in westernEurope from stalagmite data. Nature 421, 833e837.

Gupta, A.K., Anderson, D.M., Overpeck, J.T., 2003. Abrupt changes in the Asiansouthwest monsoon during the Holocene and their links to the North AtlanticOcean. Nature 421, 354e357.

Henderson, G.M., 2006. Climate e caving in to new chronologies. Science 313, 620e622.

Hendy, C.H., 1971. Isotopic geochemistry of speleothems: 1. Calculations of effectson different modes of formation on isotopic composition of speleothems andtheir applicability as paleoclimatic indicators. Geochimica Et CosmochimicaActa 35, 801.

Ivanochko, T.S., Ganeshram, R.S., Brummer, G.J.A., Ganssen, G., Jung, S.J.A.,Moreton, S.G., Kroon, D., 2005. Variations in tropical convection as an amplifierof global climate change at the millennial scale. Earth and Planetary ScienceLetters 235, 302e314.

Johnson, K.R., Hu, C.Y., Belshaw, N.S., Henderson, G.M., 2006. Seasonal trace-element and stable-isotope variations in a Chinese speleothem: the potentialfor high-resolution paleomonsoon reconstruction. Earth and Planetary ScienceLetters 244, 394e407.

Kim, S.T., O’Neil, J.R., 1997. Equilibrium and nonequilibrium oxygen isotope effects insynthetic carbonates. Geochimica Et Cosmochimica Acta 61, 3461e3475.

Kropelin, S., Verschuren, D., Lezine, A.M., 2008. Response to comment on “Climate-driven ecosystem succession in the Sahara: the past 6000 years”. Science 322.

Lachniet, M.S., 2009. Climatic and environmental controls on speleothem oxygen-isotope values. Quaternary Science Reviews 28, 412e432.

Lezine, A.M., Tiercelin, J.J., Robert, C., Saliege, J.F., Cleuziou, S., Inizan, M.L.,Braemer, F., 2007. Centennial to millennial-scale variability of the Indianmonsoon during the early Holocene from a sediment, pollen and isotope recordfrom the desert of Yemen. Palaeogeography, Palaeoclimatology, Palaeoecology243, 235e249.

Lezine, A.M., Robert, C., Cleuziou, S., Inizan, M.L., Braemer, F., Saliege, J.F.,Sylvestre, F., Tiercelin, J.J., Crassard, R., Mery, S., Charpentier, V., Steimer-Herbet, T., 2010. Climate change and human occupation in the Southern Ara-bian lowlands during the last deglaciation and the Holocene. Global andPlanetary Change 72, 412e428.

McDermott, F., 2004. Palaeo-climate reconstruction from stable isotope variationsin speleothems: a review. Quaternary Science Reviews 23, 901e918.

McMillan, E.A., Fairchild, I.J., Frisia, S., Borsato, A., McDermott, F., 2005. Annual traceelement cycles in calcite-aragonite speleothems: evidence of drought in thewestern Mediterranean 1200-1100 yr BP. Journal of Quaternary Science 20,423e433.

Mühlinghaus, C., Scholz, D., Mangini, A., 2009. Modelling fractionation of stableisotopes in stalagmites. Geochimica Et Cosmochimica Acta, 7275e7289.

Neff, U., Burns, S.J., Mangini, A., Mudelsee, M., Fleitmann, D., Matter, A., 2001. Strongcoherence between solar variability and the monsoon in Oman between 9 and6 kyr ago. Nature 411, 290e293.

Paillard, D., Labeyrie, L., Yiou, P., 1996. Macintosh program performs time-seriesanalysis. Eos, Transactions American Geophysical Union 77, 379.

Parker, A.G., Goudie, A.S., Stokes, S., White, K., Hodson, M.J., Manning, M., Kennet, D.,2006. A record of Holocene climate change from lake geochemical analyses insoutheastern Arabia. Quaternary Research 66, 465e476.

Polag, D., Scholz, D., Muehlinghaus, C., Spoetl, C., Schroeder-Ritzrau, A., Segl, M.,Mangini, A., 2010. Stable isotope fractionation in speleothems: laboratory ex-periments. Chemical Geology 279, 31e39.

Preusser, F., Radies, D., Matter, A., 2002. A 160,000-year record of dune devel-opment and atmospheric circulation in southern Arabia. Science 296, 2018e2020.

Radies, D., Hasiots, S.T., Preusser, F., Neubert, E., Matter, A., 2005. Paleoclimaticsignificance of Early Holocene faunal assemblages in wet interdune deposits ofthe Wahiba Sand Sea, sultanate of Oman. Journal of Arid Environments 62, 109e125.

Rasmussen, S.O., Andersen, K.K., Svensson, A.M., Steffensen, J.P., Vinther, B.M.,Clausen, H.B., Siggaard-Andersen, M.L., Johnsen, S.J., Larsen, L.B., Dahl-Jensen, D., Bigler, M., Rothlisberger, R., Fischer, H., Goto-Azuma, K.,Hansson, M.E., Ruth, U., 2006. A new Greenland ice core chronology for the lastglacial termination. Journal of Geophysical Research-atmospheres 111.

Rozanski, K., Araguasaraguas, L., Gonfiantini, R., 1992. Relationship between long-term trends of 18O isotope composition of precipitation and climate. Science258, 981e985.

Scholte, P., De Geest, P., 2010. The climate of Socotra Island (Yemen): a first-timeassessment of the timing of the monsoon wind reversal and its influence onprecipitation and vegetation patterns. Journal of Arid Environments 74, 1507e1515.

Scholz, D., Hoffmann, D.L., 2011. StalAge e an algorithm designed for construction ofspeleothem age models. Quaternary Geochronology 6, 369e382.

Shakun, J.D., Burns, S.J., Fleitmann, D., Kramers, J., Matter, A., Al-Subary, A., 2007.A high-resolution, absolute-dated deglacial speleothem record of Indian Oceanclimate from Socotra Island, Yemen. Earth and Planetary Science Letters 259,442e456.

Sirocko, F., Sarnthein, M., Erlenkeuser, H., Lange, H., Arnold, M., Duplessy, J.C., 1993.Century-scale events in monsoonal climate over the past 24,000 years. Nature364, 322e324.

Smith, B., Epstein, S., 1970. Two categories of 13C/12C ratios for higher plants. PlantPhysiology 47, 380e384.

Vinther, B.M., Clausen, H.B., Johnsen, S.J., Rasmussen, S.O., Andersen, S.S.,Buchardt, S.L., Dahl-Jensen, D., Seierstad, I.K., Siggaard-Andersen, M.L.,Steffensen, J.P., Svensson, A., Olsen, J., Heinemeier, J., 2006. A synchronizeddating of three Greenland ice cores throughout the Holocene. Journal of Geo-physical Research-part D-atmospheres 111, 11.

Waliser, D.E., Gautier, C., 1993. A satellite-derived climatology of the ITCZ. Journal ofClimate 6, 2162e2174.

Wang, Y.J., Cheng, H., Edwards, R.L., An, Z.S., Wu, J.Y., Shen, C.C., Dorale, J.A., 2001.A high-resolution absolute-dated Late Pleistocene monsoon record from HuluCave, China. Science 294, 2345e2348.

Wang, Y.J., Cheng, H., Edwards, R.L., He, Y.Q., Kong, X.G., An, Z.S., Wu, J.Y., Kelly, M.J.,Dykoski, C.A., Li, X.D., 2005. The Holocene Asian monsoon: links to solarchanges and North Atlantic climate. Science 308, 854e857.

Zhang, P., Cheng, H., Edwards, R.L., Chen, F., Wang, Y., Yang, X., Liu, J., Tan, M.,Wang, X., Liu, J., An, C., Dai, Z., Zhou, J., Zhang, D., Jia, J., Jin, L., Johnson, K.R.,2008. A test of climate, sun, and culture relationships from an 1810-year Chi-nese cave record. Science 322, 940e942.

M. Van Rampelbergh et al. / Quaternary Science Reviews 65 (2013) 129e142142

Maite Van Rampelbergh
82
Maite Van Rampelbergh
Chapter 3: Spelothems from Socotra
Maite Van Rampelbergh
Page 93: Maïté Van Rampelbergh

Chapter  4:  Cave  Monitoring  

  83  

Chapter  4        

Monitoring  of  the  Proserpine  stalagmite      “The  present  is  the  key  to  the  past”  (Charles  Lyell,  1830)    Understanding  how  the  present-­‐day  cave  system  works  may  be  of  large  interest  when  interpreting  speleothem  proxy  variations  in  the  recent  past.  As  discussed  in  Chapter  2,  a  multitude  of   factors  can  affect  speleothem  δ18O  and  δ13C  values,  chemical  composition,  growth  rate,  layering,  layer  thickness  and  calcite  fabric,  to  name   only   the   most   important   (Baker   et   al.,   1998;   Dreybrodt,   1999,   2008;  Mühlinghaus  et  al.,  2009;  Scholz  et  al.,  2009;  Dreybrodt,  2011;  Deininger  et  al.,  2012;   Dreybrodt,   2012).   In   particular,   speleothem   records   from   mid-­‐latitude  temperate   regions   have   shown   to   be   difficult   to   interpret   in   terms   of   a   single  climate   proxy   (Baker   et   al.,   2011).   Cave  monitoring   data   can   reveal   important  links   that   can   be   used   to   understand  how  proxies   record   climate   (Frisia   et   al.,  2000;  Spötl  et  al.,  2005;  Mattey  et  al.,  2008;  Riechelmann  et  al.,  2011).      The  Proserpine   stalagmite  grows   in   the  well-­‐known  and  easily   accessible  Han-­‐sur-­‐Lesse   showcave,   located   the   south   of   Belgium,   and   has   been   suggested   to  reflect  climate  variations  up  to  seasonal  scales  as  indicated  by  its  clear  seasonal  layering   (Verheyden  et   al.,   2006).  To   correctly   interpret   variations   in  δ18O  and  δ13C  values,  layer  thickness  and  calcite  fabric  (see  Chapter  5),  a  cave  monitoring  campaign  was  set  up  to  investigate  the  possible  environmental  factors  affecting  these  proxies.  Between  October  2012  and  January  2014,  the  Proserpine  drip  site  was  monitored  on  a  biweekly  basis.  Different   cave  parameters  were  measured  such  as  cave  air  and  drip  water  temperature,  the  water  discharge  amount,  pCO2  and  δ13CCO2  values  of  cave  air,  rainwater  δ18O  and  δD  values,  drip  water  pH,  δ18O,  δD  and  δ13CDIC  values,  and  the  δ18O  and  δ13C  values  of  freshly  farmed  cave  calcite.  Results   indicate   that   all   cave   parameters   vary   seasonally   between   a   ‘summer  mode’   from   June   to  December   and   a   ‘winter  mode’   from  December   to   June.  Of  major  importance  for  the  interpretation  of  the  proxies  measured  in  Chapter  5  is  that   the   δ18O   and   δ13C   values   of   present-­‐day   calcite   are   deposited   in   isotopic  

Page 94: Maïté Van Rampelbergh

Chapter  4:  Cave  Monitoring    

 84  

equilibrium  with   the   drip  water.   The   δ18O   values   vary   seasonally   and   become  more  negative   in  summer  (June,   July  and  August)  due  to  the   increased  cave  air  and  drip  water  temperature.  The  δ13C  values  vary  seasonally  in  response  to  the  seasonal  changes  in  prior  calcite  precipitation  (PCP).  In  summer,  the  absence  of  water   recharge   increases   PCP,   which   causes   higher   δ13CDIC   values   of   the   drip  water  and  of  the  calcite  deposited  from  the  drip  water.  δ18O  and  δ13C  values  were  measured  at  high  resolution  in  a  small  part  of  the  Proserpine  core  to  investigate  how  they  vary  according  to  the  layering.  Results  showed  that  dark  layers  display  low  δ18O  values  and  high  δ13C  values.  In  the  clearly  visible  yearly  layer  couplets,  darker   calcite   is   interpreted   to   reflect   summer   conditions   and   slower   growth  while  the  whiter,  more  porous  layers  are  formed  in  winter  when  calcite  growth  rate  increases.      REFERNCES    Baker,  A.,  Genty,  D.,  Dreybrodt,  W.,  Barnes,  W.  L.,  Mockler,  N.  J.,  and  Grapes,  J.:  Testing  theoretically  predicted  stalagmite  growth  rate  with  Recent  annually  laminated  samples:  Implications  for  past  stalagmite  deposition,  Geochimica  Et  Cosmochimica  Acta,  62,  393-­‐404,  1998.  

Baker,  A.,  Wilson,  R.,  Fairchild,  I.  J.,  Franke,  J.,  Spoetl,  C.,  Mattey,  D.,  Trouet,  V.,  and  Fuller,  L.:  High  resolution  delta  O-­‐18  and  delta  C-­‐13  records  from  an  annually  laminated  Scottish  stalagmite  and  relationship  with  last  millennium  climate,  Glob.  Planet.  Change,  79,  303-­‐311,  2011.  

Deininger,  M.,  Fohlmeister,  J.,  Scholz,  D.,  and  Mangini,  A.:  Isotope  disequilibrium  effects:  The  influence  of  evaporation  and  ventilation  effects  on  the  carbon  and  oxygen  isotope  composition  of  speleothems  -­‐  A  model  approach,  Geochimica  Et  Cosmochimica  Acta,  96,  57-­‐79,  2012.  

Dreybrodt,  W.:  Chemical  kinetics,  speleothem  growth  and  climate,  Boreas,  28,  347-­‐356,  1999.  

Dreybrodt,  W.:  Comment  on  "Oxygen  isotopes  in  calcite  grown  under  cave-­‐analogue  conditions"  by  C.C.  Day  and  G.M.  Henderson,  Geochimica  Et  Cosmochimica  Acta,  85,  383-­‐387,  2012.  

Dreybrodt,  W.:  Comments  on  processes  contributing  to  the  isotope  composition  of  13C  and  18O  in  calcite  depostied  in  speleothems,  Acta  Carsologica,  40,  2011.  

Dreybrodt,  W.:  Evolution  of  the  isotopic  composition  of  carbon  in  a  calcite  precipitating  H2O-­‐CO2-­‐CaCO3  solution  and  the  related  isotopic  composition  of  calcite  in  stalagmites,  Geochimica  Et  Cosmochimica  Acta,  72,  4712-­‐4724,  2008.  

Frisia,  S.,  Borsato,  A.,  Fairchild,  I.  J.,  and  McDermott,  F.:  Calcite  fabrics,  growth  mechanisms,  and  environments  of  formation  in  speleothems  from  the  Italian  Alps  and  southwestern  Ireland,  Journal  of  Sedimentary  Research,  70,  1183-­‐1196,  2000.  

Page 95: Maïté Van Rampelbergh

Chapter  4:  Cave  Monitoring  

  85  

Mattey,  D.,  Lowry,  D.,  Duffet,  J.,  Fisher,  R.,  Hodge,  E.,  and  Frisia,  S.:  A  53  year  seasonally  resolved  oxygen  and  carbon  isotope  record  from  a  modem  Gibraltar  speleothem:  Reconstructed  drip  water  and  relationship  to  local  precipitation,  Earth  and  Planetary  Science  Letters,  269,  80-­‐95,  2008.  

Mühlinghaus,  C.,  Scholz,  D.,  and  Mangini,  A.:  Modelling  fractionation  of  stable  isotopes  in  stalagmites,  Geochimica  Et  Cosmochimica  Acta,  doi:  10.1016/j.gca.2009.09.010,  2009.  7275-­‐7289,  2009.  

Riechelmann,  D.  F.  C.,  Schroeder-­‐Ritzrau,  A.,  Scholz,  D.,  Fohlmeister,  J.,  Spoetl,  C.,  Richter,  D.  K.,  and  Mangini,  A.:  Monitoring  Bunker  Cave  (NW  Germany):  A  prerequisite  to  interpret  geochemical  proxy  data  of  speleothems  from  this  site,  Journal  of  Hydrology,  409,  682-­‐695,  2011.  

Scholz,  D.,  Muehlinghaus,  C.,  and  Mangini,  A.:  Modelling  delta  C-­‐13  and  delta  O-­‐18  in  the  solution  layer  on  stalagmite  surfaces,  Geochimica  Et  Cosmochimica  Acta,  73,  2592-­‐2602,  2009.  

Spötl,  C.,  Fairchild,  I.  J.,  and  Tooth,  A.  F.:  Cave  air  control  on  dripwater  geochemistry,  Obir  Caves  (Austria):  Implications  for  speleothem  deposition  in  dynamically  ventilated  caves,  Geochimica  Et  Cosmochimica  Acta,  69,  2451-­‐2468,  2005.  

Verheyden,  S.,  Baele,  J.-­‐M.,  Keppens,  E.,  Genty,  D.,  Cattani,  O.,  Hai,  C.,  Edwards,  L.,  Hucai,  Z.,  Van  Strijdonck,  M.,  and  Quinif,  Y.:  The  proserpine  stalagmite  (Han-­‐sur-­‐Lesse  cave,  Belgium):  Preliminary  environmental  interpretation  of  the  last  1000  years  as  recorded  in  a  layered  speleothem,  Geologica  Belgica,  9,  245-­‐256,  2006.                                                

Page 96: Maïté Van Rampelbergh

Chapter  4:  Cave  Monitoring    

 86  

 

Page 97: Maïté Van Rampelbergh

Clim. Past, 10, 1–15, 2014www.clim-past.net/10/1/2014/doi:10.5194/cp-10-1-2014© Author(s) 2014. CC Attribution 3.0 License.

Monitoring of a fast-growing speleothem site from theHan-sur-Lesse cave, Belgium, indicates equilibriumdeposition of the seasonal �18O and �13C signals in the calcite

M. Van Rampelbergh1, S. Verheyden1,2, M Allan3, Y. Quinif4, E. Keppens1, and P. Claeys1

1Earth System Sciences, Vrije Universiteit Brussel (VUB), Pleinlaan, 1050, Brussels, Belgium2Royal Belgian Institute of Natural Sciences, Geological Survey, Direction Earth and History of Life, Jennerstraat 13,1000, Brussels, Belgium3AGEs, Départment de Géologie, Université de Liège, Allée du 6 Août, B18 Sart-Tilman, 4000, Liège, Belgium4Faculté Polytechnique, Université de Mons, Rue de Houdain 9, 7000, Mons, Belgium

Correspondence to: M. Van Rampelbergh ([email protected])

Received: 21 March 2014 – Published in Clim. Past Discuss.: 22 April 2014Revised: 15 September 2014 – Accepted: 16 September 2014 – Published:

Abstract. Speleothems provide paleoclimate information onmultimillennial to decadal scales in the Holocene. However,seasonal or even monthly resolved records remain scarce.Such records require fast-growing stalagmites and a goodunderstanding of the proxy system on very short timescales.The Proserpine stalagmite from the Han-sur-Less cave (Bel-gium) displays well-defined/clearly visible darker and lighterseasonal layers of 0.5 to 2 mm thickness per single layer,which allows a measuring resolution at a monthly scale.Through a regular cave monitoring, we acquired a good un-derstanding of how �18O and �13C signals in modern calcitereflect climate variations on the seasonal scale. From Decem-ber to June, outside temperatures are cold, inducing low caveair and water temperature, and bio-productivity in the soil islimited, leading to lower pCO2 and higher �13C values of theCO2 in the cave air. From June to December, the measuredfactors display an opposite behavior.

The absence of epikarst water recharge between May andOctober increases prior calcite precipitation (PCP) in the va-dose zone, causing drip water to display increasing pH and�13C values over the summer months. Water recharge of theepikarst in winter diminishes the effect of PCP and as a resultthe pH and �13C of the drip water gradually decrease. The�18O and �13C signals of fresh calcite precipitated on glassslabs also vary seasonally and are both reflecting equilibriumconditions. Lowest �18O values occur during the summer,when the �13C values are high. The �18O values of the cal-

cite display seasonal variations due to changes in the cave airand water temperature. The �13C values reflect the seasonalvariation of the �13CDIC of the drip water, which is affectedby the intensity of PCP. This same anticorrelation of the�18O versus the �13C signals is seen in the monthly resolvedspeleothem record that covers the period between 1976 and1985 AD. Dark layers display lower �18O and higher �13Cvalues. The cave system varies seasonally in response to theactivity of the vegetation cover and outside air temperaturebetween a “summer mode” lasting from June to Decemberand a “winter mode” from December to June. The low �18Oand high �13C values of the darker speleothem layers indi-cate that they are formed during summer, while light layersare formed during winter. The darker the color of a layer,the more compact its calcite structure is, and the more nega-tive its �18O signal and the more positive its �13C signal are.Darker layers deposited from summer drip water affected byPCP are suggested to contain lower Ca2+ concentration. Ifindeed the calcite saturation represents the main factor driv-ing the Proserpine growth rate, the dark layers should growslower than the white layers.

Published by Copernicus Publications on behalf of the European Geosciences Union.

Maite Van Rampelbergh
Chapter 4: Cave Monitoring
Maite Van Rampelbergh
87
Maite Van Rampelbergh
Maite Van Rampelbergh
Page 98: Maïté Van Rampelbergh

2 M. Van Rampelbergh: Monitoring of a fast-growing speleothem site

1 Introduction

In the past 25 years, speleothem records have provided im-portant information on past climate variations on multimil-lennial to decadal scales (e.g., Genty et al., 2003; McDer-mott, 2005; Verheyden et al., 2008b; Wang et al., 2008; VanRampelbergh et al., 2013). With the increasing number ofstudies on cave calcite deposition dynamics (e.g., Dreybrodt,1999, 2008; Verheyden et al., 2008a; Lachniet, 2009; Scholzet al., 2009; Ruan and Hu, 2010; Oster et al., 2012) and withthe help of modern analytical tools (Fairchild et al., 2006;Spötl and Mattey, 2006; Jochum et al., 2012), progress hasbeen made to measure resolution at sub-seasonal and evenbi-monthly scales. However, only a few studies, so far, havereached such high temporal resolution for the measured prox-ies (Treble et al., 2003; Mattey et al., 2008) mainly due tofollowing two limitations.

The first limitation to study the paleoclimate at the sea-sonal scale from stalagmites is that their growth rate needsto be significantly high (around 1 mm yr�1) to deposit thicklayers allowing monthly resolved time series. Speleothemgrowth rates vary according to different factors, such as dripwater rate and calcium ion concentration (Baker et al., 1998;Dreybrodt, 1999), rendering the estimation of an average ratedifficult. Generally, stalagmites increase at 10–100 µm yr�1

in cool temperate climates and at 300–500 µm yr�1 in sub-tropical climates (Fairchild et al., 2006), clearly showing thatfast-growing (more than 1 mm yr�1) speleothems are trulyexceptional.

A second limitation is that a good understanding of thecave system is needed to understand what the measured prox-ies are reflecting. The interaction between the climate param-eters, the soil, the host rock and the cave environment needsto be well understood for the studied cave and time frame.On classical multimillennial and centennial timescales, theprocesses influencing the stable isotopes of oxygen and car-bon are well established (e.g., Fairchild et al., 2006; Bakeret al., 2007). However, local cave-specific effects affect sea-sonally or even monthly resolved �18O and �13C signals. Forstudies at seasonal scales, a detailed study of the cave dy-namics is required (Mattey et al., 2008) in order to under-stand which factors drive the isotopic signals, and at whichintensity. In the last few years, different authors have tried tomodel how �18O and �13C signals are affected by tempera-ture, drip rate, amount of CO2 degassing or residence time ofthe water film on the surface of the stalagmite (Mühlinghauset al., 2007, 2009; Dreybrodt and Scholz, 2011; Deininger etal., 2012). Results of these studies have provided importantprogress in the understanding of the isotope system in caveenvironments. However, cave-monitoring programs remainof crucial importance to test in real cave environments the hy-potheses derived from the models. Different cave-monitoringstudies have been set up all over the world to better un-derstand these seasonal and sub-seasonal processes (Gentyand Deflandre, 1998; Spötl et al., 2005; Mattey et al., 2008;

Riechelmann et al., 2011). Only few of them have providedanswers on the isotope fractionation processes occurring be-tween the drip water and recent precipitated calcite due tothe complexity of the system and the variety of the differentspecific environments (Verheyden et al., 2008a; Tremaine etal., 2011; Riechelmann et al., 2013).

Previous studies of the Han-sur-Lesse karst system showthat the cave responds seasonally to external climate fac-tors and that it is well suited for high-resolution speleothem-based climate reconstructions (Genty and Quinif, 1996; Ver-heyden et al., 2006, 2008a). In the Han-sur-Lesse cave, thehigh growth rate (up to 2.1 mm yr�1) and clear seasonalbanding of the “Proserpine” stalagmite make it possible to re-construct climate variations at the seasonal scale (Verheydenet al., 2006). In this study, we report results of a cave environ-ment monitoring carried out once every 2 weeks for 1 year(2013) that shows how oxygen and carbon isotope signals ob-tained from the Proserpine banding reflect climate variationsat the seasonal scale. The results are then compared to high-resolution �18O and �13C signals measured on the 10 thickestlayers from the upper 10 cm of the Proserpine, which coverthe period from 1976 to 1985 AD. This approach improvesthe knowledge and accuracy for the use of �18O and �13Csignals in speleothems at the seasonal scale.

2 Study area and hydrological setting

The Han-sur-Lesse cave is located within Givetian lime-stones of the Dinant synclinorium and is the largest andbest-developed karst system in Belgium (Delvaux De Fenffe,1985). The Lesse river formed the cave within a hill calledthe “Massif de Boine” entering the karst system at the “Gouf-fre de Belvaux” and exiting approximately 24 h later throughthe “Trou de Han” (Fig. 1). The cave has been exploited sincethe mid-19th century as a touristic attraction and is character-ized by large chambers and well-developed speleothem for-mations. The cave monitoring and speleothem sampling forthis study is carried out in the “Salle-Du-Dôme” chamber.This 150 m wide and 60 m high chamber formed by collapseand is the largest of the whole cave system. The Proserpinestalagmite is easily reachable following the tourist path intothe cave for approximately 700 m from the cave’s exit at theTrou de Han (Fig. 1). It has a large tabular shape with a rela-tively horizontal slightly undulating surface of about 1.5 m2

and is fed by a continuous high drip water flow that dripson mainly four sites. The epikarst thickness above the cavechamber is estimated to be 40 m (Quinif, 1988). Two pas-sages connect the Salle-Du-Dôme to the neighboring cham-bers, and the Lesse river flows at the bottom of the chamber.

The mean annual precipitation at the nearest meteorolog-ical station of Han-sur-Lesse is 844 mm yr�1, and the meanannual air temperature is 10.3 �C (Royal Meteorological In-stitute Belgium). While the temperature displays a well-marked seasonality with cool summers and mild winters, the

Clim. Past, 10, 1–15, 2014 www.clim-past.net/10/1/2014/

Maite Van Rampelbergh
Chapter 4: Cave Monitoring
Maite Van Rampelbergh
Maite Van Rampelbergh
88
Maite Van Rampelbergh
Page 99: Maïté Van Rampelbergh

M. Van Rampelbergh: Monitoring of a fast-growing speleothem site 3

Figure 1. The Han-sur-Lesse karst system, with the Han-sur-Lesse cave as the northern cave system and the Père Noël cave (Verheyden etal., 2008a) as the southern cave system. The Lesse river enters the cave system at the “Gouffre de Belvaux” and exits 24 h later at the “Troude Han”. The studied speleothem is located in the “Salle-du-Dôme” and grows on a pile of debris. Figure adapted after Quinif (1988).

rainfall is spread all over the entire year. According to theKöppen and Geiger classification, the climate above the Han-sur-Lesse cave is considered as a warm temperate, fully hu-mid climate with cool summers (Kottek et al., 2006). For thestudied period lasting between November 2012 and January2014, air temperature was at its lowest between Decemberand March and highest between July and September. Thecoldest temperature of �4.2 �C was reached on 15 March2013. Such an unusually cold March is an exception and doesnot represent the average weather conditions for the studiedregion. The plant coverage above the cave consists of C3-type vegetation with oaks, beech and hazel trees. The soilis approximately 40 cm tick and consists of silty stony soilwith more than 50 % limestone fragments (Belgian Geolog-ical Survey map). The area above the cave is part of a pro-tected natural reserve, preserved from direct human influencefor more than 50 years.

3 Methods

Between October 2012 and January 2014, the cave was vis-ited every 2 weeks to record environmental parameters. Tomake sure that the measured parameters are closely reflect-ing the natural conditions and to guarantee a consistent mea-surement campaign, the cave parameters were always mea-sured around 09:00 a.m. before the first visitor enters thecave. To investigate the visitors’ possible influence on themeasured cave parameters, a test was carried out by mea-suring these parameters before visitors were allowed into thecave and after the passage of different groups. Cave air andwater temperature were measured with a HANNA HI955501thermometer with a precision of 0.2 �C. Air temperature wastaken directly above the stalagmite. The drip water tempera-ture and pH were determined in a small natural pool (6 cm

wide and 3 cm deep) formed on the stalagmite’s surface,where drip water is continuously falling in. The extremelyshort residence time of the water in this small pool guaran-tees that the temperature and pH suffer minimal alteration.The pH of the drip water was measured with a HANNAHI991300 sensor (precision of 0.01 pH). The concentrationof CO2 in the cave air was obtained using an ACCURO640 000 manual Dräger pump with a standard deviation of 10to 15 %. Three times per visit, pCO2 values were measuredat the same spot, right above the surface of the speleothem,and reported as an average of the three values. The drip waterdischarge (volumetric flow rate, here given in volume (mL)of water reaching the speleothem surface per minute) abovethe Proserpine stalagmite was measured in a graded cylinderafter collecting the drip water during 10 min in an inflatablesoft plastic swimming pool (1.77 m2) which was placed onthe stalagmite’s surface to collect the drips from the four dripsites of the Proserpine. Drip water samples for �18O and �Dmeasurements were collected using a container placed at thesurface of the stalagmite, where the drip fell from the stalac-tite approximately 30 m above the stalagmite. Water sampleswere stored in fully filled glass bottles in a cool and darkenvironment. Rainwater samples for �18O and �D measure-ments were collected in a garden close to the cave using athermos bottle and sampled every 15 days between Novem-ber 2012 and January 2014. To avoid evaporation processes,the rainwater was collected using a funnel with a raised edgeand connected to a tube reaching the bottom of the thermosbottle. The funnel was attached to the bottle through a her-metic cap. Glass bottles were fully filled with the collectedrainwater and stored in a dark fridge until being analyzed.

The �18O and �D composition of the waters were mea-sured using a PICARRO L2130-i Cavity Ring-Down Spec-trometer at the Vrije Universiteit Brussel (VUB). For ev-

www.clim-past.net/10/1/2014/ Clim. Past, 10, 1–15, 2014

Maite Van Rampelbergh
Chapter 4: Cave Monitoring
Maite Van Rampelbergh
Maite Van Rampelbergh
89
Maite Van Rampelbergh
Page 100: Maïté Van Rampelbergh

4 M. Van Rampelbergh: Monitoring of a fast-growing speleothem site

ery sample 1.4 mL of water was used for the measure-ments. Every sample was injected and analyzed 10 consec-utive times. The measured values were then corrected us-ing house standards with a strongly different isotopic com-position. The first house standard, called DO1, has a �18Ovalue of �0.79 ± 0.04 ‰ and a �D value of �26.2 ± 0.2 ‰and was made by collecting the damp of boiling water. Thesecond house standard, called DO2, is Milli-Q water witha �18O composition of �7.38 ± 0.04 ‰ and a �D value of�48.8 ± 0.2 ‰. The third house standard, called DO3, is wa-ter from Antarctic glacier ice that was filtered and has a �18Ocomposition of �14.77 ± 0.04 ‰ and a �D composition of�105.1 ± 0.2 ‰. The most positive (DO1) and the most neg-ative (DO3) house standards are used to obtain a two-pointcalibration line. The DO2 house standard, with intermedi-ate values, serves as a “target” or “control” point. By usingthese three standards, we can correct the measured valuesfor a lateral difference as well as for a stretch that can occurin the measurement range. All three working standards weremade in the lab and calibrated against the international stan-dards VSMOW2 (Vienna Standard Mean Ocean Water 2),GISP (Greenland Ice Sheet Precipitation) and SLAP2 (Stan-dard Light Antarctic Precipitation 2). These three interna-tional standards were used to correct the house standards inthe same way the DO1, DO2 and DO3 standards are usedto correct the measurements. The calibration curve was ob-tained using VSMOW2 and SLAP2, and the GISP standardwas used as target. Every house standard was measured 55times on the PICARRO L2130-i. The collected rain and dripwater samples were analyzed two times in different order.The reported values are the average of the two measurementsand reported in per mill VSMOW. Analytical uncertainties(2� ) equal 0.07 ‰ for the measured �18O values and 0.5 ‰for the measured �D values.

The isotopic composition of the cave air CO2 was mea-sured from samples collected by filling vacuum 2 L glasscontainers. To avoid “human” contamination, these sampleswere taken at the beginning of every cave visit. The CO2 wasextracted from the container using a manual extraction lineat the VUB. The extracted CO2 was then analyzed for itsisotopic composition on a Thermo Delta plus XL mass spec-trometer in dual-inlet mode. The standard deviation of thethree measurements reports the error on the measured �13Cvalue. The 2� values average 0.6 ‰. All values are reportedin per mill VPDB (Vienna Pee Dee Belemnite).

Samples for the analyses of the �13C composition of thedissolved inorganic carbon (DIC) in the water were collectedby filling 12 mL gastight glass tubes all the way to the top toavoid air CO2 contamination. A drop of HgCl2 was immedi-ately added and the bottle hermetically closed and stored ina dark and cool environment until being analyzed. The daybefore the analysis, a headspace was created in the bottle bytaking out 3 mL of water, while bubbling He through the sep-tum. Once the headspace was formed, H3PO4 was added andthe sample shaken overnight to convert all DIC species into

CO2. The CO2 gas was then extracted from the bottle andmeasured for its �13C composition. Samples were duplicatedand measured immediately after sampling and 1 month laterto test whether degassing processes affect the DIC composi-tion. This sampling and storing method was tested against themethod described by Spötl et al. (2005) and delivers similarresults within the analytical uncertainties. The �13C compo-sition of the DIC was measured on a Flash EA 1112 deviceconnected to a Delta V plus mass spectrometer. The injectedCO2 is measured against a house standard that consists ofCO2 gas with a �13C composition of �34.07 ‰. The mea-surement series starts with five house standard injections. Af-ter a series of five samples a new house standard injectionis measured to correct the drift. The house standard is cal-ibrated against two international standards, IAEA-CH6 (su-crose) and IAEA-CH7 (polyethylene). All measurements arereported in per mill VPDB with an analytical uncertainty of0.4 ‰ (2� ).

During every visit, three glass slabs were placed on thesurface of the stalagmite to study current calcite depositionconditions on the surface of the stalagmite. All three slabswere positioned very near to where the speleothem core wasdrilled and collected during the next visit. Each time, all threecollected slabs were always completely covered with calcite.The freshly precipitated calcite was then scraped off from theentire slab. Five aliquots per slab were taken from the col-lected powder and measured for their �18O and �13C compo-sition. The reported value per cave visit is the average of theresults of all 15 aliquots of the three slabs collected duringeach visit.

In January 2011, the Proserpine was sampled by drillinga 1 m long core in the middle of the large stalagmite. Aslab was cut from the middle of the core and polished withAl2O3 powder. Layer counting established the age model ofthe upper 16 mm of the core, knowing that one dark and onelight layer are deposited every year (Verheyden et al., 2006).Layer counting was carried out on high-resolution scans us-ing Adobe Photoshop, by counting on the slab itself and byusing the Merchantek MicroMill microscope. Samples to testthe evolution of the isotopic composition of the individuallayers were milled in nine consecutive layer couplets, wherethe dark and white layers were the largest. Samples weredrilled every 40 µm over a length of 4 mm with a MerchantekMicroMill, giving a temporal resolution of approximately 1sample a month. The glass slab and speleothem calcite pow-ders were measured for their �18O and �13C composition us-ing a Kiel III device coupled on a Thermo Delta plus XL. Allvalues were corrected using the international calcite powderstandard NBS-19 and reported in per mill VPDB. Analyticaluncertainties (2� ) were 0.12 ‰ for �13C and 0.16 ‰ for�18O.

Clim. Past, 10, 1–15, 2014 www.clim-past.net/10/1/2014/

Maite Van Rampelbergh
Chapter 4: Cave Monitoring
Maite Van Rampelbergh
90
Maite Van Rampelbergh
Page 101: Maïté Van Rampelbergh

M. Van Rampelbergh: Monitoring of a fast-growing speleothem site 5

4 Results

The drip water feeding the Proserpine falls from a smalldrapery-shaped stalactite, indicating that only a small partof the dissolved calcite precipitates from the drip water whenhanging on the ceiling of the cave. The drip falls approxi-mately 30 m before reaching the surface of the Proserpine.The Proserpine grows under a “flow” or continuous “rain”that falls on the surface of the stalagmite at four points. Thecore used in this study was sampled in the center of thespeleothem, where water flows on the stalagmite during thewhole year.

Figure 2 presents the results of the cave-monitoring cam-paign from November 2012 to January 2014 (15 months)together with precipitation amounts and air temperatures atthe RMI station of Han-sur-Lesse (Fig. 2a and b, respec-tively) for the same period. No difference in the measuredcave parameters is observed before the start of the visitsand after a large number of visitors entered the chamber.The cave air temperature (Fig. 2c) in the Salle-Du-Dômevaries seasonally over a range of ca. 4 �C, with highest val-ues reaching up to 14.5 �C in August and lowest value of10.5 �C in February. After the warm month of August, tem-perature values decrease gradually until December. From De-cember through the end of May, temperatures remain sta-ble and vary around an average of 11 �C. In July, the tem-perature increases quickly to reach the warmest values inAugust.Drip water temperatures (Fig. 2d) follow a similartrend as the cave air temperature (Fig. 2c) but are on average0.5 �C colder.

The drip flow of the Proserpine was always measured forthe whole stalagmite, thus including the four drip sites, tohave an idea of how much water is dripping on the wholesurface of the speleothem. The water flow above the Proser-pine stalagmite (Fig. 2e) averages 161 mL min�1. It sharplyincreases during early winter (December) to an average valueabove 200 mL min�1to remain high throughout winter andspring until early summer (June), when it decreases again toaround 100 mL min�1. With a discharge of 3 ⇥ 10�3 L s�1

and a coefficient of variation of approximately 2, the dripsite can be characterized as a “percolation stream” accord-ing to the classification of Smart and Friederich (1987).In early June, discharge record shows a short maximum to280 mL min�1, which is most probably related to the heavy-rainfall period at the end of May 2013 (Fig. 2a). Superim-posed on this seasonal cycle, very short events of increaseddrip flow are observed within 24 h following a heavy-rainfallevent.

The drip water pH (Fig. 2f) varies between 8.4 and 7.9,decreasing in spring (sharply in May) and gradually increas-ing back at the end of summer and throughout autumn (fromSeptember through January). The heavy-rainfall period at theend of May 2013 (Fig. 2a) seems to correspond with a pHdecrease below 8.0. The pCO2 values (Fig. 2g) remain rela-tively stable around 500 ppm throughout much of the year ex-

cept for a marked increase in the summer, reaching 1000 ppmin July and August. The �13C signature (Fig. 2h) of thecave air varies around an average of �19.5 ‰, displayingan anticorrelation to that of the pCO2 concentrations. The�13C of the DIC (Fig. 2i) varies between �12.2 ± 0.3 ‰ and�11.0 ± 0.3 ‰, increases between June and November anddecreases between December and May.

Rainwater �18O values (Fig. 3c) average �8.18 ± 0.07 ‰,and the �D values (Fig. 3d) �55.52 ± 0.5 ‰. The �18O and�D signals increase by 3 ‰ and 30 ‰, respectively, duringthe summer months, presumably due to temperature effect.One larger drop (red arrow in Fig. 3c and d) of 9 ‰ for the�18O and of 90 ‰ for the �D signal occurs at the beginningof March.

The drip water �18O and �D values (Fig. 3e and f)weakly vary around an average of �7.65 ± 0.07 ‰ and�50.1 ± 0.5 ‰, respectively. These values appear slightlyhigher compared to the yearly average �18O and �D valuesof the rainwater. The drip water isotopic records of oxygenand hydrogen are well correlated and remain stable through-out the year with the exception of one small but meaningfulnegative excursion in July and August of 0.06 ‰ for �18Oand of 0.5 ‰ for �D (red arrow in Fig. 3e and f). The rangeof these shifts is on the order of the analytical uncertainties(0.07 ‰ for �18O and 0.5 ‰ for �D), but they are recorded byat least four consecutive measurements, suggesting that theyare significant.

The �18O signal of the calcite recovered from glass slabsplaced on top of the stalagmite (Fig. 2l) remains stable at�6.5 ± 0.16 ‰ most of the year but decreases to more neg-ative values of �7.1 ± 0.16 ‰ during summer (JJA) (red ar-row in Fig. 2l). The slabs’ calcite �13C signal (Fig. 2m) re-mains relatively constant at �10 ± 0.12 ‰ except for a bulgefrom August through January, with maximal �13C values of�9.0 ± 0.12 ‰ at the end of October (blue arrow in Fig. 2m).The two isotopic signatures are decoupled, suggesting thatdifferent forcing factors affect these signals.

The individual layers of the Proserpine stalagmite also dis-play an anticorrelation pattern between the oxygen and car-bon isotopic signals. The �18O composition oscillates aroundan average �6.5 ± 0.16 ‰ over a range of 0.9 ‰. The �13Cvaries around an average �8.4 ± 0.12 ‰ and over a range of2.4 ‰. Both oxygen and carbon isotopic signals measured inthe stalagmite correspond to the values measured on the glassslabs. At the end of a dark layer (dotted lines in Fig. 4) �18Ovalues reach their minimum while the �13C values reach theirmaximum, illustrating the anticorrelation pattern between the�18O and �13C values at seasonal level.

www.clim-past.net/10/1/2014/ Clim. Past, 10, 1–15, 2014

Maite Van Rampelbergh
Chapter 4: Cave Monitoring
Maite Van Rampelbergh
Maite Van Rampelbergh
91
Maite Van Rampelbergh
Page 102: Maïté Van Rampelbergh

6 M. Van Rampelbergh: Monitoring of a fast-growing speleothem site

Figure 2. Precipitation intensities (a) and air temperatures (b) aremeasured at the Han-sur-Lesse station of the Royal MeteorologicalInstitute. All measured cave parameters – the cave air and drip wa-ter temperature (c and d), drip flow (e), drip water pH (f), cave airpCO2 (g), �13CCAVE AIR CO2 (h), �13CDIC (i), �18ORAINWATER(j), �18ODRIP WATER (k), and the �18O (l) and �13C (m) of freshlydeposited calcite on glass slabs – were measured during the cave-monitoring campaign from November 2012 through January 2014.Cave-monitoring results show that the cave conditions vary season-ally between a “summer mode” (red shadow) lasting from June toDecember and a “winter mode” (blue shadow) lasting from Decem-ber to June. The water �18O values (j and k) are reported in per millVSMOW. �13CCAVE AIR CO2 (h), �13CDIC (i), and �13C and �18Oof freshly deposited calcite on the glass slabs (l and m) are reportedin per mill VPDB.

Figure 3. Precipitation (a) and air temperatures (b) are measuredat the Han-sur-Lesse station of the Royal Meteorological Institute.The �18O and �D of the rainwater (c and d) and the �18O and �D ofthe drip water (e and f) are all reported in per mill VSMOW. At theend of March a cold temperature peak and prolonged snowfall causethe rainwater �18O to display a sharp drop. This negative spike inthe 18Orain water in March can be found back in the 18Odrip water inAugust, indicating that at least part of the infiltrating water reachesthe cave 5 to 6 months later.

5 Discussion

5.1 Forcing of the rain and drip water �18O and�D variations

Generally, rainwater �18O values at a specific location varydue to temperature changes, variation in the amount of rain-fall, fluctuations in the source of the rainwater or cloud track(Rozanski et al., 1992). The rainwater �18O signal increasesby a few per mill during the summer months, when air tem-perature is higher (Fig. 3a and c). A single larger drop indrip water �18O of about 0.1 ‰ occurs in March 2013 andis indicated by a red arrow in Fig. 3. Although it is only ofthe size of the analytical uncertainty, we consider it mean-ingful because several points support it and it is the onlyexcursion of this magnitude. This larger drop does not cor-respond with a decrease in temperature or with an increasein rainfall amount. A modification in rainwater source couldbe a plausible explanation but is not supported by changesin wind direction during that month based on RMI data. Onthe other hand, in March 2013 an unusually late snow layercovered the area for several weeks (RMI data). The observeddecrease in �18O is then most probably related to the collec-tion of isotopically light snow in the sampling bottle.

The average �18O and �D compositions of the drip wa-ter are slightly higher compared to the average �18O and �D

Clim. Past, 10, 1–15, 2014 www.clim-past.net/10/1/2014/

Maite Van Rampelbergh
Chapter 4: Cave Monitoring
Maite Van Rampelbergh
Maite Van Rampelbergh
92
Maite Van Rampelbergh
Page 103: Maïté Van Rampelbergh

M. Van Rampelbergh: Monitoring of a fast-growing speleothem site 7

compositions of the rainwater. This slight increase in � valuesis possibly due to evaporation of precipitated meteoric waterfrom the surface before percolating into the epikarst, leadingto slightly increased �18O and �D values in the vadose wa-ter and in the drip water. Effective precipitation calculationsusing the Thornthwaite (1948) equation on the Han-sur-LessRMI data set for the years of 2011 and 2012 indicate thatJanuary is the only month where no evapotranspiration af-fects the isotopic composition of the water (Bonniver, 2011).Higher evaporation of a part of the precipitating water dur-ing the summer half year is most probably responsible forthe slightly more positive �18O and �D values in the drip wa-ter compared to the average rainwater �18O and �D values.

The �18O and �D compositions of the drip water dis-play almost no variations throughout the year, indicating thatthe water residence time is sufficiently long to homogenizeits isotopic composition (Fig. 2j). No detailed hydrologi-cal study of the Han-sur-Lesse cave system was carried outduring this study. However, the water residence time in theepikarst is supposed to be on the order of months to years.In general, depending on the epikarst thickness and epikarstflow systems, the transfer time of the drip water from thesurface, via the soil and epikarst, to the cave can vary frommonths to several decades. The epikarst thickness above theHan-sut-Lesse cave (40 m) is similar to that observed in shal-low caves such as Ernesto Cave in Italy or Bunker Cave inGermany with epikarst thicknesses between 15 and 30 m.Drip water residence times in such caves are suggested tovary from months to years (Kluge et al., 2010; Miorandi etal., 2010). Considering similar flow patterns, the Han-sur-Lesse residence time of the water in the epikarst may also beon the order of months to years such as Ernesto and Bunkercaves rather than on the order of decades such as observed indeep caves such as Monte Corchia (�1187 m) (Piccini et al.,2008).

The Proserpine drip water �18O displays a small negativeexcursion of 0.1 ‰ in July and August (Fig. 3e). This smalldrop in the drip water composition is most probably relatedto the strong decrease in isotopic composition of the rainwa-ter in March due to a single brief last snowfall. However,the intense mixing of the percolating vadose water in theepikarst reduces the �18O shift of about 8 ‰ in the meteoricto a hardly detectable one of about 0.1 ‰ in the drip water.The presence of this �18O minimum suggests that possiblya small part of the infiltrating water reaches the cave within5 to 6 months, albeit strongly mixed with the larger epikarstreservoir. A similar functioning was suggested in the PèreNoël cave, which also forms part of the Han-sur-Lesse karstsystem (Verheyden et al., 2008a). Uranine-tracing tests of thePère Noël epikarst showed first tracer occurrence 200 h afterinjection, with long restitution time of the curve going upto > 600 h containing superimposed smaller uranine concen-tration peaks (Bonniver, 2011). They described the epikarstas a large reservoir where a part of the infiltrating waterhas a fast flow to the cave while another part is stored for

longer time spans. A similar situation may be present in theepikarst above the Proserpine. However, a more detailed un-derstanding of the epikarst system with additional tracer testsis needed. So far, the hydrological system above the Proser-pine is interpreted as a piston flow system with the first wa-ters coming through after half a year.

5.2 Seasonal variations of the cave atmosphere anddrip flow

A large number of visitors enter the cave and the Salle-du-Dôme chamber every day. Such a large number of visitorscan induce an artificial increase in temperature, pCO2 levelsor a decrease in humidity (Baker and Genty, 1998). The stud-ied Salle-du-Dôme chamber has a height of 60 m and a widthof 150 m, for a total volume of 124 000 m3, and a river flow-ing at the bottom of the chamber causing good air mixing.Due to the large size and the good ventilation of the cham-ber, the effect of the visitors on the measured parameters isexpected to be negligible. This was confirmed by a series oftests where similar cave parameter values were measured be-fore and after groups visited the chamber. For the Salle-du-Dôme chamber, we consider the measured values to reflectthe natural conditions of the cave atmosphere. No cave en-trances connected to the studied Salle-du-Dôme have beenartificially enlarged.

Based on our observations, the temperatures of cave air,drip water and outside air all follow the same seasonal cycle.However, the air temperature, which varies between 20 �C insummer and 0 �C in winter outside the cave, only varies be-tween 14 �C and 11 �C inside the cave, respectively. Apartfrom the influence of the external temperature, the temper-ature regulation of the Salle-du-Dôme chamber is probablyalso influenced by the Lesse river flowing at the bottom of thechamber. Due to the water flow, a good mixing of the cave airis induced, causing the outside air to enter the cave chambermore easily. In addition, the water of the river transports en-ergy from the outside (gained before it entered the cave) tothe inner cave. This heat can then be released in the cave withmore impact in the warm summer months than in the wintermonths. Since no temperature record of the river is available,the effect of the Lesse river on the cave air temperature can-not be quantified.

Compared to other Belgian caves where values up to15 000 ppmv are measured for cave air CO2 (Verheyden,2001; Ek and Godissart, 2014), the CO2 content in the Salle-du-Dôme chamber, which hardly ever reaches 1000 ppmv, in-dicates that exchanges between cave air and external air mustbe relatively important. The Lesse river and the connectionsof the Salle-du-Dôme with neighboring chambers are mostprobably influencing the cave ventilation. In the Salle-du-Dôme, CO2 values fall close to outside air pCO2, and valuesof 400 to 600 ppmv are measured during much of the year(Fig. 2g). Only during summer, higher pCO2 values up to1000 ppmv are measured in the chamber, which corresponds

www.clim-past.net/10/1/2014/ Clim. Past, 10, 1–15, 2014

Maite Van Rampelbergh
Chapter 4: Cave Monitoring
Maite Van Rampelbergh
Maite Van Rampelbergh
93
Maite Van Rampelbergh
Page 104: Maïté Van Rampelbergh

8 M. Van Rampelbergh: Monitoring of a fast-growing speleothem site

to the often observed summer increase in cave air pCO2(Baker and Genty, 1998; Spötl et al., 2005; Riechelmann etal., 2011). In general, even higher amplitude variations areobserved in most seasonal dynamically ventilated caves, suchas St Michael’s cave in Gibraltar with values varying between500 and 6000 ppmv (Mattey et al., 2008), Obir Cave in Aus-tria with values varying between 400 and 1500 ppmv (Spötlet al., 2005) Crag Cave (Ireland) with values between 1100and 8000 ppmv (Baldini et al., 2008). In the Han-sur-Lessecave, lower pCO2 values with smaller variation ranges areobserved, similar to what is observed in year-round well-ventilated caves such as Bunker Cave (Riechelmann et al.,2011). This confirms the former suggestion that the Han-sur-Lesse cave air is most probably well mixed year-round dueto good ventilation, due to the three known entrances and theLesse river, which continuously flows through it. The dif-ference between seasonally ventilated caves and year-roundventilated caves is also visible in the �13C composition of thecave air (Riechelmann et al., 2011). The �13C composition ofthe cave air in seasonally dynamically ventilated caves variesover a larger range around 10 ‰ (Spötl et al., 2005; Matteyet al., 2008) compared to year-round ventilated caves, suchas Bunker Cave, where a variation range of only 3 ‰ is ob-served (Riechelmann et al., 2011). The Han-sur-Lesse cavedisplays a �13C variation of 4 ‰, corresponding with a levelthat is expected for a year-round well-ventilated cave. Al-though the cave is well ventilated, a more negative �13C sig-nature of the CO2 air down to �21.9 ‰ is measured in theSalle-du-Dôme in spring and summer (Fig. 2h), suggestingenhanced input of C3-plant-derived soil CO2 into the caveatmosphere.

Seasonal variations are also seen in the flow rates abovethe stalagmite, with less water dripping on the stalagmiteduring spring and summer (Fig. 2e). Drip rate monitoringin the Père Noël cave, which is part of the Han-sur-Lessecave system, also demonstrated seasonal drip rate variations,with higher discharge amounts in winter compared to sum-mer (Genty and Deflandre, 1998). This seasonal variationin discharge above the Han-sur-Lesse cave is mostly relatedto seasonal variations in evapotranspiration. This lowers thequantity of water feeding the epikarst during summer, reduc-ing the amount of water that can reach the cave. In winter,the situation reverses: the activity of the vegetation cover di-minishes and air temperatures lowers, reducing evapotran-spiration, all of which allows more water to enter the soiland the epikarst. The pressure on the piston flow system israised and more water is pushed into the cave, increasing thedischarge above the stalagmite. Effective precipitation calcu-lation of the Han-sur-Lesse karst aquifer indicates that fromJune to September the net evaporation or evapotranspirationis larger than the amounts of precipitation. The karst aquiferis thus only recharged during winter and spring: from Oc-tober to May (Bonniver, 2011). This seasonal recharge isclearly visible in the evolution of the drip rate. A gradual

decrease in drip rate occurs from June to November, whiledrip rate clearly increases from December to May (Fig. 2e).

5.3 Seasonal variations in the pH and �13C of thedrip water

The drip water pH and �13CDIC display similar seasonal vari-ations with an increase during summer and a decrease duringwinter. Synchronous seasonal variations in pH and �13CDICdriven by cave air pCO2 variations are commonly observedin caves (Spötl et al., 2005; Mattey et al., 2008). The de-gree of CO2 degassing can influence the pH and �13CDIC ofthe drip water; removal of CO2 from the water during de-gassing increases its pH, and removal of light 12C isotopesfrom the drip water during degassing increases its �13CDIC.The degree of CO2 degassing of the drip water is mainlydriven by the pCO2 difference between the cave air (withlow pCO2) and the drip water (with high pCO2) (Mühling-haus et al., 2009; Dreybrodt and Scholz, 2011; Deininger etal., 2012). This process is most pronounced in caves witha large seasonal variation in cave air pCO2 such as in theseasonally ventilated St Michael’s Cave in Gibraltar (Mat-tey et al., 2008) or Obir Cave in Austria (Spötl et al., 2005).The monitoring data indicate that the similar seasonal varia-tions of the drip water pH and �13CDIC are not a response tothe seasonal variations of cave air pCO2. The Han-sur-Lessecave is a year-round well-ventilated cave with a much smallerpCO2 seasonal variation range, roughly between 400 and1000 ppmv (Fig. 2g). If present in the Han-sur-Lesse cave,the effect of degree of CO2 degassing on the drip water pHand �13CDIC is expected to be rather small. More importantis that, if the degree of CO2 degassing is driving the pH and�13CDIC variations at the Proserpine drip site, they must an-ticorrelate with the pH and �13CDIC of the drip water. Suchanticorrelation is not visible in the results (Fig. 2f, g and i);the pCO2 peaks in summer, while the pH and �13CDIC dis-play a more gradual decrease. Furthermore, modeling resultshave shown that the effect of degree of CO2 degassing on the�13C and pH of the drip water is negligible for fast drip sitessuch as the Proserpine (Mühlinghaus et al., 2009; Dreybrodtand Scholz, 2011; Deininger et al., 2012). The effect of de-gassing due to seasonal pCO2 variations in the cave air isthus most probably not driving the variations observed in thepH and the �13CDIC of the drip water.

Another factor that can seasonally increase both the pHand �13CDIC of the drip water (and of the deposited calcite)is prior calcite precipitation (PCP). PCP is a common pro-cess occurring in karst aquifers (e.g., Fairchild et al., 2000,2006; Verheyden et al., 2008a; Riechelmann et al., 2011).When downward-percolating water encounters a zone withlower pCO2, degassing occurs and calcite can precipitate.This process is enhanced during drier periods and lower dis-charge as aerated zones increase in the epikarst and residencetime of the water becomes longer (Fairchild et al., 2000).The CO2 degassing in the epikarts causes both the �18O and

Clim. Past, 10, 1–15, 2014 www.clim-past.net/10/1/2014/

Maite Van Rampelbergh
Chapter 4: Cave Monitoring
Maite Van Rampelbergh
Maite Van Rampelbergh
94
Page 105: Maïté Van Rampelbergh

M. Van Rampelbergh: Monitoring of a fast-growing speleothem site 9

�13C signals in the remaining water to increase. However, ifthe transition time of the solution until reaching the cave islonger than several days, the oxygen isotope equilibrium ofthe CO2–H2O–CaCO3 system in the water is re-establishedagain (Dreybrodt and Scholz, 2011). The net effect is thus anincrease in �13C and pH of the drip water, while no variationis visible in the �18O of the drip water. Furthermore, effectiveprecipitation data show that no water recharges the Han-sur-Lesse epikarst between June and the end of September (Bon-niver, 2011). This lower epikarst recharge is also visible ina decrease of drip flow between June and October (Fig. 2e).The decreased recharge of the epikarst is suggested to in-crease the aerated zone in the epikarst and thus enhance PCPgradually from June throughout September. The increase ofrecharge in winter decreases the effect of PCP, and the dripwater pH and �13C display a gradual decreasing trend. PCPis thus strongly suggested to cause the seasonal variations inpH and �13C of the Proserpine drip water.

The occurrence of PCP can be indicated by an increaseddrip water �13C composition compared to the soil water �13Cand/or by a negative relationship between the Mg / Ca ratioor Sr / Ca of the drip water and the Ca2+ concentration anddrip rate (Fairchild and Treble, 2009). However, no data onthe evolution of the Ca, Mg or Sr concentration in the Pros-erpine drip water or the �13C composition of the soil waterare available, making it difficult to estimate the importance ofthe process. Another argument in support of the PCP effect isthe formation of a stalactite above the stalagmite. However,during this formation of PCP, and due to the high drip rateof the Proserpine, the transition time is not long enough tore-establish the increased �18O signal in the water and boththe �13C and the �18O values of the drip water are increased(Dreybrodt and Scholz, 2011). Since the �18O and �13C sig-nals of the Proserpine drip water evolve differently, stalac-tite growth can increase both signals but cannot be respon-sible for the seasonal variations. Furthermore, the draperyabove the Proserpine is estimated to be about 0.5 to 1 m longand is thus rather small compared to the 2 to 3 m height ofthe Proserpine stalagmite, which displays very high growthrates up to 2 mm yr�1 (Verheyden et al., 2006). It is thus notcertain whether the stalactite is actively growing in presenttimes. Also, due to the continuous water flow, calcite willmost probably not have the time to precipitate stalactite cal-cite before falling on the stalagmite. Dating and analysis ofthe stalactite above the Proserpine would provide the missinginformation. However, its location 30 m above the Proserpinemakes it difficult to be sampled.

In summary, the cave system is subdivided into a “win-ter mode” lasting from December to June and a “summermode” from June to December (Fig. 2). During the wintermode, cave air and drip water temperature and pCO2 are lowand �13C values of the cave air CO2 are high. The plant cov-erage above the cave inactively facilitates water recharge ofthe epikarst reservoir, leading to a decrease of PCP, whichgradually decreases the pH and the �13CDIC of the drip wa-

ter. During the summer mode, cave air and drip water tem-peratures increase, the pCO2 increases and the �13C valuesof the cave air CO2 decrease. The plant coverage above thecave reactivates, leading to lower water recharge, which en-hances the PCP effect, leading to increased pH and �13CDICvalues in the drip water.

5.4 The �18O of the precipitated calcite reflectstemperature variations

The �18O of calcite deposited on the glass slabs varies sea-sonally with more negative values during summer (Fig. 2l).If the calcite is deposited in isotopic equilibrium with its dripwater, these variations can be caused by changes in the �18Oof the drip water and/or by changes in temperature that af-fect the fractionation factor between the drip water and theprecipitating calcite (Dreybrodt and Scholz, 2011). If not de-posited in equilibrium, the seasonal �18O variations on theglass slabs would be due to disequilibrium effects that re-quire further investigation.

A first step in understanding the �18O system of the Salle-du-Dôme demands determining whether the calcite is de-posited in equilibrium or not. In speleothems, this is tra-ditionally done by applying the Hendy test (Hendy, 1971),which compares the �18O and �13C values in the center ofthe stalagmite with those on the sides within a single growthlayer. However, since we work on a drill core taken “in themiddle” of a 1–2 m wide speleothem, the Hendy test cannotbe applied. As an alternative, we calculated equilibrium con-ditions of the deposited calcite using the calcite–water frac-tionation factor. Different authors have proposed fractiona-tion factors based on three different approaches, summarizedin Table 1.

The first approach, using laboratory experiments, has beentested in different studies, each giving another value for awater–calcite fractionation factor. The most-used laboratory-established fractionation factors remain the ones by O’Neilet al. (1969) later modified by Friedman and O’Neil (1977),the relationship of Kim and O’Neil (1997) later modified byKim et al. (2007), and the results of Tarutani et al. (1969)and of Jimenez-Lopez et al. (2001) (Table 1). A second ap-proach to determine fractionation factors is established by us-ing theoretical models such as the ones from Horita and Clay-ton (2007) and from Chacko and Deines (2008). A third ap-proach consists of using cave-monitoring data to make an av-erage of the in-cave-observed fractionation factors (Demenyet al., 2010; Tremaine et al., 2011). Tremaine et al. (2011)established such a “cave calcite” relationship by doing a bestfit through the data on a large number of modern caves atdifferent latitudes, altitudes and temperatures.

Applying the different fractionation factors to our datashows that the measured �18O signals of the glass-slab cal-cite correspond within 1 ‰ with the calculated values, sug-gesting that it is deposited close to equilibrium with the dripwater (Table 2). However, variations occur between the re-

www.clim-past.net/10/1/2014/ Clim. Past, 10, 1–15, 2014

Maite Van Rampelbergh
Chapter 4: Cave Monitoring
Maite Van Rampelbergh
Maite Van Rampelbergh
95
Maite Van Rampelbergh
Page 106: Maïté Van Rampelbergh

10 M. Van Rampelbergh: Monitoring of a fast-growing speleothem site

Table 1. A selection of the most commonly used water–calcite oxygen fractionation factors. Laboratory and theoretical approaches differfrom the relationships found in cave settings.

Author Method 1000 ⇥ ln↵ d↵/dT (‰ �C�1)

O’Neil et al. (1969) modified Laboratory 2.78(106T �2) � 2.89 �0.24by Friedman and O’Neil (1977)Kim and O’Neil (1997) modified Laboratory 18.03(106T �2) � 31.17 �0.22by Kim et al. (2007)Chacko and Deines (2008) constructed Theoretical calculation 2.57333(106T �2) � 0.869 �0.22relation by Tremaine et al. (2011)Horita and Clayton (2007) Calculations compared with 0.9521(106T �2) + 11.59(103T �1) � 21.56 �0.23

experimental resultsTremaine et al. (2011) Linear best fit through large (16.01 ± 0.65) ⇥ (103T �1) � (24.6 ± 2.2) �0.18

number of cave studiesDemeny et al. (2010) Cave-monitoring results 17500 ⇥ T �1 � 29.89 �0.22

Hungarian cave

sults derived from different methods. Among the laboratory-established relationships, that of O’Neil et al. (1969) mod-ified by Friedmann and O’Neil (1977) best corresponds toour observations with on average 0.25 ‰ difference with thevalues measured on the glass slabs. This is also confirmed inother studies where this experimental fractionation factor isconsidered to give the best approximation for in-cave obser-vations (McDermott et al., 2005; Riechelmann et al., 2013).The theoretical values (Horita and Clayton, 2007; Chackoand Deines, 2008) suggest a more negative �18O compositionfor the deposited calcites. This is also the case for other stud-ies where theoretical results seem to overestimate the frac-tionation factor (Demeny et al., 2010; Tremaine et al., 2011).For our results the best agreement is found by applying theTremaine et al. (2011) fractionation factor, which is not sur-prising since the latter is based on experimental studies oncalcite formed in caves. However, the good match betweenthe measured values and those calculated from the Tremaineet al. (2011) relationship does not constitute a proof of equi-librium condition. The fractionation factor from Tremaine etal. (2011) is derived from the natural system, and it is ques-tionable whether real equilibrium does exist in nature.

The small difference between the measured data and theO’Neil et al. (1969) relationship modified by Friedmann andO’Neil (1977) indicates that the deposition of calcite occursnear oxygen isotopic equilibrium in the Salle-du-Dôme. Sea-sonal variations of �18O observed in calcite are likely causedby variations in the drip water �18O composition and/or in thetemperature-dependent fractionation factor (Fairchild et al.,2006). However, within analytical uncertainty, the �18O com-position of the drip water remains constant throughout theyear. Consequently, variations in the fractionation factor dueto temperature changes in the cave air likely explain the sea-sonal pattern seen in the �18O composition of the glass slab.If the most commonly accepted temperature dependence ofthe water–calcite fractionation factor for the oxygen isotopesof 0.247 ‰ 1 �C�1 (O’Neil et al., 1969) is used, our mea-

sured net difference of the �18O values of the glass-slab cal-cite (i.e., �18O range of the glass-slab calcite minus the �18Orange of the drip water) of 0.58 ‰ would result from a 2.3 �Cvariation in the drip water temperature. This temperature cor-responds well with the 2–2.5 �C temperature range measuredin the drip waters. This correspondence constitutes a strongconfirmation of both the isotopic equilibrium and the temper-ature dependence of the calcite �18O.

To summarize, the �18O composition of the glass-slab cal-cite (Fig. 2l) is deposited very close to equilibrium with itsdrip water. Seasonal variations in �18O composition of thecalcite on the glass slabs are caused by the very similar sea-sonal temperature variation of the cave air and the drip water(Fig. 2c and d). A temperature increase of 1 �C correspondswith a decrease of 0.20 ‰ in isotopic composition of the de-posited calcite. The warmer the cave air, the more negativethe �18O composition of the formed calcite. No disequilib-rium processes are active, as they would shift the isotopiccomposition to heavier values with increasing temperature.

5.5 The calcite �13C reflects equilibrium conditionsand is driven by PCP

The carbon isotopic composition of calcite deposited in equi-librium with its drip water depends on (i) the �13C of the DICin the drip water and (ii) the temperature-dependent fraction-ation factor between the DIC and the deposited calcite fora pH range around 7. The average �13C values of the DICin the Han-sur-Lesse drip water (Fig. 2i) display a season-ality between �12.2 ‰ at the end of winter and �11 ‰ atthe end of summer. The �13C of the deposited glass-slab cal-cite are on average 1 ‰ heavier than the �13C of the dripwater and display similar seasonal variations (r = 0.62 andp = 0.0017). The fractionation factor between the DIC andthe deposited calcite as determined by Emrich et al. (1970)and Dulinski and Rozanki (1990) are estimated to correspondbest with the natural speleothem depositional conditions, and

Clim. Past, 10, 1–15, 2014 www.clim-past.net/10/1/2014/

Maite Van Rampelbergh
Chapter 4: Cave Monitoring
Maite Van Rampelbergh
Maite Van Rampelbergh
96
Maite Van Rampelbergh
Page 107: Maïté Van Rampelbergh

M. Van Rampelbergh: Monitoring of a fast-growing speleothem site 11

Table 2. Calculated �18O values for calcite deposited in equilibrium with its drip water for different authors. Values were calculated usingthe average drip water composition of �7.65 ‰ for the lowest, highest and average measured temperatures at the Salle-du-Dôme. The secondpart of the table shows the difference between the calculated �18O values and the �18O values measured on the glass slabs. Best results aregiven by the “in-cave” relationship of Tremaine et al. (2011).

�18ODripWater = �7.65 ‰ Predicted �18O Difference with glass-slab valuesTmin Tmid Tmax �18Omin �18Omid �18Omax

Author 10.5 �C 12 �C 14.5 �C �6.25 �6.55 �6.98

O’Neil et al. (1969) �6.44 �6.80 �7.39 0.19 0.25 0.41modified by Friedmann and O’Neil (1977)Kim and O’Neil (1997) �5.71 �6.04 �6.59 �0.54 �0.51 �0.39modified by Kim et al. (2007)Horita and Clayton (2007) �6.96 �7.30 �7.85 0.71 0.75 0.87Chacko and Deines (2008) �6.98 �7.32 �7.86 0.73 0.77 0.88Demeny et al. (2010) �6.30 �6.62 �7.15 0.05 0.07 0.17Tremaine et al. (2011) �6.26 ± 0.09 �6.55 ± 0.08 �7.04 ± 0.06 0.01 ± 0.09 0 ± 0.08 0.06 ± 0.06

both suggest " values around 1 ‰ for temperatures between5 and 20 �C. The 1 ‰ enrichment of the �13C calcite on theglass slabs compared to the drip water �13C indicates that the�13C is deposited in equilibrium with the drip water.

Our hypothesis that the deposition of the calcite occurs inisotopic equilibrium is also confirmed by the different evolu-tion of the �18O (Fig. 2l) and the �13C (Fig. 2m) values of theglass-slab calcite, which indicates that both proxies evolveindependently under the influence of different factors. Em-rich et al. (1970) and Dulinski and Rozanki (1990) both sug-gest a small temperature dependence of the fractionation fac-tor of about 0.07 ‰/ �C. The ⇠ 2.5 �C seasonal temperaturevariation in the Salle-Du-Dôme causes a change of hardly0.17 ‰ in the calcite composition. The seasonal temperaturevariations are thus too small to significantly influence the car-bon fractionation factor and thus the isotopic composition ofthe deposited calcite. Furthermore typical disequilibrium ef-fects such as increased drip interval and stronger pCO2 de-gassing of the drip water due to longer residence times havebeen shown not to affect the �13C of calcite deposited un-der fast-flow sites (Dreybrodt and Scholz, 2011; Deininger etal., 2012). The �13C composition of the deposited Proserpinecalcite is thus considered deposited in equilibrium with thedrip water. The seasonal variations in calcite �13C are causedby seasonal variations in the drip water �13C, which relatesprobably to the intensity of PCP.

5.6 Variations in the �18O and �13C of the stalagmitereflect seasonal variations

The Proserpine stalagmite displays clear lamination formedby alternating dark, compact layers and white, more porouslayers (Verheyden et al., 2006). The seasonal character of thelayering in the Proserpine stalagmite, with one dark and onewhite layer deposited every year, was already demonstratedby Verheyden et al. (2006). However, these authors were notable to determine the correspondence between layer type and

season. In our studied core, 10 additional layer couplets (=dark + white layer) can be counted, compared to the Proser-pine core of Verheyden et al. (2006) drilled in 2001. The 10additional layer couplets, counted over a period of 10 years,confirm that one layer couplet is deposited every year. Layercouplets establish the age model of the laminated part of thestalagmite. Twenty-six layer couplets are counted from thetop of the stalagmite to the start of the isotope sampling, in-dicating that the youngest analyzed layer formed in 1985 AD.The isotopic measurements were conducted on nine consec-utive layer couplets and consequently run from 1985 to 1976AD (Fig. 5).

A first conclusion from the isotopic analyses of the indi-vidual layers is that the �13C and �18O signals display anopposite behavior within one layer (Fig. 4). When the �13Csignal reaches its maximum value at the end of a dark layer,the �18O value arrives at its minimum value. This anticorre-lation is also seen in the cave-monitoring results where �13Cvalues increase through summer while the �18O values arelow. This indicates that dark layers are most probably de-posited in summer, while light layers are formed in winter.Furthermore, a link between the variation range of the �13Cand �18O signals and the color intensity of a layer (darkerdark layers or lighter white layers) may be possible. The lay-ers around 1977 AD display a stronger visual color contrastbetween dark and white layers compared to the layers around1985 AD (Fig. 4). In parallel, a larger difference in the iso-topic composition of two consecutive layers is observed forthat year. More compact and darker layers have more neg-ative �18O and more positive �13C values while whiter andmore porous layers have more negative �13C and more posi-tive �18O values.

The difference in calcite between white and dark may pos-sibly be related to seasonal changes in growth rate of the cal-cite. The mean annual growth rate of speleothems primarilydepends on the drip water calcium ion concentration (Gentyet al., 2001). In addition, there is a correlation between the

www.clim-past.net/10/1/2014/ Clim. Past, 10, 1–15, 2014

Maite Van Rampelbergh
Chapter 4: Cave Monitoring
Maite Van Rampelbergh
Maite Van Rampelbergh
97
Page 108: Maïté Van Rampelbergh

12 M. Van Rampelbergh: Monitoring of a fast-growing speleothem site

Figure 4. The �18O and �13C signals, reported in per mill VPDB,from the Proserpine stalagmite anticorrelate at monthly scale. Notethat, in the scanned figure of the Proserpine slab, dark and com-pact layers become whiter due to the translucent light of the scanwhile the white and porous layers become dark. At the end of adark compact layer (dotted line), the �18O values reach their min-imum values, while the �13C values reach their maximum. Darklayers are formed during summer, when cave temperatures are high(leading the low �18O) and PCP increases (leading to high �13C).The clearer the lamination, the larger the amplitude of the variationsin the isotopic composition.

measured growth rate and the temperature due to the inter-relationship between calcium ion concentration, soil pCO2and surface temperature. However, the latter link is not al-ways true since PCP and/or lack of soil cover can decreasethe calcium ion concentration of the water and still repre-sent high temperatures (Genty et al., 2001). Discharge alsoaffects growth rate but is more pronounced for slowly drip-ping sites. For faster-drip sites, such as the Proserpine, cal-cite saturation index is the main factor driving the growthrate changes (Genty et al., 2001). Unfortunately, no dataon the seasonal evolution of the calcite saturation index atthe Proserpine grow site are available. However, seasonalvariations in the drip water �13C values of the Proserpinehave shown that PCP affects its geochemistry in summer.PCP lowers the calcite saturation index (Genty et al., 2001;Riechelmann et al., 2013). The increase of the PCP effect insummer strongly suggests that the Proserpine summer watercontains less dissolved Ca2+, which may be responsible fora decreased growth rate in summer.

The observations gained from combined monitoring ob-servation and stable isotopic analyses answer the remain-ing question in Verheyden et al. (2006). Darker layers re-flect summer cave conditions, and white layers reflect win-ter cave conditions. In addition, in the summer, the proba-bly lower calcite saturation index of the drip water due to

Figure 5. Detail of the upper 16 mm of the Proserpine core drilled in2011. Ten additional dark layers are counted compared to the coreof Verheyden et al. (2006), confirming the seasonal character of thelayering. Ages are based on layer counting, with one dark and onewhite layer deposited every year. The frame indicates the locationof the monthly resolved �18O and �13C measurements.

enhanced PCP causes a decreased in the calcite growth ratein the speleothem. Darker and more compact calcite is sug-gested to reflect slower growth rate conditions. Further in-vestigation in the seasonal cycle of the carbonate-dissolvedion concentration of the Proserpine drip water is necessaryto confirm this hypothesis.

6 Conclusions

Through a biweekly cave monitoring and a high-resolutionstable isotopic record of a recent finely laminated part ofa growing speleothem, the seasonal variation of the caveenvironment and how these parameters are recorded in thespeleothem proxies are documented.

1. The temperature effect clearly influences the �18O and�D composition of the rainwater with increasing valuesduring the summer months. The �18O and �D compo-sition of the drip water remains constant, indicating along residence time of the water in the epikarst and awell-mixed aquifer. A residence time of more than oneyear is assumed with the first waters percolating intothe cave 5 to 6 months after entering the soil. Such longwater residence time in the epikarst combined with a re-action time of 24 h to a heavy-rainfall event indicates apiston flow hydrology in the epikarst.

2. The records of the different measured factors in andoutside the cave suggest that the physico-chemical be-havior of the Han-sur-Lesse cave closely responds tothe seasonally varying external climate. The cave inte-rior climatology varies between a summer mode lastingfrom June to December and a winter mode from De-cember to June. During the summer mode, the cave airand drip water temperatures are higher, the drip flow

Clim. Past, 10, 1–15, 2014 www.clim-past.net/10/1/2014/

Maite Van Rampelbergh
Chapter 4: Cave Monitoring
Maite Van Rampelbergh
Maite Van Rampelbergh
98
Page 109: Maïté Van Rampelbergh

M. Van Rampelbergh: Monitoring of a fast-growing speleothem site 13

decreases the cave air pCO2 increases. During the win-ter mode the temperature cools down and the vegetationcover becomes less active, leading to an opposite be-havior of the measured factors compared to the summermode.

3. Seasonal variation in drip water pH and �13C valuemost probably relate to the seasonality in PCP intensity,driven by seasonal changes in water discharge. Duringsummer, lower recharge will cause stronger PCP, whichgradually increases the pH and �13C composition of thedrip water between June and September. In winter, thevadose zone gets recharged with water and PCP de-creases. The pH and �13C composition of the drip waterdecreases gradually from October trough May.

4. Calcite precipitated on glass slabs indicates more neg-ative �18O and more positive �13C values in summer.Both isotopic signals are deposited in equilibrium withtheir drip water. The seasonal variations of the �18Ovalues of precipitated calcite are caused by the sea-sonal temperature variations inside the cave. The sea-sonal variations of the �13C values are mainly controlledby the seasonal variation in drip water �13C caused byseasonally varying PCP.

5. The studied part of the Proserpine stalagmite displaysseasonal layering with one dark and one white layer de-posited every year. The opposite seasonal behavior ofthe �18O and the �13C signals on the glass slabs is alsovisible in the isotopic signals of the individual layersin the stalagmite. Dark layers have low �18O and high�13C values, while white layers have high �18O and low�13C values. Dark layers are formed during the cave’ssummer mode, while white layers grow calcite duringthe cave’s winter mode. The clearer the lamination andcoloring of the layers, the larger the amplitudes of thevariations in the isotopic composition.

Acknowledgements. We thank the “Domaine des Grottes de Han”for allowing us to sample the stalagmite and visit the cave on abiweekly basis, Etienne Lannoy for sampling the rain and dripwater during the monitoring period, Claire Mourgues and DavidVerstraeten for the help in the lab and Mr. Van Dierendonck for hisinterest and support. P. Claeys thanks the Hercules Foundation andResearch Foundation Flanders (FWO, project G-0422-10).

Edited by: J. Guiot

References

Baker, A. and Genty, A.: Environmental pressures on conservingcave speleothems: effects of changing surface land use and in-creased cave tourism, J. Environ. Manage., 53, 165–175, 1998.

Baker, A., Genty, D., Dreybrodt, W., Barnes, W. L., Mockler, N. J.,and Grapes, J.: Testing theoretically predicted stalagmite growthrate with Recent annually laminated samples: implications forpast stalagmite deposition, Geochim. Cosmochim. Ac., 62, 393–404, 1998.

Baker, A., Asrat, A., Fairchild, I. J., Leng, M. J., Wynn, P. M.,Bryant, C., Genty, D., and Umer, M.: Analysis of the climatesignal contained within delta O-18 and growth rate parametersin two Ethiopian stalagmites, Geochim. Cosmochim. Ac., 71,2975–2988, 2007.

Baldini, J. U. L., McDermott, F., Hoffmann, D. L., Richards, D. A.,and Clipson, N.: Very high-frequency and seasonal cave atmo-sphere pCO2 variability: Implications for stalagmite growth andoxygen isotope-based paleoclimate records, Earth Planet. Sci.Lett., 272, 118–129, 2008.

Bonniver, I.: Etude Hyrogeologique et dimmensionnement parmodelisation du “systeme-tracage” du reseau karstique the Han-sur-Lesse (Massif de Boine, Belgique), 2011, Geologie, FUNDPNamur, Namur, 93–97, 2011.

Chacko, T. and Deines, P.: Theoretical calculation of oxygen iso-tope fractionation factors in carbonate systems, Geochim. Cos-mochim. Ac., 72, 3642–3660, 2008.

Deininger, M., Fohlmeister, J., Scholz, D., and Mangini, A.: Isotopedisequilibrium effects: The influence of evaporation and ventila-tion effects on the carbon and oxygen isotope composition ofspeleothems – A model approach, Geochim. Cosmochim. Ac.,96, 57–79, 2012.

Delvaux De Fenffe, D.: Geologie et tectonique du Parc de Lesse etLomme au bord sud du Bassin de Dinant (Rochefort, Belgique),Bulletins de la Societe Belge de Geologie, 94, 81–95, 1985.

Demeny, A., Kele, S., and Siklosy, Z.: Empirical equations for thetemperature dependence of calcite-water oxygen isotope frac-tionation from 10 to 70 �C, Rapid Commun. Mass Sp., 24, 3521–3526, 2010.

Dreybrodt, W.: Chemical kinetics, speleothem growth and climate,Boreas, 28, 347–356, 1999.

Dreybrodt, W.: Evolution of the isotopic composition of carbon ina calcite precipitating H2O-CO2-CaCO3 solution and the relatedisotopic composition of calcite in stalagmites, Geochim. Cos-mochim. Ac., 72, 4712–4724, 2008.

Dreybrodt, W. and Scholz, D.: Climatic dependence of stable car-bon and oxygen isotope signals recorded in speleothems: Fromsoil water to speleothem calcite, Geochim. Cosmochim. Ac., 75,734–752, 2011.

Dulinski, M. and Rozanski, K.: Formation of 12C/13C isotope ratiosin spsleothems: a semi-dynamic model, Radiocarbon, 32, 7–16,1990.

Ek, C. and Godissart, J.: Carbon dioxide in cave air and soil airin some karstic areas of Belgium, A prospective view, Geolog.Belgica, 17, 102–106, 2014.

Emrich, K., Ehhalt, D. H., and Volgel, J. C.: Carbon isotope frac-tionation during the precipitation of calciumcarbonates, EarthPlanet. Sci. Lett., 8, 363–371, 1970.

Fairchild, I. J., Borsato, A., Tooth, A. F., Frisia, S., Hawkesworth,C. J., Huang, Y. M., McDermott, F., and Spiro, B.: Controls on

www.clim-past.net/10/1/2014/ Clim. Past, 10, 1–15, 2014

Maite Van Rampelbergh
Chapter 4: Cave Monitoring
Maite Van Rampelbergh
Maite Van Rampelbergh
Maite Van Rampelbergh
99
Maite Van Rampelbergh
Page 110: Maïté Van Rampelbergh

14 M. Van Rampelbergh: Monitoring of a fast-growing speleothem site

trace element (Sr-Mg) compositions of carbonate cave waters:implications for speleothem climatic records, Chem. Geol., 166,255–269, 2000.

Fairchild, I. J., Smith, C. L., Baker, A., Fuller, L., Spotl, C., Mat-tey, D., McDermott, F., and Eimp: Modification and preserva-tion of environmental signals in speleothems, Earth-Sci. Rev., 75,105–153, 2006.

Fairchild, I. J. and Treble, P. C.: Trace elements in speleothems asrecorders of environmental change, Quat. Sci. Rev., 28, 449–468,2009.

Friedman, I. and O’Neil, J. R.: Compilation of stable isotope frac-tionation factors of geochemical interest, Geolog. Surv. Prof. Pa-per 440-KK, 177, 1977.

Genty, D., Baker, A., and Vokal, B.: Intra- and inter-annual growthrate of modern stalagmites, Chem. Geol., 176, 191–212, 2001.

Genty, D., Blamart, D., Ouahdi, R., Gilmour, M., Baker, A.,Jouzel, J., and Van-Exter, S.: Precise dating of Dansgaard–Oeschger climate oscillations in western Europe from stalagmitedata, Nature, 421, 833–837, 2003.

Genty, D. and Deflandre, G.: Drip flow variations under a stalactiteof the Pere Noel cave (Belgium). Evidence of seasonal variationsand air pressure constraints, J. Hydrol., 211, 208–232, 1998.

Genty, D. and Quinif, Y.: Annually laminated sequences in the in-ternal structure of some Belgian stalagmites – Importance for pa-leoclimatology, J. Sediment. Res., 66, 275–288, 1996.

Hendy, C. H.: Isotopic geochemistry of speleothems: 1. Calcula-tions of effects on different modes of formation on isotopic com-position of speleothems and their applicability as paleoclimaticindicators, Geochim. Cosmochim. Ac., 35, 801–824, 1971.

Horita, J. and Clayton, R. N.: Comment on the studies of oxygenisotope fractionation between calcium carbonates and water atlow temperatures by Zhou and Zheng (2003, 2005), Geochim.Cosmochim. Ac., 71, 3131–3135, 2007.

Jimenez-Lopez, C., Caballero, E., Huertas, F. J., and Ro-manek, C. S.: Chemical, mineralogical and isotope behavior, andphase transformation during the precipitation of calcium carbon-ate minerals from intermediate ionic solution at 25 �C, Geochim.Cosmochim. Ac., 65, 3219–3231, 2001.

Jochum, K. P., Scholz, D., Stoll, B., Weis, U., Wilson, S. A., Yang,Q., Schwalb, A., Boerner, N., Jacob, D. E., and Andreae, M.O.: Accurate trace element analysis of speleothems and biogeniccalcium carbonates by LA-ICP-MS, Chem. Geol., 318, 31–44,2012.

Kim, S.-T., Mucci, A., and Taylor, B. E.: Phosphoric acid fraction-ation factors for calcite and aragonite between 25 and 75 �C: re-visited, Chem. Geol., 246, 135–146, 2007.

Kim, S. T. and O’Neil, J. R.: Equilibrium and nonequilibriumoxygen isotope effects in synthetic carbonates, Geochim. Cos-mochim. Ac., 61, 3461–3475, 1997.

Kluge, T., Riechelmann, D. F. C., Wieser, M., Spötl, C., Sultenfuss,J., Schroder-Ritzrau, A., Niggemann, S., and Aeschbach-Hertig,W.: Dating cave drip water by tritium, J. Hydrol., 394, 396–406,2010.

Kottek, M., Grieser, J., Beck, C., Rudolf, B., and Rubel, F.: Worldmap of the Koppen-Geiger climate classification updated, Mete-orologische Zeitschrift, 15, 259–263, 2006.

Lachniet, M. S.: Climatic and environmental controls onspeleothem oxygen-isotope values, Quat. Sci. Rev., 28, 412-432,2009.

Mattey, D., Lowry, D., Duffet, J., Fisher, R., Hodge, E., and Frisia,S.: A 53 year seasonally resolved oxygen and carbon isotoperecord from a modem Gibraltar speleothem: Reconstructed dripwater and relationship to local precipitation, Earth Planet. Sci.Lett., 269, 80–95, 2008.

McDermott, F.: Centennial-scale Holocene climate variability re-vealed by a high-resolution speleothem �18O record from SWIreland (9 November, pg 1328, 2001), Science, 309, 1816–1816,2005.

McDermott, F., Schwarcz, H. P., and Rowe, P. J.: Isotopesin speleothems, in: Isotopes in Paleoenvironmental Research,edited by: Leng, M. J., Springer Netherlands, 185–226, 2005.

Miorandi, R., Borsato, A., Frisia, S., Fairchild, I. J., and Richter, D.K.: Epikarst hydrology and implications for stalagmite captureof climate changes at Grotta di Ernesto (NE Italy): results fromlong-term monitoring, Hydrol. Proc., 24, 3101–3114, 2010.

Mühlinghaus, C., Scholz, D., and Mangini, A.: Modelling stalag-mite growth and �13C as a function of drip interval and tempera-ture, Geochim. Cosmochim. Ac., 71, 2780–2790, 2007.

Mühlinghaus, C., Scholz, D., and Mangini, A.: Modelling fraction-ation of stable isotopes in stalagmites, Geochim. Cosmochim.Ac., 73, 7275–7289, doi:10.1016/j.gca.2009.09.010, 2009.

O’Neil, J. R., Clayton, R. N., and Mayeda, T. K.: Oxygen isotopefractionation in divalent metal carbonates J. Chem. Phys., 51,5547–5558, 1969.

Oster, J. L., Montanez, I. P., and Kelley, N. P.: Response of a moderncave system to large seasonal precipitation variability, Geochim.Cosmochim. Ac., 91, 92–108, 2012.

Piccini, L., Zanchetta, G., Drysdale, R. N., Hellstrom, J., Isola,I., Fallick, A. E., Leone, G., Doveri, M., Mussi, M., Mantelli,F., Molli, G., Lotti, L., Roncioni, A., Regattieri, E., Meccheri,M., and Vaselli, L.: The environmental features of the MonteCorchia cave system (Apuan Alps, central Italy) and their effectson speleothem growth, International Journal of Speleology, 37,153–172, 2008.

Quinif, Y.: Une nouvelle topographie de la Grotte de Han, Lapiazhors serie “Special Han”, 15–18, 1988.

Riechelmann, D. F. C., Deininger, M., Scholz, D., Riechelmann, S.,Schroeder-Ritzrau, A., Spoetl, C., Richter, D. K., Mangini, A.,and Immenhauser, A.: Disequilibrium carbon and oxygen isotopefractionation in recent cave calcite: Comparison of cave precipi-tates and model data, Geochim. Cosmochim. Ac., 103, 232–244,2013.

Riechelmann, D. F. C., Schroeder-Ritzrau, A., Scholz, D.,Fohlmeister, J., Spoetl, C., Richter, D. K., and Mangini, A.: Mon-itoring Bunker Cave (NW Germany): A prerequisite to interpretgeochemical proxy data of speleothems from this site, J. Hydrol.,409, 682–695, 2011.

Rozanski, K., Araguasaraguas, L., and Gonfiantini, R.: Relationshipbetween long-term trends of 18O isotope composition of precip-itation and climate, Science, 258, 981–985, 1992.

Ruan, J. and Hu, C.: Seasonal variations and environmental controlson stalagmite calcite crystal growth in Heshang Cave, CentralChina, Chinese Sci. Bull., 55, 3929–3935, 2010.

Scholz, D., Muehlinghaus, C., and Mangini, A.: Modelling �13Cand �18O in the solution layer on stalagmite surfaces, Geochim.Cosmochim. Ac., 73, 2592–2602, 2009.

Smart, P. L. and Friedrich, H.: Water movement and storage in theunsaturated zone of amaturely karstified aquifer, Mendip Hills,

Clim. Past, 10, 1–15, 2014 www.clim-past.net/10/1/2014/

Maite Van Rampelbergh
Chapter 4: Cave Monitoring
Maite Van Rampelbergh
Maite Van Rampelbergh
100
Maite Van Rampelbergh
Page 111: Maïté Van Rampelbergh

M. Van Rampelbergh: Monitoring of a fast-growing speleothem site 15

England, Proceedings of the Conference on Environ-mentalProblems in Karst Terrains and Teir Solution, Bowling Green,Kentucky, National Water Well Association, 57–87, 1987.

Spötl, C., Fairchild, I. J., and Tooth, A. F.: Cave air controlon dripwater geochemistry, Obir Caves (Austria): Implicationsfor speleothem deposition in dynamically ventilated caves,Geochim. Cosmochim. Ac., 69, 2451–2468, 2005.

Spötl, C. and Mattey, D.: Stable isotope microsampling ofspeleothems for palaeoenvironmental studies: A comparison ofmicrodrill, micromill and laser ablation techniques, Chem. Geol.,235, 48–58, 2006.

Tarutani, T., Clayton, R. N., and Mayeda, T. K.: Effect of poly-morphism and magnesium substitution on oxygen isotope frac-tionnation between calcium carbonate and water, Geochim. Cos-mochim. Ac., 33, 987–996, 1969.

Thornthwaite, C.: An approaxh toward a rational classification ofclimate, Geographical Review, 38, 55–94, 1948.

Treble, P., Shelley, J. M. G., and Chappell, J.: Comparison of highresolution sub-annual records of trace elements in a modern(1911–1992) speleothem with instrumental climate data fromsouthwest Australia, Earth Planet. Sci. Lett., 216, 141–153, 2003.

Tremaine, D. M., Froelich, P. N., and Wang, Y.: Speleothem calcitefarmed in situ: Modern calibration of delta �18O and �13C paleo-climate proxies in a continuously-monitored natural cave system,Geochim. Cosmochim. Ac., 75, 4929–4950, 2011.

Van Rampelbergh, M., Fleitmann, D., Verheyden, S., Cheng, H.,Edwards, L., De Geest, P., De Vleeschouwer, D., Burns, S. J.,Matter, A., Claeys, P., and Keppens, E.: Mid- to late HoloceneIndian Ocean Monsoon variability recorded in four speleothemsfrom Socotra Island, Yemen, Quat. Sci. Rev., 65, 129–142, 2013.

Verheyden, S.: Speleothems as palaeoclimatic archives, Isotopicand geochemical study of the cave environment and its Late Qua-ternary records, Unpubl. PhD Thesis, Vrije Universiteit Brussel,Belgium, 132 p., 2001.

Verheyden, S., Baele, J.-M., Keppens, E., Genty, D., Cattani, O.,Hai, C., Edwards, L., Hucai, Z., Van Strijdonck, M., andQuinif, Y.: The proserpine stalagmite (Han-sur-Lesse cave,Belgium): preliminary environmental interpretation of the last1000 years as recorded in a layered speleothem, Geol. Belg., 9,245–256, 2006.

Verheyden, S., Genty, D., Deflandre, G., Quinif, Y., and Kep-pens, E.: Monitoring climatological, hydrological and geochem-ical parameters in the Pere Noel cave (Belgium): implication forthe interpretation of speleothem isotopic and geochemical time-series, Int. J. Speleol., 37, 221–234, 2008a.

Verheyden, S., Nader, F. H., Cheng, H. J., Edwards, L. R., andSwennen, R.: Paleoclimate reconstruction in the Levant regionfrom the geochemistry of a Holocene stalagmite from the Jeitacave, Lebanon, Quaternary Res., 70, 368–381, 2008b.

Wang, Y., Cheng, H., Edwards, R. L., Kong, X., Shao, X., Chen, S.,Wu, J., Jiang, X., Wang, X., and An, Z.: Millennial- andorbital-scale changes in the East Asian monsoon over the past224 000 years, Nature, 451, 1090–1093, 2008.

www.clim-past.net/10/1/2014/ Clim. Past, 10, 1–15, 2014

Maite Van Rampelbergh
Chapter 4: Cave Monitoring
Maite Van Rampelbergh
Maite Van Rampelbergh
101
Maite Van Rampelbergh
Page 112: Maïté Van Rampelbergh

Chapter  4:  Cave  Monitoring  

102  

   

Page 113: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium    

  103  

Chapter  5        

A   500-­‐year   speleothem   multiproxy  record   from   the   Han-­‐sur-­‐Lesse   cave,  Belgium      Different   studies   have   indicated   that   speleothem   proxies   successfully   record  climate   variations   in  Western   Europe   (Frisia   et   al.,   2003;  Mangini   et   al.,   2005;  Verheyden  et  al.,  2006;  Mattey  et  al.,  2008;  Baker  et  al.,  2011;  Fohlmeister  et  al.,  2012).   In   Belgium,   the   Proserpine   stalagmite   from   the   Han-­‐sur-­‐Lesse   cave,  displays   extremely   high-­‐growth   rates   (up   to   2   mm/y)   and   seasonal   layering  allowing   to   measure   different   proxies   at   seasonal   scale   back   to   2000   BP  (Verheyden   et   al.,   2006).   In   this   PhD,  we   focused   on   the   clearly   layered  upper  500  years  of  the  Proserpine,  containing  a  large  part  of  Little  Ice  Age  (LIA,  ±1300  to   1850   AD)   (Jones   and   Mann,   2004)   as   well   as   the   ongoing   anthropogenic  climate  warming  of  the  last  150  years  (Mann  et  al.,  1998,  1999;  2003;  2008).  The  results   of   the   cave   monitoring,   discussed   in   Chapter   4,   are   used   to   link   the  measured  variations  in  the  δ18O  and  δ13C  values,  layer  thickness  and  calcite  color  with   climate.   The   observed   variations   are   then   compared   with   historical,  instrumental  records  and  other  temperature  reconstructions  to  investigate  how  recognized   regional   climate   variations   are   recorded   in   the   (local)   speleothem  proxies.   Conclusions   are   that   the   δ18O   and   δ13C   signals,   layer   thickness   and  calcite   fabric   of   the   Proserpine   can   successfully   be   used   to   investigate   climate  variations   over   the   last   500   years.   The   recorded   variations   compare  well  with  known   regional   climate   changes.   Moreover,   the   seasonally   resolved   proxies  provide   additional   information   on   the   amplitude   of   the   seasonal   variations  during  the  LIA.  The  Proserpine  has  demonstrated  to  successfully  record  climate  at  seasonal  scale  for  the  most  recent  500  years.  Older  sections  can  most  probably  be  used  in  further  studies  to  investigate  in  more  detail  for  example  the  Medieval  Warm  Period  (±950-­‐1250  AD).        

Page 114: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium    

 104  

 REFERENCES    Baker,  A.,  Wilson,  R.,  Fairchild,  I.  J.,  Franke,  J.,  Spoetl,  C.,  Mattey,  D.,  Trouet,  V.,  and  Fuller,   L.:   High   resolution   delta  O-­‐18   and   delta   C-­‐13   records   from   an   annually  laminated   Scottish   stalagmite   and   relationship   with   last   millennium   climate,  Glob.  Planet.  Change,  79,  303-­‐311,  2011.  

Cook,   E.   R.,   D'Arrigo,   R.   D.,   and   Mann,   M.   E.:   A   well-­‐verified,   multiproxy  reconstruction   of   the   winter   North   Atlantic   Oscillation   index   since   AD   1400,  Journal  of  Climate,  15,  1754-­‐1764,  2002.  

Esper,   J.,   Frank,   D.,   Buntgen,   U.,   Verstege,   A.,   and   Luterbacher,   J.:   Long-­‐term  drought  severity  variations  in  Morocco,  Geophysical  Research  Letters,  34,  2007.  

Fohlmeister,   J.,   Schroder-­‐Ritzrau,   A.,   Scholz,   D.,   Spotl,   C.,   Riechelmann,   D.   F.   C.,  Mudelsee,   M.,   Wackerbarth,   A.,   Gerdes,   A.,   Riechelmann,   S.,   Immenhauser,   A.,  Richter,  D.  K.,   and  Mangini,  A.:  Bunker  Cave   stalagmites:   an   archive   for   central  European  Holocene  climate  variability,  Climate  of  the  Past,  8,  1751-­‐1764,  2012.  

Frisia,  S.,  Borsato,  A.,  Preto,  N.,  and  McDermott,  F.:  Late  Holocene  annual  growth  in  three  Alpine  stalagmites  records  the  influence  of  solar  activity  and  the  North  Atlantic  Oscillation  on  winter  climate,  Earth  and  Planetary  Science  Letters,  216,  411-­‐424,  2003.  

Hurrell,   J.   W.:   Decadal   trends   in   the   North   Atlantic   Osscilation-­‐   Regional  temperatures  and  precipitation  Science,  269,  676-­‐679,  1995.  

Jones,  P.  D.  and  Mann,  M.  E.:  Climate  over  past  millennia,  Reviews  of  Geophysics,  42,  2004.  

Luterbacher,   J.,   Xoplaki,   E.,   Dietrich,   D.,   Jones,   P.   D.,   Davies,   T.   D.,   Portis,   D.,  Gonzalez-­‐Rouco,   J.   F.,   von   Storch,   H.,   Gyalistras,   D.,   Casty,   C.,   and   Wanner,   H.:  Extending  North  Atlantic  Oscillation  reconstructions  back  to  1500,  Atmospheric  Science  Letters,  2,  114-­‐124,  2001.  

Mangini,   A.,   Spotl,   C.,   and   Verdes,   P.:   Reconstruction   of   temperature   in   the  Central  Alps  during  the  past  2000  yr  from  a  delta  O-­‐18  stalagmite  record,  Earth  and  Planetary  Science  Letters,  235,  741-­‐751,  2005.  

Mann,  M.  E.,  Bradley,  R.  S.,  and  Hughes,  M.  K.:  Global-­‐scale  temperature  patterns  and  climate  forcing  over  the  past  six  centuries,  Nature,  392,  779-­‐787,  1998.  

Mann,  M.  E.,  Bradley,  R.  S.,  and  Hughes,  M.  K.:  Northern  hemisphere  temperatures  during   the   past   millennium:   Inferences,   uncertainties,   and   limitations,  Geophysical  Research  Letters,  26,  759-­‐762,  1999.  

Mann,   M.   E.   and   Jones,   P.   D.:   Global   surface   temperatures   over   the   past   two  millennia,  Geophysical  Research  Letters,  30,  2003.  

Mann,  M.  E.,   Zhang,  Z.,  Hughes,  M.  K.,  Bradley,  R.   S.,  Miller,   S.  K.,  Rutherford,   S.,  and   Ni,   F.:   Proxy-­‐based   reconstructions   of   hemispheric   and   global   surface  

Page 115: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium    

  105  

temperature  variations  over  the  past  two  millennia,  Proceedings  of  the  National  Academy  of  Sciences  of  the  United  States  of  America,  105,  13252-­‐13257,  2008.  

Mattey,   D.,   Lowry,   D.,   Duffet,   J.,   Fisher,   R.,   Hodge,   E.,   and   Frisia,   S.:   A   53   year  seasonally  resolved  oxygen  and  carbon  isotope  record  from  a  modem  Gibraltar  speleothem:   Reconstructed   drip   water   and   relationship   to   local   precipitation,  Earth  and  Planetary  Science  Letters,  269,  80-­‐95,  2008.  

Olsen,   J.,   Anderson,   N.   J.,   and   Knudsen,   M.   F.:   Variability   of   the   North   Atlantic  Oscillation  over  the  past  5,200  years,  Nature  Geoscience,  5,  808-­‐812,  2012.  

Proctor,   C.   J.,   Baker,   A.,   Barnes,   W.   L.,   and   Gilmour,   R.   A.:   A   thousand   year  speleothem   proxy   record   of   North   Atlantic   climate   from   Scotland,   Climate  Dynamics,  16,  815-­‐820,  2000.  

Trouet,   V.,   Esper,   J.,   Graham,   N.   E.,   Baker,   A.,   Scourse,   J.   D.,   and   Frank,   D.   C.:  Persistent   Positive   North   Atlantic   Oscillation   Mode   Dominated   the   Medieval  Climate  Anomaly,  Science,  324,  78-­‐80,  2009.  

Verheyden,  S.,  Baele,  J.-­‐M.,  Keppens,  E.,  Genty,  D.,  Cattani,  O.,  Hai,  C.,  Edwards,  L.,  Hucai,  Z.,  Van  Strijdonck,  M.,  and  Quinif,  Y.:  The  proserpine  stalagmite  (Han-­‐sur-­‐Lesse  cave,  Belgium):  Preliminary  environmental  interpretation  of  the  last  1000  years  as  recorded  in  a  layered  speleothem,  Geologica  Belgica,  9,  245-­‐256,  2006.  

Wanner,  H.,  Bronnimann,  S.,  Casty,  C.,  Gyalistras,  D.,  Luterbacher,  J.,  Schmutz,  C.,  Stephenson,   D.   B.,   and   Xoplaki,   E.:   North   Atlantic   Oscillation   -­‐   Concepts   and  studies,  Surveys  in  Geophysics,  22,  321-­‐382,  2001.                                              

Page 116: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium    

 106  

   

Page 117: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium      

Paper  in  review  in  Climate  of  the  Past   107  

A  500-­‐year   seasonally   resolved  δ18O  and  δ13C,   layer   thickness  and  calcite  fabric  record  from  a  speleothem  deposited  in  equilibrium  of  the  Han-­‐sur-­‐Lesse  cave,  Belgium.    Maïté   Van   Rampelbergh1,   Sophie   Verheyden1,2,   Mohammed   Allan3,   Yves  Quinif4,   Hai   Cheng5,6,   Lawrence   Edwards6,   Edward  Keppens1  and   Philippe  Claeys1    1Earth   System   Sciences,   Vrije   Universiteit   Brussel   (VUB),   Pleinlaan,   B-­‐1050,  Brussels,  Belgium  2Royal  Belgian   Institute   of  Natural   Sciences,  Geological   Survey,  Direction  Earth  and  History  of  Life,  Jennerstraat  13,  B-­‐1000,  Brussels,  Belgium  3AGEs,  Départment  de  Géologie,  Université  de  Liège,  Allée  du  6  Août,  B18  Sart-­‐Tilman,  B-­‐4000,  Liège,  Belgium  4Faculté   Polytechnique,   Université   de  Mons,   Rue   de   Houdain   9,   B-­‐7000,  Mons,  Belgium  5   Institute   of   Global   Environmental   Change,   Xi'an   Jiaotong   University,   Xi'an  710049,  China    6  Department  of  Geological   Sciences,  University  of  Minnesota,  100  Union  Street  SE,  Minneapolis  MN  55455,  USA    Abstract  Speleothem   δ18O   and   δ13C   signals   have   already   proven   to   enable   climate  reconstructions   at   high   resolution.   However,   decadal   and   seasonally   resolved  speleothem   records   are   still   scarce   and   often   difficult   to   interpret   in   terms   of  climate  due   to   the  multitude  of   factors   that  can  affect   the  proxy  signals.   In   this  paper,  a  fast  growing  (up  to  2  mm/y)  seasonally  laminated  speleothem  from  the  Han-­‐sur-­‐Lesse   cave   (Belgium)   is   analyzed   for   its   δ18O   and   δ13C   values,   layer  thickness   and   changes   in   calcite   fabric.   The   studied   part   of   the   speleothem  covers  the  period  between  2001  and  1479  AD  as  indicated  by  layer  counting  and  confirmed   by   20   U/Th-­‐ages.   Recharge  mainly   occurs   during  winter   and   lesser  during  spring  and  fall  causing  the  Proserpine  proxies  to  be  seasonally  biased.  On  decadal   and   multi-­‐decadal   scale,   increased   δ18O   values   in   the   Proserpine   are  interpreted   to   reflect   drier   (and   colder)   winters.   Higher   δ13C   signals   are  interpreted   to   reflect   increased  prior   calcite  precipitation   (PCP)  and   lower   soil  activity   during   drier   (and   colder)   winters.   Thinner   layers   and   darker   calcite  relate  to  slower  growth  and  both  occur  during  drier  (and  colder)  winter  periods.  Exceptionally   dry   (and   cold)   winter   periods   induce   simultaneous   large-­‐amplitude  shifts  in  the  four  proxies.  Such  anomalies  occur  from  1565  to  1610,  at  1730,   from   1770   to   1800,   from   1810   to   1860   and   from   1880   to   1895   and  correspond   with   exceptionally   cold   periods   in   historical   and   instrumental  records   as   well   as   in   European   winter   temperature   reconstructions.   More  relative   climate   variations,   during   which   the   four   measured   proxies   vary  independently  and  display  lower  amplitude  variations,  occur  between  1479  and  1565,  between  1610  and  1730  and  between  1730  and  1770.  The  winters  during  the   two   periods   between   1479   and   1565   and   between   1730   and   1770   are  interpreted   as   relatively   wetter   (and   warmer)   and   correspond   with   warmer  periods  in  historical  data  and  in  winter  temperature  reconstructions  in  Europe.  The  winters  in  the  period  between  1610  and  1730  are  interpreted  as  relatively  

Page 118: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium    

Paper  in  review  in  Climate  of  the  Past    

108  

drier   (and   cooler)   and   correspond  with   generally   colder   conditions   in   Europe.  Interpretation  of  the  seasonal  variations  in  δ18O  and  δ13C  signals  differs  from  the  interpretation   on   decadal   and   multi-­‐decadal   scale.   Seasonal   δ18O   variations  reflect  cave  air   temperature  variations  and  suggest  a  2.5  °C  seasonality   in  cave  air   temperature  during   the   two  relatively  wetter   (and  warmer)  winter  periods  (1479-­‐1565   and   1730-­‐1770),   which   corresponds   to   the   cave   air   temperature  seasonality  observed  today.  Between  1610  and  1730,   the  δ18O  values  suggest  a  1.5   °C   seasonality   in   cave   air   temperature   suggesting   colder   summer  temperatures   during   this   drier   (and   cooler)   interval.   The   δ13C   seasonality   is  driven  by  PCP  and  suggests  generally  lower  PCP  seasonal  effects  between  1479  and  1810,  compared  to  the  PCP  seasonal  effect  observed  today.  A  short  interval  of   increased   PCP-­‐seasonality   occurs   between   1600   and   1660,   and   reflects  increased  PCP  in  summer  due  to  decreased  winter  recharge.    1.  Introduction    In   the   studied   western   European   region,   high-­‐resolution   climate   records  covering  the   last  500  years  are  scarce.  Most  climate   information  at  seasonal  or  yearly  scale  is  retrieved  from  historical  data  such  as  the  price  of  flour  or  grapes  (Van  Engelen  et  al.,  2001;  Le  Roy  Ladurie,  2004)  which  may  induce  biases  in  the  climate  record.  Therefore  it   is  necessary  to  confront   information  from  different  archives,  based  on  different  approaches.      Speleothems  have  already  often  proven  to  enable  climate  reconstruction  at  high-­‐resolution   in   Europe   (Genty   et   al.,   2003;   Baker   et   al.,   2011;  McDermott   et   al.,  2011;   Fohlmeister   et   al.,   2012;   Verheyden   et   al.,   2014).   On   millennial   and  centennial   scales,   the  δ18O   and  δ13C   variations   can   often   be   related   to   a   single  climate  proxy  such  as  temperature  or  vegetation  cover  (Spötl  and  Mangini,  2002;  Genty  et  al.,  2003;  McDermott,  2005).  However,  on  decadal  and  seasonal  scale,  a  larger   range   of   factors   can   influence   the   δ18O,   δ13C,   layer   thickness   or   calcite  fabric   of   a   speleothem   making   an   interpretation   in   terms   of   climate   more  difficult.  To  allow  reconstructing  the  climate  up  to  seasonal  variation  using  mid-­‐latitude  speleothems,  a  detailed  analysis  of  each  used  proxy  must  be  compared  with   a   multiproxy   approach.   Different   European   records   have   enabled   to  reconstruct   climate   successfully  by  using   this  approach   (e.g.  Frisia  et   al.,   2003;  Niggemann  et  al.,  2003;  Mangini  et  al.,  2005;  Mattey  et  al.,  2008;  Fohlmeister  et  al.,  2012).    Belgian   speleothems   have   the   valuable   advantage   to   often   display   a   clear  internal   layered   structure   reflecting   seasonal   variations   (Genty   and   Quinif,  1996).  The  link  between  layer  thickness  and  water  excess  in  Belgian  stalagmites  for  the  Late  Glacial  and  Holocene  period  has  clearly  been  demonstrated  by  Genty  and  Quinif  (1996).  The  δ18O  and  δ13C  signals  from  a  speleothem  sampled  in  the  Père  Noël  cave  were  interpreted  as  due  to  variations  in  cave  humidity  and  drip  rate  inducing  changes  in  the  kinetics  of  the  calcite  deposition  occurring  closer  or  less   close   to   isotopic   equilibrium.   More   negative   δ18O   and   δ13C   values   occur  during   periods   of   higher   cave   water   recharge,   when   calcite   deposition   occurs  closer   to   isotopic  equilibrium  (Verheyden  et  al.,  2008).     In   this  speleothem,   the  isotopic   (δ18O   and   δ13C)   and   geochemical   (Mg/Ca   and   Sr/Ca)   proxies   vary  

Page 119: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium      

Paper  in  review  in  Climate  of  the  Past   109  

similarly  and  record  the  climate  in  terms  of  wetter  and  drier  phases  (Verheyden  et   al.,   2014).   The   studied   Proserpine   stalagmite   is   a   large   tabular   shaped  speleothem,  growing   in   the  Han-­‐sur-­‐Lesse  cave,  which   is  part  of   the  same  cave  system  as  the  Père  Noël  cave.  A  former  study  of  the  stalagmite  (Verheyden  et  al.,  2006)  revealed  deposition  from  200  AD  to  2001  AD,  indicating  an  exceptionally  high  average  growth  rate  of  ±  1  mm/y.  The  upper  56  cm,  which  covers  the   last  522  years   is  clearly   layered.  The  similar  variability  of   the  δ18O  and  δ13C  signals  and  the  layer  thickness  was  linked  to  changes  in  effective  precipitation  (rainfall  minus   evapo-­‐transpiration).   These   proxies   therefore   have   the   potential   to   be  used  to  reconstruct  climate  in  terms  of  wetter  and  drier  phases.    In   this   paper   we   study   this   potential   more   in   detail   and   up   to   a   seasonally  resolved   timescale.   An   absolute   age   model   is   established   by   combining   layer-­‐counting   ages   with   measured   U/Th-­‐ages.   A   comparison   of   variations   in   layer  thickness,   calcite   fabric,   δ18O   and   δ13C   signals   in   the   light   of   former   studies  (Genty  and  Quinif,  1996;  Verheyden,  2001;  Genty  et  al.,  2003;  Mühlinghaus  et  al.,  2007;  Wackerbarth  et  al.,  2010;  Fohlmeister  et  al.,  2012;  Verheyden  et  al.,  2014)  and  monitoring  of  the  same  stalagmite   location  (Van  Rampelbergh  et  al.,  2014)  leads   to   a   better   understanding   of   how   these   proxies   are   related   among   them  and  how  they  reflect  climate  variations.  Comparing  the  Proserpine  climate  signal  with   winter   temperature   reconstructions   in   Europe   (Le   Roy   Ladurie,   2004;  Luterbacher  et  al.,  2004;  Dobrovolny  et  al.,  2010)   further  verifies   the  proposed  climate  interpretation.    2.  Study  area    The  Proserpine  stalagmite  is  sampled  in  the  Salle-­‐du-­‐Dôme  chamber  in  the  Han-­‐sur-­‐Lesse  cave,  southern  Belgium  (Fig.  1).  The  Han-­‐sur-­‐Lesse  cave  is  a  meander  cutting  of   the  Lesse-­‐river,  which   still   flows   through   the   cave.  The   large   rooms,  the  multiple   entrances   and   the   presence   of   the   river  make   it   a  well-­‐ventilated  cave.  Part  of  the  cave,  including  the  Salle-­‐du-­‐Dôme,  is  a  show  cave  since  the  mid  19th   century.  The  Salle-­‐du-­‐Dôme,  being   the   largest   chamber  of   the  cave  system  (150  m  wide   and   60  m   high),   is   located   under   ca.   40  m   of   Givetian   limestone  (Quinif,  1988)  with  a  C3-­‐type  vegetation  covered  soil.  The  Proserpine  stalagmite  is  a  2  m  high  stalagmite  with  a  large  tabular  shape  (with  a  horizontal  70  cm  by  150   cm   to   surface)   that   was   actively   growing   when   cored   in   2001.   A   rain   of  seepage  water  throughout  the  year  feeds  the  stalagmite.  Such  fast  growing  ‘tam-­‐tam’   shaped   stalagmites   have   the   property   to   record   climate   signals   and  environmental  information  at  high  resolution  (Perette,  2000).      The  mean  annual  precipitation  at  the  meteorological  station  of  Han-­‐sur-­‐Lesse  is  844   mm/y   and   the   mean   annual   air   temperature   averages   10.3°C   (Royal  Meteorological   Institute  Belgium,  hereafter  named  RMI)  characterizing  a  warm  temperate,   fully   humid   climate  with   cool   summers   (Kottek   et   al.,   2006).  While  the  temperature  displays  a  well-­‐marked  seasonality  with  cool  summers  and  mild  winters,  the  rainfall  is  spread  all  over  the  entire  year.  The  external  seasonality  in  temperature  causes  a  subdued  temperature  variation  within  the  Salle-­‐du-­‐Dôme  of   2   to   2.5   °C   between   summer   and   winter   (Van   Rampelbergh   et   al.,   2014).  Present-­‐day  calcite  is  deposited  in  isotopic  equilibrium  with  its  drip  water  (Van  

Page 120: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium    

Paper  in  review  in  Climate  of  the  Past    

110  

Rampelbergh  et  al.,  2014).  The  δ18O  signal  of  freshly  formed  calcite  collected  on  top  of  the  Proserpine  varies  seasonally  due  the  changes  in  cave  air  temperature.  The   δ13C   signal   varies   seasonally   due   to   changes   in   prior   calcite   precipitation  (PCP)  intensity,  driven  by  changes  in  effective  precipitation.  At  a  seasonal  scale  the  δ18O  and  δ13C  signals  display  an  opposite  behavior  with  more  negative  δ18O  values  in  summer,  when  the  δ13C  values  are   less  negative  (Van  Rampelbergh  et  al.,  2014).    

 Figure   1   a)   The   Han-­‐sur-­‐Lesse   cave   system   is   located   in   the   southern   part   of  Belgium.  The  Proserpine  stalagmite  was  sampled  in  the  Salle-­‐du-­‐Dôme  chamber  (white  square)   located  500  m   from  the  cave  exit.  b)  The  Proserpine  stalagmite  with  the  location  of  the  2  m  long  core  that  was  drilled  in  2001  at  the  spot  where  most  of  the  drip  water  falls.      3.  Methods    The  Proserpine  stalagmite  was  sampled  in  January  2001,  by  drilling  a  2  m  core  in  the   tabular   shaped   stalagmite.   The   precise   location   was   on   the   side   with   the  highest  drip  rate  but  far  enough  away  from  the  edge  to  avoid  disturbance  of  the  expected  horizontal   layering   of   the   growth   increments   (Fig.   1b).   The   core  was  cut  in  half  and  a  slab  of  1  cm  was  cut  from  the  center.  The  slabs  were  polished  by  hand  with  carbide  powder  and  finished  with  Al2O3.  The  upper  56  cm,  was  further  studied  and  cut  in  seven  parts,  numbered  I  to  VII  (Fig.  2),  to  allow  easy  handling  in  the  laboratory.  Layers  were  counted  per  part  under  the  Mercantec  Micromill  microscope  and  on  high-­‐resolution  scans  using  Adobe  Illustrator.  To  increase  the  reliability   of   the   layer   counting,   layers   were   counted   by   different   authors,   on  different  days  and  with  different  zooms  when  counted  on  computer  screen.  The  reported   layer  amount   is  given  by  the  average  of  10   layer  counting  rounds  per  part.  The   thickness  of  each   layer  was  measured  using   the  measurement   tool  of  the  Merchantec  Micromill  microscope  with  an  uncertainty  of  0.1  μm.  Samples  for  δ18O   and   δ13C   measurements   were   taken   with   a   drill   bit   of   0.3   mm   diameter  mounted  on  a  Merchantec  Micromill.  Ethanol  was  used  to  clean  the  speleothem  surface   and   drill   bit   prior   to   sampling.   Between   samplings,   drill   bit   and  speleothem   surface   were   cleaned   with   compressed   air.   Samples   were   drilled  every   0.5   mm   in   part   I   and   in   every   layer   for   the   other   parts,   in   total   867  samples.   Stable   isotope  measurements  were   carried   out   using   a  Kiel-­‐III-­‐device  coupled   on   a   Thermo  Delta   plus   XL  with   analytical   uncertainties   ≤   0.12‰   for  

Page 121: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium      

Paper  in  review  in  Climate  of  the  Past   111  

δ13C  and  ≤  0.16‰  for  δ18O.  A   total  of  20  U-­‐series  age,   among  which  8   from  a  former   study   (Verheyden   et   al.,   2006)   were   measured   at   the   University   of  Minnesota  (USA),  using  the  procedures  for  uranium  and  thorium  as  described  in  Edwards  et  al.  (1987)  and  Cheng  et  al.  (2000;  2009a;  2009b).  StalAge  (Scholz  and  Hoffmann,  2011)  was  used  to  interpolate  the  ages  between  the  U/Th-­‐age  points.  The  seasonal  character  of  the  layering  (Verheyden  et  al.,  2006;  Van  Rampelbergh  et   al.,   2014)   in   the   Proserpine   allows   using   layer   counting   to   establish   an   age  model.  The  amount  of  counted  layer  couplets  per  part  represents  the  amount  of  years   for   that   part.   The   amount   of   years   obtained   by   layer   counting   is   then  compared  with  the  amount  of  years  suggested  by  the  U/Th-­‐ages  per  part.  Results  of   both   independent   dating   methods   are   combined   to   provide   the   final   age  model.   The   uncertainties   (2σ)   on   all   reported   values   correspond   with   a   95%  confidence  interval  and  are  calculated  according  to  the  following  relation:    𝑥 − 𝑡!.!",!!! ∙ 𝑠 𝑛 ≤ 𝑥 ≤  𝑥 + 𝑡!.!",!!! ∙ 𝑠 𝑛  

 where  𝑥  is   the  arithmetic  mean  of   the   results,  n   the  number  of   replicates,   t   the  student  distribution  function  and  s  the  standard  deviation  on  the  results.   If  n  ≥  30,  t  approximates  a  normal  distribution  and  is  roughly  equal  to  2.    4.  Results    Layering   is   present   in   the   studied   upper   56   cm   of   the   Proserpine   core   and   is  formed   by   alternating   dark  more   compact   and  white  more   porous   layers.   The  seasonal   character   of   the   layering   in   the   Proserpine   stalagmite,  with   one   dark  and   one   white   layer   deposited   every   year   is   suggested   by   Verheyden   et   al.  (2006)   and   further   confirmed   by  monitoring   results   of   the   Proserpine   growth  site  (Van  Rampelbergh  et  al.,  2014).  The  Proserpine  stalagmite  displays  a  clear  sedimentological   perturbation   between   9   cm   and   10   cm   (Fig.   2).   During   this  perturbation,  calcite  deposition  is  heavily  disturbed  with  straw  pieces  embedded  in   the   calcite,   which   might   be   relics   from   fires   lit   on   the   paleo-­‐surface   of   the  stalagmite   to   illuminate   the   Salle-­‐du-­‐Dôme   (Verheyden   et   al.,   2006).   Four  proxies   were   measured   on   the   Proserpine   stalagmite:   calcite   fabric,   layer  thickness,   δ18O   and   δ13C   values.   Layer   thickness   varies   between   0.05   and   1.7  mm/layer   (Fig.  3)  and  dark   layers  are  on  average  0.05  mm  thinner   than  white  layers.  The  δ18O  values  average  -­‐6.9  ±  0.16  ‰  and  the  δ13C  values  average  -­‐10  ±  0.12  ‰.  Four   intervals   characterized  by   large  amplitude  variations  of   the   four  measured   proxies   occur   between   7   and   8   cm,   between   10.5   and   12.4   cm,  between  18  and  20  cm  and  between  34  and  36  cm  (blue  lines  Fig.  3).  Between  7  and  8  cm  and  between  34  and  36  cm,  calcite  fabric  is  dark  compact  with  almost  no  visible   layering.  During   these   two   intervals   layer   thickness  decreases   to  0.2  mm/layer  and  the  δ18O  and  δ13C  values  increase  to  values  around  -­‐6.0  ±  0.16  ‰  and   -­‐8.0   ±   0.12  ‰   respectively.   Between   10.5   and   12.4   cm,   calcite   is   heavily  altered  and  more  matte  and  whiter  compared  to  the  generally  more  translucent  calcite  fabric  of  the  Proserpine.  The  heat  of  the  fires  made  on  the  surface  of  the  stalagmite   during   the   perturbation   period  may   have   altered   the   calcite   in   this  part.  In  this  interval,  layer  thickness  decreases  to  0.2  mm/layer  and  the  δ18O  and  δ13C   values   increase   to   values   around   -­‐6.0   ±   0.16   ‰   and   -­‐6.5   ±   0.12   ‰  

Page 122: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium    

Paper  in  review  in  Climate  of  the  Past    

112  

respectively.   From   18   to   20   cm,   layering   is   heavily   undulating   with   vertically  orientated   layers   in   some   parts,   which   may   reflect   small   basin   or   rimstone  structures.   In   this   interval,   layer   thickness   decreases   to   0.4  mm/layer   and   the  δ18O   and   δ13C   values   increase   sharply   to   -­‐6.2   ±   0.16  ‰   and   -­‐7.0   ±   0.12  ‰  respectively.   With   the   exception   of   the   four   intervals   characterized   by  simultaneous  large  amplitude  variations  of  the  four  measured  proxies,  the  time-­‐series  can  be  subdivided  in  two  parts.  For  the  part  above  the  perturbation  (part  I),  calcite  fabric  is  generally  darker  and  more  compact.  The  δ18O  values  average  -­‐6.6   ±   0.16   ‰   and   δ13C   values   average   -­‐10   ±   0.12   ‰.   Both   display   a   good  correlation  as  indicated  by  a  Spearman’s  correlation  coefficient  of  ρ=  0.811  (p=  8.86   x   10-­‐44).   Layer   thickness   in   part   I   averages   0.3   mm/layer   and   displays  similar  variations  as   the   isotopes  with   thicker   layers   corresponding  with  more  negative   isotopic   values.   The   parts   below   the   perturbation   (parts   II   to   VII)  display  more  negative  δ18O  values  at   -­‐7.0  ±  0.12  ‰  while   the  δ13C  values  vary  around   the   same   mean   of   -­‐10   ±   0.12   ‰.   A   lower   Spearman’s   correlation  coefficient  between  the  δ18O  and  δ13C  signals  is  calculated  for  these  parts  (parts  II  to  VII)  (ρ=  0.37,  p=  9.54  x  10-­‐24).  Below  the  perturbation,  layer  thickness  varies  between  0.5  and  1  mm/layer  and  displays  similar  variations  as  the  δ18O  values.  In  lower  part  II  and  the  upper  part  III  (14  -­‐  18.5  cm)  and  for  the  most  of  part  V,  part  VI  and  VII   (38   -­‐  56  cm),   the  δ18O  signal   is  generally  more  negative   (-­‐7.5  ±  0.16  ‰)  and  the  layer  thickness  increases  to  0.8  mm/layer  (Fig.  3).  In  the  lower  part  III  and  part  IV  (18.5  and  38  cm),  the  δ18O  values  increase  to  -­‐6.6  ±  0.16  ‰  and  the   layer   thickness  decreases   to  0.5  mm/layer,  while  no  general  particular  change  is  observed  for  the  δ13C  values.  Sampling  for  the  stable  isotopes  was  done  layer  per   layer   in  the  parts  II   to  VII  and  reflects  seasonal  variations   in  the  δ18O  and   δ13C   signals.   The   δ13C   seasonality   evolves   differently   from   the   δ18O  seasonality.  A   larger  δ18O-­‐seasonality  of  0.5  ‰  occurs   in   the   lower  part   II   and  upper  part  III  (14  -­‐  18.5  cm)  and  for  the  most  of  part  V,  part  VI  and  VII  (38  -­‐  56  cm),  while   in   lower  part   III   to   IV  (18.5  -­‐  32  cm),   the  δ18O  seasonality   lowers  to  0.25  ‰.  For  δ13C,  the  overall  seasonality  averages  at  0.7  ‰.  An  increase  in  δ13C  seasonality  to  1.5  ‰  occurs  at  32  cm  and  is  followed  by  a  gradual  decrease  until  27  cm  when  the  seasonality  returns  to  0.7  ‰.                

Page 123: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium      

Paper  in  review  in  Climate  of  the  Past   113  

 Figure  2.  The  upper  laminated  56  cm  of  the  Proserpine  core  with  the  description  of   the   calcite   fabric.   The   blue   bars   indicate   intervals   during   which   calcite  deposition  is  disturbed  or  calcite  fabric  is  very  dark  compact  or  white  matte.      

 Figure   3.   The   δ18O   and   δ13C   signals   (‰   VPDB)   and   layer   thickness   of   the  Proserpine   core  plotted   against   distance   from   top.   Blue   bars   indicate   intervals  during   which   the   calcite   fabric,   δ18O   and   δ13C   signals   and   layer   thickness   all  display  simultaneous  large  amplitude  variations.    

Page 124: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium    

Paper  in  review  in  Climate  of  the  Past    

114  

Eight  U/Th-­‐ages  that  were  previously  published  by  some  of  us  (Verheyden  et  al.,  (2006)  are  used  and  numbered  1,  2,  7,  8,  15,  17,  18  and  19,  and  marked  in  light  grey  in  Table  1.  Twelve  new  U/Th  ages  measured  in  this  study  are  listed  in  black  in   Table   1   and   correspond   well   with   the   previously   measured   ages.   Layer  counting  ages  were  carried  out  per  part  (i.e.  part  I  to  part  VII)  and  are  listed  in  Table   2   (column   5)   together  with   their   2σ   uncertainty   range.   To   compare   the  two   independent   age  methods   (layer   counting  method   and  U/Th-­‐age  method),  the   U/Th-­‐age   points   have   to   be   interpolated   to   obtain   an   age   for   the   top   and  bottom  of  each  part.  The   interpolation  of   the  measured  U/Th-­‐ages  was  carried  out   using   StalAge   and   top   and   bottom   ages   of   each   part   are   listed   in   Table   2  (column   3).   The   difference   between   the   top   and   the   bottom   age   of   each   part  provides   the  number   of   years   of   that   part   (Table   2,   column  4).   The   amount   of  years  per  part  derived  from  the  U/Th-­‐ages  display  larger  2σ  uncertainties  for  the  parts  I,  II  and  III  (∼  70)  compared  to  the  parts  IV  to  VII  where  uncertainties  are  smaller   (∼   30).   The   amount   of   years   per   part   derived   from   the   layer   counting  display  2σ  uncertainties  of  ∼  7,  being  smaller  than  the  uncertainties  on  the  U/Th-­‐ages.  The  obtained  amount  of  layers  per  part  correspond  for  the  two  methods  in  the   parts   I,   II,   II,   V   and   VII.   Note   that,   the   U/Th-­‐age   method   suggests   much  smaller  amount  of  years  (Table  2,  columns  4  and  5)  in  the  parts  IV  and  VI.      

 Table   1.   U/Th   measurements   (University   of   Minnesota)   of   the   Proserpine  stalagmite.  All  ages  are  converted  to  before  2013.  Ages  number  1,  2,  7,  8,  15,  17,  18  and  19,  marked  in  light  grey  are  the  U/Th-­‐ages  from  Verheyden  et  al.,  2006.    The   growth   rates   per   part   derived   from   the   U/Th-­‐ages   are   listed   in   Table   2,  column  6.  The   growth   rates  per  part   derived   from   the   layer   counting   ages   are  listed   in  Table  2,   column  7.  The  growth  rates  per  part  based  on   layer  counting  increase  in  two  steps:  they  are  low  at  0.6  mm/y  in  part  I,  higher  around  1  mm/y  in  part  II,  III  and  IV,  and  very  high  at  2  mm/year  in  the  parts  V,  VI,  and  VII.  The  growth  rates  per  part  derived  from  the  U/Th-­‐ages  display  much  larger  variations  between   the  different  parts,  with   exceptionally  high  growth   rates  of   5.6  mm/y  for  the  part  IV  and  of  6.5  mm/y  for  part  VI.    

Page 125: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium      

Paper  in  review  in  Climate  of  the  Past   115  

 Table   2.   Comparison   between   the   layer   counting   ages   and  U/Th-­‐ages   per   part  together  with  their  growth  rates.  The  interpolated  U/Th-­‐ages  for  the  top  and  the  bottom   of   each   part  were   obtained   using   StalAge.   The   interpolated   U/Th   ages  and  amount  of  years  per  part  are  reported  with  their  2σ  uncertainty  range.    5.  Discussion    5.1  Speleothem  Age  model    Two  independent  geochronological  methods  are  used  to  establish  the  age  model  of  the  Proserpine:  StalAge  based  on  20  U/Th-­‐ages  and  layer  counting.  Due  to  the  interruption  in  calcite  deposition  between  9  and  10  cm,  the  layer  counting  ages  cannot  be  used  to  count  the  years  back  from  present  until  56  cm.  To  compare  the  U/Th-­‐ages  and  the  layer  counting  ages,  the  amount  of  years  must  be  determined  for  each  part   (Table  2,   columns  4  and  5).  Results   show   that   the   layer   counting  method   displays   smaller   uncertainties.   Both   independent   geochronological  methods   deliver   similar   ages  with   the   exception   of   parts   IV   and  VI,  where   the  U/Th-­‐ages  suggest  a  lower  number  of  years.  The  U/Th-­‐ages  indicate  that  Part  IV  was  deposited  in  19  ±  33  years  while  the  layer  counting  indicates  a  total  of  105  ±  7  years  (Table  2).  The  U/Th-­‐ages  suggest  that  Part  VI  was  deposited  in  13  ±  27  years  while   the   layer  counting   indicates  a   total  of  42  ±  10  years  (Table  2).  The  number   of   years   obtained   by   layer   counting   in   the   two   parts   IV   and   VI   is  considered  more  probable  compared  to   the  number  of  years  obtained  by  U/Th  ages.  Based  on  in-­‐situ  monitoring  of  the  Proserpine  drip  site  demonstrating  the  seasonal  character  of  the  layering  and  the  good  agreement  of  the  layer  counting  and  the  U/Th  ages  in  most  of  the  other  parts,  the  layer  counting  model  is  seen  as  the  most  accurate  to  establish  the  chronology.  Furthermore,  the  U/Th  ages  give  improbable  high  growth  rates  (∼  6  mm/y)  for  the  parts  IV  and  VI  (Table  2).    Using   the   layer   counting   ages,   the   Proserpine   age  model   is   subdivided   in   two  parts:  the  part  above  the  perturbation  and  the  part  below  the  perturbation.  The  age   of   part   I   above   the   perturbation   can   be   obtained   by   simply   counting   the  layers  back   from  2011.  This   leads   to  an  age  of  1857  ±  6  AD   for   the  end  of   the  perturbation  (Fig.  4).  Below  the  perturbation  (at  10  cm),  the  age  of  the  onset  of  the   perturbation   has   to   be   estimated   in   order   to   restart   the   layer   counting  downwards.   This   is   carried   out   by   counting   the   layers   back   upward   from   the  U/Th-­‐age  located  closest  below  the  perturbation  (=1798  ±  45  AD).  By  doing  this,  a  total  of  12  ±  2  layer-­‐couplets  are  obtained,  indicating  that  the  age  of  the  onset  of  the  perturbation  is  estimated  at  1810  ±  45  AD  (Fig.  4).  The  good  estimation  of  

Page 126: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium    

Paper  in  review  in  Climate  of  the  Past    

116  

this  age  is  confirmed  by  the  fact  that  StalAge  suggests  an  age  of  1810  ±  48  AD  for  the   onset   of   the   perturbation.   Furthermore,   a   14C-­‐date   on   a   straw   piece  embedded   in   the   perturbed   calcite   indicates   an   age   interval   of   1760   to   1810  (probability  of  95.4  %)   (Verheyden  et   al.,   2006)   also   suggesting   a   similar   time  window   for   the   perturbation.   The   age   of   1810   ±   45   AD   is   consequently  considered  a  good  estimation  of  the  onset  of  the  perturbation.  This  age  is  used  to  restart  layer  counting  downwards.  Since  the  uncertainties  on  the  layer  counting  ages   are   determined   per   part,   the   uncertainty   on   the   age  model   increases   per  older  part  according  to  the  propagation  of  uncertainties  on  a  sum  (Table  3).  The  age  obtained  for  the  bottom  of  the  laminated  part  of  the  Proserpine  stalagmite  at  56  cm  is  1479  ±  48  AD  (Fig.  4).      

 Figure   4.   Age-­‐depth   model   of   the   Proserpine   based   on   layer   counting   ages  reported  with  their  2σ  uncertainty.  The  onset  of  the  perturbation  is  estimated  by  counting   the   layers   back   up   from   the   U/Th-­‐age   located   closest   below   the  perturbation.   U/Th-­‐ages   are   plotted   in   light   grey   in   the   age-­‐depth   graph  with  their   2σ   uncertainty.   Location   of   U/Th   samples   on   the   Proserpine   core   is  indicated  by  the  black  dots.  All  ages  are  reported  in  years  AD.      

Page 127: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium      

Paper  in  review  in  Climate  of  the  Past   117  

 Table  3.  Uncertainties  on  the  counted  layers  per  part  below  the  perturbation  (II  to  VII)  together  with  the  uncertainties  on  the  obtained  ages  (AD)  per  part  using  the  age  of  1810  ±  45  AD  as  starting  point  for  the  age  model  counting.    5.2   Factors   driving   decadal   and   multi-­‐decadal   changes   in   the   measured  proxies      Variations  in  δ18O  values  of  speleothems  deposited  in  equilibrium  with  their  drip  water  relate  mainly  to  changes  in  air  temperature,  rainfall  amount  and/or  source  of  the  rainfall  (Fairchild  et  al.,  2006).  Rainfall  sources  often  imply  δ18O  shifts   in  the  order  of  several  ‰  (Fleitmann  et  al.,  2007)  while  the  δ13C  values  and  layer  thickness   values   remain   unchanged.   The   large-­‐scale   δ18O   variations   in   the  Proserpine  are   in  the  order  of  1  to  2  ‰  and  always  occur  together  with   large-­‐scale   δ13C   variations   of   the   same   order   and   a   decrease   in   layer   thickness  indicating  that  the  source  effect  is  most  probably  not  responsible  for  these  δ18O  variations.  In  temperate  regions  speleothem  δ18O  values  often  display  a  difficult  link  with  surface  air  temperature  due  to  the  inverse  effect  of  temperature  on  the  rainwater   δ18O   compared   to   the   calcite   δ18O.   The   relation   between   surface   air  temperature   and   rainwater   δ18O   varies   between   ∼   0.1   and   0.3   ‰/1   °C   for  Central  Europe  (Schmidt  et  al.,  2007).  The  temperature  dependent  fractionation  during   calcite   formation  within   the   cave   acts   in   the   opposite   direction,   and   is  around   -­‐0.2  ‰/1   °C   for   the   Proserpine   drip   site   as   suggested   by   monitoring  results  (Van  Rampelbergh  et  al.,  2014).  The  net  effect  of  air  temperature  changes  on   the   Proserpine   δ18O   signal  may   thus   vary   between  ∼   -­‐0.1   and   0.1  ‰/1   °C  considering  that   the  temperature  dependence  of   the  rainwater  of  ∼  0.1  and  0.3  ‰/1   °C   is   also   valid   for   Belgium.   Consequently,   the   temperature   effect   most  probably  only  has  a  minor  influence  on  the  decadal  and  multi-­‐decadal  variations  in   the   Proserpine   δ18O   signal.   In   the   studied   region,   heavier   δ18O   values   have  been   observed   to   correspond   to   drier   periods   and   thus   reflecting   the   amount  effect   (Verheyden,   2001).   Variations   in   the   Proserpine   δ18O  may   thus   possibly  relate  to  changes  in  wetter  or  drier  conditions.    If   recharge   is   seasonally   biased,   the   decadal   and  multi-­‐decadal  δ18O   variations  may  be  caused  by  variations  in  air  temperature  and/or  by  rainfall  amount  during  a   certain   season.  Hydrological   studies   of   the  Han-­‐sur-­‐Lesse   epikarst   show   that  recharge  mostly  occurs  between  spring  and  fall  with  largest  amounts  of  recharge  

Page 128: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium    

Paper  in  review  in  Climate  of  the  Past    

118  

in  winter   (Bonniver,   2011).  Rainfall  δ18O  data   shows   that  winter   rainfall   has   a  lower   isotopic   composition   compared   to   the   rainfall   from   other   seasons   (Van  Rampelbergh   et   al.,   2014).   During   a   period   of   lower   winter   recharge,   less  isotopically  light  (winter)  water  is  added  to  the  epikarst  reservoir  compared  to  the  heavier  spring  and  fall  water  and  the  total  δ18O  composition  of  the  epikarst  water   increases,   causing   increased   δ18O   values   in   the   speleothem.   Periods   of  increased   δ18O   values   in   the   Proserpine   record   may   thus   be   reflecting   drier  winter  periods  and  vice  versa.  The  relation  between  lower  drip  water  δ18O  and  higher  winter  recharge  amounts  can  be  illustrated  by  drip  water  monitoring  data  over   several   years.   Although   no   such   data   is   available,   winter   recharge   is  considered  the  main  factor  determining  the  δ18O  values  of  the  Proserpine.  More  positive   δ18O   values   are   interpreted   to   reflect   drier   winter   periods   and   vice  versa.   Furthermore,   a   good   Spearman   correlation   can   be   established   between  lower   winter   precipitation   intensities   (DJF)   and   lower   winter   temperatures  (DJF)  measured  by  the  RMI  since  1833  (ρ  =  0.47  and  p  =  3.99  x  10-­‐11)  suggesting  that  drier  winters  correspond  to  colder  winters.  More  negative  δ18O  values  in  the  Proserpine   may   thus   possibly   reflect   drier   winter   conditions   that   are   most  probably   also   colder.   A   similar   interpretation   is   used   for   the   decadal   and  centennial  δ18O  variations  measured  of  a  German  speleothem  with  similar  yearly  temperature   and   yearly   precipitation   amounts   as   the   Proserpine   growth   site  (Wackerbarth  et  al.,  2010;  Fohlmeister  et  al.,  2012).      Since   no  major   vegetation   changes   (mainly   C3-­‐vegetation)   occurred   above   the  cave  for  studied  period  and  site,  changes  in  δ13C  values  might  relate  to  changes  in  soil   activity   (Genty   et   al.,   2003;   Fohlmeister   et   al.,   2012)   and/or   Prior   Calcite  Precipitation   (PCP)   (Fairchild   et   al.,   2000).   Plant-­‐CO2   has   a   lower   isotopic  signature   compared   to   atmospheric   CO2   (δ13C   of   C3-­‐vegetation   is   between   -­‐20  and  -­‐25‰,  while  in  atmospheric  CO2  it  evolved  from  –7  ‰  to  –8  ‰  during  the  studied  period).  A  reduced  plant-­‐CO2   input   in  the  soil  due  to   lower  soil  activity  will   increase   the  δ13C   of   the   soil-­‐CO2   reservoir   and   consequently   the   dissolved  inorganic   carbon   (DIC)   in   the   epikarst   water.   During   PCP,   calcite   is   deposited  from  the  percolating  epikarst  water  before  entering  the  cave  as  drip  water.  This  process  mostly   occurs   during   drier   periods  when   aerated   zones   become  more  important  in  the  epikarst.  PCP  causes  a  simultaneous  increase  in  the  δ13C  and  in  the   Mg/Ca   and   Sr/Ca   composition   of   the   drip   water   and   speleothem   calcite  (Fairchild  et  al.,  2000).  Although  no  Mg/Ca  and  Sr/Ca  ratios  are  measured  in  the  Proserpine,  which  makes   it  difficult   to  evaluate   the  process  of  PCP,  monitoring  results  have  clearly  demonstrated  that  PCP  is  an  important  process  in  the  Han-­‐sur-­‐Lesse   epikarst   (Van   Rampelbergh   et   al.,   2014).   Both   effects,   being   soil  activity   and   PCP   act   in   the   same   direction   and   both   cause   the   δ13C   values   to  increase  during  drier  periods.  Since  drier  periods  in  the  cave  are  caused  by  lower  winter   recharge   periods,   increased   δ13C   values   are   interpreted   to   reflect   drier  and  most  probably  also  colder  winter  periods.    Disequilibrium  processes  due  to  a  stronger  pCO2  gradient  between  the  cave  air  and   drip  water   and/or   due   to   longer   drip   intervals  may   cause   simultaneously  increased   δ18O   and   δ13C   values   (Mühlinghaus   et   al.,   2009;   Scholz   et   al.,   2009;  Deininger  et  al.,  2012).  Under  the  present-­‐day  conditions,  pCO2  levels  of  the  cave  

Page 129: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium      

Paper  in  review  in  Climate  of  the  Past   119  

air   in   the   Salle-­‐du-­‐Dôme   are   low   year-­‐round   and   equal   the   outside   air   values.  pCO2  levels  may  change  over  time  due  to  changes  in  ventilation  patterns,  which  may  change  over  time  due  to  new  cave  openings.  No  such  new  openings  that  may  have  affected  the  Salle-­‐du-­‐Dôme  ventilation  occurred  in  the   last  500  years.  The  effect  of  changing  pCO2  gradient  on  the  drip  water  δ18O  and  δ13C  values  over  the  studied  period  is  thus  unlikely.  Longer  drip  intervals  due  to  decreased  drip  flow  may  be  possible.  However,  under  the  present-­‐day  conditions,  a  continuous  drip  water  flow  feeds  the  stalagmite,  which  inhibits  disequilibrium  effects  related  to  longer  drip  interval  (Mühlinghaus  et  al.,  2009).  The  drip  discharge  consequently  needs   to   be   sufficiently   decreased,   beneath   a   certain   threshold   value,   to   allow  disequilibrium   processes   to   be   present.   Since   recharge   occurs   in   winter  (Bonniver,   2011),   a   decreased   drip   discharge   is   expected   to   relate   with  significantly  drier  winters,   that  are  also  colder.  Furthermore,  during  periods  of  lower   drip   discharge,   PCP   will   occur   and   further   increase   the   δ13C   signal.  Decreased   drip   discharge   due   to   significantly   drier   (and   colder)   winters   will  consequently   cause   increased   correlating   δ18O   and   δ13C   values   with   a   larger  increase  in  δ13C  values  compared  to  the  δ18O  values,  the  latter  being  not  affected  by  PCP.      Layer   thickness   and   calcite   fabric   in   the   Proserpine   are   expected   to   relate   to  growth  rate,  with  thinner  layers  and  darker  calcite  formed  under  slower  growth.  Growth  rate  is  primarily  dependent  on  two  factors;  the  discharge  amount,  which  is   expected   to   lower   during   drier   (and   colder)   winter   periods   and   the   cave  seepage  water  calcium  ion  concentration  (Genty  et  al.,  2001).  The  cave  seepage  water  calcium  ion  concentration  depends  on  mainly  two  factors.  The  first  factor,  being  the  soil  pCO2  is  expected  to  increases  during  warmer  and  wetter  periods.  Higher  soil  pCO2   increase   the  amount  of  CO2  dissolved   in   the  soil  water.  Water  containing  higher  CO2  amounts  more  easily  dissolves  CaCO3,  which  increases  its  calcium  ion  concentration.  The  second  factor  determining  seepage  water  calcium  ion  concentration  is  the  intensity  of  PCP.  PCP  mostly  occurs  during  dry  periods  and  decreases   the   Ca2+   concentration   of   the  drip  water   due   to   precipitation   of  calcite   in   the   epikarst.   Cave  monitoring   results   show   that   PCP   is   an   important  process   in   the   Han-­‐sur-­‐Lesse   epikarst   that   becomes   more   intense   during   the  drier  summer  season  (Van  Rampelbergh  et  al.,  2014).  During  drier  periods,  most  probably  cased  by  drier   (and  colder)  winter  periods,   soil   activity  will  decrease  and  PCP  will  increase,  both  causing  lower  calcium  ion  concentration  of  the  drip  water.   A   lower   calcium   ion   concentration   and   a   lower   drip   discharge   during  drier   (and   colder)   winters   will   both   cause   slower   growth   of   the   calcite   and  consequently  thinner  layers  and  darker  calcite.      To   conclude,   decadal   and   centennial   changes   in   the   proxies   (δ18O   and   δ13C  signals,   layer   thickness   and   calcite   color)   reflect   changes   in   drier   (and   colder)  versus  wetter  (and  warmer)  winters.  Exceptionally  dry  (and  cold)  winters  shift  the  drip  discharge  below  a  certain  threshold  value,  which  causes  the  proxies  to  display  simultaneous  large  amplitude  shifts.  During  such  exceptionally  dry  (and  cold)   winter   periods,   the   δ18O   and   δ13C   values   increase,   layer   thickness   will  decrease   and   calcite   fabric   will   become   darker   and/or   disturbed.   When   the  

Page 130: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium    

Paper  in  review  in  Climate  of  the  Past    

120  

discharge  threshold  is  not  reached,  calcite  is  deposited  close  to  equilibrium  and  the  four  proxies  may  vary  independently.      5.3  Anomalies  in  the  proxy  records    Proserpine  calcite  deposited  in  equilibrium  with  its  drip  water  has  a  δ18O  value  of   -­‐6.7   ±   0.16  ‰   and   a   δ13C   value   of   -­‐10   ±   0.12  ‰   (Van   Rampelbergh   et   al.,  2014).  Four  periods  where  the  δ18O  and  δ13C  values  abruptly  increase  away  from  the  present-­‐day  equilibrium  occur  in  the  Proserpine  from  1565  to  1610,  at  1730,  from  1770  to  1800  and  from  1880  to  1895  and  are  interpreted  as  anomalies  in  the  record  (blue  bars  Fig.  5).  During  these  anomalies   layer   thickness  decreases  below  0.2  mm/layer  and  calcite  fabric  is  disturbed  or  very  dark  and  compact.  As  indicated   by   the   detailed   analysis   of   the   climatic   factors   affecting   the   different  used   proxies,   as   soon   as   a   certain   threshold   value   is   reached,   the   four   proxies  display  simultaneous  large-­‐amplitude  changes  and  reflect  exceptionally  dry  (and  cold)  winter  periods.  No  calcite  was  deposited  between  1810  and  1860,  which  strongly  suggests  that  too  little  water  was  dripping  on  the  Proserpine  during  that  period.   Therefore,   this   period   is   also   interpreted   as   an   anomaly   reflecting  exceptionally  dry  (and  cold)  winters.  A  total  of  five  anomalies  are  suggested  by  the  Proserpine  proxies  and  last  between  1565  and  1610,  at  1730,  between  1770  and  1800,  between  1810  and  1860  and  between  1880  and  1895  (blue  bars  Fig.  5).  The  five  anomalies  suggesting  exceptionally  dry  (and  cold)  winter  conditions  correspond  with  known  cold  and/or  dry  periods   in  historical  and   instrumental  archives   and   in   winter   temperature   reconstructions   from   Europe   and   Central  Europe  (Fig.  5):          

• Between  1565  and  1610  winter  temperatures  in  Europe  (Luterbacher  et  al.,  2004)  and  Central  Europe  (Dobrovolny  et  al.,  2010)  were  low  (Fig.  5,  f  and  g).  Historical  data  of  France,  Belgium  and  the  Netherlands  indicate  icy  cold  winters,   harsh   famines,   low  numbers  of   child  births   and  weddings,  and   the  outbreak  of   the  plague  with   its  worst   years   from  1562   to  1570  (Le  Roy  Ladurie,  2004).  The  shift  to  cold  and  dry  conditions  at  1565  AD  is  interpreted   as   the   onset   of   the   second   pulse   of   the   Little   Ice   Age   (LIA,  ±1300-­‐1850)   (Le   Roy   Ladurie,   2004)   and   is   nicely   recorded   in   the  Proserpine  proxies  as  a  shift  to  drier  (and  colder)  winters.  Between  1590  and  1600,  the  Proserpine  proxies  suggest  a  shorter  wetter  (and  warmer)  interval   as   indicated   by   the   more   negative   δ18O   and   δ13C   values   and  thicker  layers  (Fig.  5  a,  b  and  c).    A  similar  decade  of  warmer  conditions  between   1590   and   1600   is   also   reported   in   winter   temperature  reconstructions   from   Europe   (Luterbacher   et   al.,   2004),   Central   Europe  (Dobrovolny   et   al.,   2010)   and   from   historical   archives   (Le   Roy   Ladurie,  2004).    

• At   1730,   the   abrupt   shift   in   the  measured   proxies   suggests   a   short   but  exceptionally   dry   (and   cold)   winter   period.   Considering   the   age  uncertainty   of   ±   45   years   for   this   period   (Fig.   5),   the   dry   (and   cold)  conditions  suggested  by  the  Proserpine  at  1730  ±  45  AD,  most  probably  relate   to   the  exceptionally  cold  and  dry  decade  between  1690  and  1700  AD   recorded   in   historical   archives   (Le   Roy   Ladurie,   2004)   and   by  

Page 131: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium      

Paper  in  review  in  Climate  of  the  Past   121  

extremely   low  winter  temperatures   in  Europe  (Luterbacher  et  al.,  2004)  and  Central  Europe  (Dobrovolny  et  al.,  2010)  (Fig.  5,  f  and  g).    

• Between  1770  and  1800,  the  Proserpine  proxies  suggest  a  dry  (and  cold)  winter   period   that   corresponds   to   a   known   period   of   colder  winters   in  Europe  (Fig.  5,   f  and  g)   (Le  Roy  Ladurie,  2004;  Luterbacher  et  al.,  2004;  Dobrovolny  et  al.,  2010).    

• The   exceptionally   dry   (and   cold)   winter   conditions   between   1810   and  1860,  as  suggested  by  the  Proserpine,  correspond  nicely  with  decreased  winter   temperatures   in   Europe   (Luterbacher   et   al.,   2004)   and   Central  Europe    (Dobrovolny  et  al.,  2010)  (Fig.  5,  f  and  g).  Historical  climate  data  from   France,   Belgium   and   the   Netherlands   indicate   that   this   interval  corresponds   with   the   third   and   last   cold   pulse   of   the   LIA   and   is  characterized  by  exceptionally  cold  winters  and  warm  summers  (Le  Roy  Ladurie,  2004).    

• The  most  recent  dry  (and  cold)  period  recorded  in  the  Proserpine  (1880  and   1895)   corresponds   with   colder   winter   temperatures   and   lower  winter   precipitation   amounts   as  measured  by   the  RMI   in  Belgium   since  1833   (Fig.   5,   d   and   e).   The   temperature   drop   is   clearly   visible   in   the  winter   temperature   reconstruction   from   Europe   (Luterbacher   et   al.,  2004)  (Fig.  5,  f).  A  decrease  in  precipitation  has  also  been  recorded  in  the  England   and  Wales   precipitation   record,  where   this   period   is   known   as  very   dry   with   peak   dry   years   at   1884,   1887   and   1893   (Nicholas   and  Glasspoole,  1931).    

   The  exact  forcing  behind  these  five  dry  (and  cold)  winter  periods  is  still  a  matter  of   discussion.   The  most   trivial   forcing   of   the  western   European   climate   is   the  variation  in  winter  North  Atlantic  Oscillation  (NAO)  (Trouet  et  al.,  2009).  During  a  negative  winter  NAO  phase,  westerlies  winds  are  forced  over  southern  Europe,  which  may  cause  drier  and  colder  winter  conditions  over  Belgium.  However,  the  five   dry   (and   cold)   winter   periods   observed   in   the   Proserpine   do   not   always  correspond  with  negative  winter  NAO  phases   (Trouet   et   al.,   2009).  Other   than  negative  NAO  phases,  lower  solar  irradiance  combined  with  the  input  of  volcanic  gasses  in  the  atmosphere  may  also  be  responsible  for  decreased  temperatures  in  Europe.  Such  is  probably  the  case  for  the  cold  and  dry  period  between  1810  and  1860   (third  pulse  of   the  LIA).   In   this  period,   solar   insolation  decreased  during  the   Dalton   Minimum   (1790-­‐1810,   Mann,   2002)   and   the   Tamborra   volcano  (Indonesia)   erupted   in   1815.   Combination   of   negative   NAO   conditions  (Luterbacher   et   al.,   2001),   the   eruption  of   the  Krakatoa  volcano   (Indonesia)   in  1883   and   lower   sunspot   activity   (Lassen   and   Friischristensen,   1995)   are  most  probably  responsible  for  the  exceptionally  dry  (and  cold)  winter  period  between  1880  and  1895.        

Page 132: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium    

Paper  in  review  in  Climate  of  the  Past    

122  

   Figure   5.   The   (a)   δ18O   and   (b)   δ13C   values   (‰  VPDB)   and   (c)   layer   thickness  plotted   against   (d)   the   instrumental   winter   temperature   (DJF)   and   (e)   winter  precipitation   (DJF)   record   of   the   Belgian   Royal   Meteorological   Institute   (RMI)  measured   in   Brussels   since   1833   (f)   the   winter   temperature   reconstruction  based  on  multiple  proxies  in  Europe  (Luterbacher  et  al.,  2004)  and  (g)  the  winter  temperature   reconstruction   derived   from   documentary   and   instrumental  evidence  in  Central  Europe  (Dobrovolny  et  al.,  2010).  Five  exceptionally  dry  (and  cold)  winter  periods  suggested  by  the  Proserpine  are  indicated  by  blue  bars  and  correspond   with   clear   cold   periods   in   instrumental   records   and   winter  temperature   reconstructions   in   Europe   and   Central   Europe.   Two   periods   of  relatively   wetter   (and   warmer)   winters   occur   from   1479   and   1565   and   from  1730  to  1770  and  corresponds  with  known  warmer  intervals.  Between  1610  and  1730  the  Proserpine  suggests  relatively  drier  (and  colder)  winter  periods,  which  correspond  with  colder  winter  conditions  in  Europe  and  Central  Europe.        

Page 133: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium      

Paper  in  review  in  Climate  of  the  Past   123  

5.4  More  relative  warmer  and  wetter  and  colder  and  drier  periods    In   contrast   to   the   five   anomaly   periods   where   the   four   proxies   suggest  exceptionally   dry   (and   cold)   winter   conditions,   the   remaining   parts   of   the  Proserpine  stalagmite  display  more  relative  variations.  Between  2001  and  1860,  above   the   perturbation,   the   δ18O   and   δ13C   values   display   a   bulge   with   most  negative   values   around  1930.   Layer   thickness   follows   the   same  evolution  with  the  thickest  layers  around  1930  indicating  an  evolution  to  wetter  (and  warmer)  winters   up   to   1930   followed   by   an   evolution   to   drier   (and   colder)   winters   to  2001.   This   observation   in   the   Proserpine   proxies   does   not   correspond   with  instrumental  winter   precipitation   and   temperature   data  measured   by   the   RMI  since  1833  nor  with  European  winter  temperature  reconstructions  (Luterbacher  et  al.,  2004)  (Fig.  5).  Calcite  is  darker  in  this  part  due  to  the  incorporation  of  soot  from  torches  used  to  illuminate  the  chamber  during  cave  visits  (Verheyden  et  al.,  2006).   Soot   incorporation   in   the   calcite   structure   may   hamper   the   calcite  deposition   and   overprint   lower-­‐amplitude   climate   variations.   However,   large-­‐amplitude  variations  such  as   the  dry  (and  cold)  winter  anomaly  between  1880  and  1895  are  still  visible  within  this  part,  indicating  that  the  climate  signal  is  not  fully  overprinted.  The  possible  effects  of  soot  on  δ18O  and  δ13C  values  and  layer  thickness   need   further   investigation   to   allow   deriving   low-­‐amplitude   climate  variations  in  the  part  above  the  perturbation.      Below   the   perturbation,   and   with   exception   of   the   anomaly   periods,   the  measured  proxy  signals  can  be  subdivided   in  three  periods;  between  1479  and  1565,  between  1610  and  1730  and  between  1730  and  1770   (Fig  5  a,  b  and  c).  Between  1479  and  1565  and  between  1730  and  1770,  more  negative  δ18O  values  and  thicker  layers  indicate  relatively  wetter  (and  warmer)  winter  conditions.  In  between   the   two   latter   periods   (1610-­‐1730),   the   δ18O   values   become   less  negative   and   layers   become   thinner   indicating   relatively   drier   (and   cooler)  winter  conditions.  During  the  three  above  described  periods  (1479-­‐1565,  1610-­‐1730,   1730-­‐1770),   the   δ13C   values   display   no   variations   indicating   no   major  changes   in   soil   activity   or   PCP   intensity.   Only   during   the   relatively   drier   (and  colder)  winter  period  between  1610  and  1730,   the  δ13C  values  display   a  weak  gradual  increase  from  1700  to  1730.  The  relatively  dry  (and  cool)  conditions  in  the   period   between  1610   and  1730  may  have   caused   lower   soil   activity   and   a  gradual  increase  in  prior  calcite  precipitation,  which  gradually  augment  the  δ13C  signal.      The   two  periods  with   relatively  wetter   (and  warmer)  winters   (1479-­‐1565  and  1730-­‐1770)   interrupted   by   a   period   with   drier   (and   cooler)   winters   (1610-­‐1730)  observed  in  the  Proserpine  are  also  recorded  in  the  winter  temperatures  reconstructions   of   Europe   (Luterbacher   et   al.,   2004)   and   Central   Europe  (Dobrovolny   et   al.,   2010)(Fig.   5)   and   in   historical   archives   (Le   Roy   Ladurie,  2004).  The  relatively  drier  (and  cooler)  winter  period  between  1610  and  1730  corresponds   to   colder  winter   conditions   in   Europe   and   Central   Europe   and   is  referred  as   the   second  pulse  of   the  LIA   (Le  Roy  Ladurie,  2004).  This   relatively  cooler  interval  may  relate  to  the  Maunder  Minimum,  being  a  period  of  decreased  solar   activity   between   1640   and   1714.   However,   lower   solar   irradiance   alone  

Page 134: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium    

Paper  in  review  in  Climate  of  the  Past    

124  

cannot   be   responsible   for   the   cooler   conditions   between   1610   and   1730.   The  exact  forcing  of  this  second  pulse  of  the  LIA  is  still  a  matter  of  discussions.      5.5  Seasonality  in  δ18O  and  δ13C  values    The  δ18O  and  δ13C  values  were  measure  at   a   seasonal   scale  between  1479  and  1810  and  clearly  display  seasonal  variations  (Fig.  6).   Interpretation  of   the  δ18O  and  δ13C  variations  on  a  seasonal  scale  strongly  differs  from  the  interpretation  of  these   proxies   on   decadal   and   multi-­‐decadal   scale.   Whereas   the   decadal   and  multi-­‐decadal   variations   in  δ18O   and  δ13C   vary   in   phase   and   reflect   changes   in  drier   (and   colder)   versus  wetter   (and  warmer)  winters,   the   seasonal  δ18O   and  δ13C  values  vary   in  anti-­‐phase.   Seasonal  δ18O  variations  are  driven  by   seasonal  cave   air   temperature   changes  with   a   temperature  dependence   of   -­‐0.2  ‰/1   °C  (Van  Rampelbergh  et  al.,  2014).  Higher  cave  air  temperatures  in  summer  lead  to  lower  δ18O  values  of  the  formed  calcite.  The  seasonal  variation  in  δ13C  values  is  driven   by   the   seasonal   change   in   PCP   intensity,  with   stronger   PCP   in   summer  leading  to  increased  calcite  δ13C  values  (Van  Rampelbergh  et  al.,  2014).      The   seasonality   in  δ18O  measured  during   the   two  wetter   (and  warmer)  winter  periods   (1479-­‐1565   and   1730-­‐1770),   equals   0.5   ‰,   which   is   similar   to   the  present-­‐day  conditions  (Van  Rampelbergh  et  al.,  2014)  and  corresponds  with  a  2  to  2.5  °C  seasonality   in  cave  air   temperature.  Between  1610  and  1730,  winters  are   relatively   drier   (and   cooler),   and   the   δ18O   seasonality   lowers   to   0.25  ‰  corresponding   with   a   1   to   1.5   °C   cave   air   temperature   seasonality.   Lower  summer  temperatures  during  this  cold  LIA  period  are  most  probably  responsible  for  the  lower  cave  air  seasonality.      The  δ13C  signal  mostly  displays  a  seasonality  of  0.7  ‰  being  smaller  than  the  1  ‰  seasonality   in  δ13C   values   observed  under   the   present-­‐day   conditions   (Van  Rampelbergh  et  al.,  2014).  At  1600,  the  δ13C  seasonality  increases  to  1.5  ‰  and  displays  a  gradual  decreasing  trend  back  to  0.7  ‰  at  1660.  The  increase  in  δ13C  seasonality   between   1600   and   1660   also   corresponds   with   an   interval   where  layers   are   thinner   (∼   0.4   mm/layer)   but   clearly   alternating   between   dark  compact   and   white   porous   layers.   This   suggests   well-­‐expressed   wet   winter  conditions   and   dry   summer   conditions   in   the   cave.   The   relatively   drier   (and  colder)  winter  conditions  in  the  period  between  1610  and  1730  cause  the  yearly  water   recharge   (occurring  mostly   in  winter)   to   be   lower   compared   to   the   two  periods  with  wetter  (and  warmer)  winters  (1479-­‐1565  and  1730-­‐1770).  A  lower  recharge   during   winter   will   consequently   lead   to   drier   cave   conditions   in  summer,  and  increase  the  effect  of  PCP.  Increased  PCP  in  summer  due  to  lower  winter   recharge   is   interpreted   to   be   responsible   for   the   increased   δ13C  seasonality  and  the  clear  layering  between  1600  and  1660.  

Page 135: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium      

Paper  in  review  in  Climate  of  the  Past   125  

 Figure  6.  A  decrease  in  δ18O  seasonality  in  the  drier  (and  cooler)  period  between  1610   and   1730   (blue)   indicates   lower   cave   air   temperature   seasonality   than  during   the   wetter   (and   warmer)   periods   (1479-­‐1565   and   1730-­‐1770)   (red).  Seasonality   in   the   δ13C   signal   is   higher   between   1600   and   1660   and   indicates  more  intense  PCP  during  summer  (green),  which  is  caused  by  decreased  winter  recharge.    All  values  are  in  ‰  VPDB.    6.  Conclusions    

1. A  multiproxy   approach   using   δ18O   and   δ13C   values,   layer   thickness   and  calcite   fabric   of   the   Proserpine   stalagmite   from   the   Han-­‐sur-­‐Less   cave,  Belgium,  successfully  reconstructs   the  climate  over  the   last  522  years   in  terms  of  drier  (and  colder)  versus  wetter  (and  warmer)  winters.  

2. Thinner   layers  and  darker  calcite  correspond   to  periods  with  decreased  growth   rate,   driven   by   lower   recharge   and   stronger   PCP   effects   during  drier   (and   colder)  winters.  Higher  δ18O  values   are   interpreted   to   reflect  drier  (and  colder)  winters,  due  to  the  decreased  input  of  winter  recharge  water   with   light   isotopic   composition.   Higher   δ13C   values   reflect   lower  soil  activity  and  increased  PCP  during  drier  (and  colder)  winter  periods.    

3. Anomalies  in  the  measured  proxies  occur  when  discharge  drops  under  a  certain  threshold  value.  During  these  anomalies,  the  δ18O  and  δ13C  values  increase   away   from   isotopic   equilibrium,   layers   become   thin   and   the  calcite  becomes  very  dark  or  disturbed.  Such  periods  occur  between  1565  and   1610,   around   1730,   between   1770   and   1800,   between   1810   and  1860   and   between   1880   and   1895   and   are   interpreted   as   reflecting  exceptionally   dry   (and   cold)   winter   conditions.   The   exceptionally   dry  (and  cold)  periods   found   in   the  Proserpine   speleothem  correspond  well  with   known   dry   and   cold   periods   in   historical,   instrumental   and/or  temperature  reconstruction  records  from  Europe.  

4. Less   exceptional   variations  occur  between  1479  and  1565  and  between  1730   and   1770,   with   more   negative   δ18O   values   and   thicker   layers  reflecting  two  relatively  wetter  (and  warmer)  winters.  Less  negative  δ18O  values,  still  reflecting  equilibrium  conditions,  and  thinner  layers  between  1610  and  1730  are  interpreted  to  reflect  a  period  of  relatively  drier  (and  cooler)  winters.  The   two  relatively  wetter   (and  warmer)  winter  periods  correspond   with   warmer   periods   in   European   winter   temperature  reconstructions   and   historical   data   from   Belgium,   the   Netherland   and  France.   The   drier   (and   cooler)   winter   period   between   1610   and   1730  

Page 136: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium    

Paper  in  review  in  Climate  of  the  Past    

126  

corresponds   with   relatively   colder   conditions   in   winter   temperature  reconstructions  and  historical  data.  

5. Seasonally  resolved  isotopic  signals  successfully  record  seasonal  changes  in   cave   air   temperature   and  PCP.  The  δ18O   signals   suggest   a  2   to  2.5   °C  cave   air   temperature   seasonality   between   1479   and   1565   and   between  1730  and  1770,  which  is  similar  to  the  seasonality  in  cave  air  temperature  observed   today.   Between   1610   and   1730,   corresponding   with   a   period  with  drier   (and   cooler)  winters,   the   seasonality   in   cave  air   temperature  decreases   to   1   to   1.5°C.   The   δ13C   seasonal   changes   suggest   that   the  seasonality   in   discharge   was   lower   than   observed   today   with   a   short  interval   of   increased   seasonality   between   1600   and   1660   reflecting  stronger  summer  PCP-­‐effects  due  to  decreased  winter  recharge.  

 7.  Acknowledgements  We   thank   the   ‘Domaine   des   Grottes   de   Han’   for   allowing   us   to   sample   the  stalagmite  and  to  carry  out  other  fieldwork,  Mr  Van  Dierendonck  for  his  interest  and   support.   PC   thanks   the   Hercules   Foundation   and   Research   Foundation    Flanders  (FWO,  project  G-­‐0422-­‐10).    8.  References    Baker,  A.,  Wilson,  R.,  Fairchild,  I.  J.,  Franke,  J.,  Spoetl,  C.,  Mattey,  D.,  Trouet,  V.,  and  Fuller,  L.:  High  resolution  delta  O-­‐18  and  delta  C-­‐13  records  from  an  annually  laminated  Scottish  stalagmite  and  relationship  with  last  millennium  climate,  Glob.  Planet.  Change,  79,  303-­‐311,  2011.  

Bonniver,  I.:  Etude  Hyrogeologique  et  dimmensionnement  par  modelisation  du  "systeme-­‐tracage"  du  reseau  karstique  the  Han-­‐sur-­‐Lesse  (Massif  de  Boine,  Belgique),  2011.  Geologie,  FUNDP  Namur,  Namur,  p  93  to  97  pp.,  2011.  

Cheng,  H.,  Edwards,  R.  L.,  Broecker,  W.  S.,  Denton,  G.  H.,  Kong,  X.,  Wang,  Y.,  Zhang,  R.,  and  Wang,  X.:  Ice  Age  Terminations,  Science,  326,  248-­‐252,  2009a.  

Cheng,  H.,  Edwards,  R.  L.,  Hoff,  J.,  Gallup,  C.  D.,  Richards,  D.  A.,  and  Asmerom,  Y.:  The  half-­‐lives  of  uranium-­‐234  and  thorium-­‐230,  Chemical  Geology,  169,  17-­‐33,  2000.  

Cheng,  H.,  Fleitmann,  D.,  Edwards,  R.  L.,  Wang,  X.  F.,  Cruz,  F.  W.,  Auler,  A.  S.,  Mangini,  A.,  Wang,  Y.  J.,  Kong,  X.  G.,  Burns,  S.  J.,  and  Matter,  A.:  Timing  and  structure  of  the  8.2  kyr  BP  event  inferred  from  delta  O-­‐18  records  of  stalagmites  from  China,  Oman,  and  Brazil,  Geology,  37,  1007-­‐1010,  2009b.  

Deininger,  M.,  Fohlmeister,  J.,  Scholz,  D.,  and  Mangini,  A.:  Isotope  disequilibrium  effects:  The  influence  of  evaporation  and  ventilation  effects  on  the  carbon  and  oxygen  isotope  composition  of  speleothems  -­‐  A  model  approach,  Geochimica  Et  Cosmochimica  Acta,  96,  57-­‐79,  2012.  

Dobrovolny,  P.,  Moberg,  A.,  Brazdil,  R.,  Pfister,  C.,  Glaser,  R.,  Wilson,  R.,  van  Engelen,  A.,  Limanowka,  D.,  Kiss,  A.,  Halickova,  M.,  Mackova,  J.,  Riemann,  D.,  Luterbacher,  J.,  and  Boehm,  R.:  Monthly,  seasonal  and  annual  temperature  

Page 137: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium      

Paper  in  review  in  Climate  of  the  Past   127  

reconstructions  for  Central  Europe  derived  from  documentary  evidence  and  instrumental  records  since  AD  1500,  Climatic  Change,  101,  69-­‐107,  2010.  

Edwards,  R.  L.,  Chen,  J.  H.,  and  Wasserburg,  G.  J.:  238U-­‐234U-­‐230Th-­‐232Th  systematics  and  the  precise  measurement  of  time  over  the  past  500  000  years,  Earth  and  Planetary  Science  Letters,  81,  175-­‐192,  1987.  

Fairchild,  I.  J.,  Borsato,  A.,  Tooth,  A.  F.,  Frisia,  S.,  Hawkesworth,  C.  J.,  Huang,  Y.  M.,  McDermott,  F.,  and  Spiro,  B.:  Controls  on  trace  element  (Sr-­‐Mg)  compositions  of  carbonate  cave  waters:  implications  for  speleothem  climatic  records,  Chemical  Geology,  166,  255-­‐269,  2000.  

Fairchild,  I.  J.,  Smith,  C.  L.,  Baker,  A.,  Fuller,  L.,  Spötl,  C.,  Mattey,  D.,  McDermott,  F.,  and  Eimp:  Modification  and  preservation  of  environmental  signals  in  speleothems,  Earth-­‐Science  Reviews,  75,  105-­‐153,  2006.  

Fleitmann,  D.,  Burns,  S.  J.,  Mangini,  A.,  Mudelsee,  M.,  Kramers,  J.,  Villa,  I.,  Neff,  U.,  Al-­‐Subbary,  A.  A.,  Buettner,  A.,  Hippler,  D.,  and  Matter,  A.:  Holocene  ITCZ  and  Indian  monsoon  dynamics  recorded  in  stalagmites  from  Oman  and  Yemen  (Socotra),  Quaternary  Science  Reviews,  26,  170-­‐188,  2007.  

Fohlmeister,  J.,  Schroder-­‐Ritzrau,  A.,  Scholz,  D.,  Spötl,  C.,  Riechelmann,  D.  F.  C.,  Mudelsee,  M.,  Wackerbarth,  A.,  Gerdes,  A.,  Riechelmann,  S.,  Immenhauser,  A.,  Richter,  D.  K.,  and  Mangini,  A.:  Bunker  Cave  stalagmites:  an  archive  for  central  European  Holocene  climate  variability,  Climate  of  the  Past,  8,  1751-­‐1764,  2012.  

Frisia,  S.,  Borsato,  A.,  Preto,  N.,  and  McDermott,  F.:  Late  Holocene  annual  growth  in  three  Alpine  stalagmites  records  the  influence  of  solar  activity  and  the  North  Atlantic  Oscillation  on  winter  climate,  Earth  and  Planetary  Science  Letters,  216,  411-­‐424,  2003.  

Genty,  D.,  Baker,  A.,  and  Vokal,  B.:  Intra-­‐  and  inter-­‐annual  growth  rate  of  modern  stalagmites,  Chemical  Geology,  176,  191-­‐212,  2001.  

Genty,  D.,  Blamart,  D.,  Ouahdi,  R.,  Gilmour,  M.,  Baker,  A.,  Jouzel,  J.,  and  Van-­‐Exter,  S.:  Precise  dating  of  Dansgaard-­‐Oeschger  climate  oscillations  in  western  Europe  from  stalagmite  data,  Nature,  421,  833-­‐837,  2003.  

Genty,  D.  and  Quinif,  Y.:  Annually  laminated  sequences  in  the  internal  structure  of  some  Belgian  stalagmites  -­‐  Importance  for  paleoclimatology,  Journal  of  Sedimentary  Research,  66,  275-­‐288,  1996.  

Kottek,  M.,  Grieser,  J.,  Beck,  C.,  Rudolf,  B.,  and  Rubel,  F.:  World  map  of  the  Koppen-­‐Geiger  climate  classification  updated,  Meteorologische  Zeitschrift,  15,  259-­‐263,  2006.  

Lassen,  K.  and  Friischristensen,  E.:  Variability  of  Solar-­‐Cycle  length  during  the  past  5  centuries  and  the  apparent  association  with  terrestrial  climate  Journal  of  Atmospheric  and  Terrestrial  Physics,  57,  835-­‐845,  1995.  

Page 138: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium    

Paper  in  review  in  Climate  of  the  Past    

128  

Le  Roy  Ladurie,  E.:  Histoire  humaine  et  comparée  du  climat,  Canucules  et  Glaciers  XIII  et  XVIII  sciècles,  2004.  

Luterbacher,  J.,  Dietrich,  D.,  Xoplaki,  E.,  Grosjean,  M.,  and  Wanner,  H.:  European  seasonal  and  annual  temperature  variability,  trends,  and  extremes  since  1500,  Science,  303,  1499-­‐1503,  2004.  

Luterbacher,  J.,  Xoplaki,  E.,  Dietrich,  D.,  Jones,  P.  D.,  Davies,  T.  D.,  Portis,  D.,  Gonzalez-­‐Rouco,  J.  F.,  von  Storch,  H.,  Gyalistras,  D.,  Casty,  C.,  and  Wanner,  H.:  Extending  North  Atlantic  Oscillation  reconstructions  back  to  1500,  Atmospheric  Science  Letters,  2,  114-­‐124,  2001.  

Mangini,  A.,  Spötl,  C.,  and  Verdes,  P.:  Reconstruction  of  temperature  in  the  Central  Alps  during  the  past  2000  yr  from  a  delta  O-­‐18  stalagmite  record,  Earth  and  Planetary  Science  Letters,  235,  741-­‐751,  2005.  

Mann,  M.  E.:  The  Earth  system:  physical  and  chamical  dimensions  of  global  environmental  change.  In:  Encyclopedia  of  Global  Environmental  Change,  MacCracken,  M.  C.  and  Perry,  J.  S.  (Eds.),  John  Wiley  &  Sons,  Chichester,  2002.  

Mattey,  D.,  Lowry,  D.,  Duffet,  J.,  Fisher,  R.,  Hodge,  E.,  and  Frisia,  S.:  A  53  year  seasonally  resolved  oxygen  and  carbon  isotope  record  from  a  modem  Gibraltar  speleothem:  Reconstructed  drip  water  and  relationship  to  local  precipitation,  Earth  and  Planetary  Science  Letters,  269,  80-­‐95,  2008.  

McDermott,  F.:  Centennial-­‐scale  Holocene  climate  variability  revealed  by  a  high-­‐resolution  speleothem  delta  O-­‐18  record  from  SW  Ireland  (9  Nov,  pg  1328,  2001),  Science,  309,  1816-­‐1816,  2005.  

McDermott,  F.,  Atkinson,  T.  C.,  Fairchild,  I.  J.,  Baldini,  L.  M.,  and  Mattey,  D.  P.:  A  first  evaluation  of  the  spatial  gradients  in  delta  O-­‐18  recorded  by  European  Holocene  speleothems,  Glob.  Planet.  Change,  79,  275-­‐287,  2011.  

Mühlinghaus,  C.,  Scholz,  D.,  and  Mangini,  A.:  Modelling  fractionation  of  stable  isotopes  in  stalagmites,  Geochimica  Et  Cosmochimica  Acta,  doi:  10.1016/j.gca.2009.09.010,  2009.  7275-­‐7289,  2009.  

Mühlinghaus,  C.,  Scholz,  D.,  and  Mangini,  A.:  Modelling  stalagmite  growth  and  delta  C-­‐13  as  a  function  of  drip  interval  and  temperature,  Geochimica  Et  Cosmochimica  Acta,  71,  2780-­‐2790,  2007.  

Nicholas,  F.  J.  and  Glasspoole,  J.:  General  monthly  rainfall  over  England  and  Wales,  1727  to  1931,  British  Rainfall,  1931.  299-­‐306,  1931.  

Niggemann,  S.,  Mangini,  A.,  Mudelsee,  M.,  Richter,  D.  K.,  and  Wurth,  G.:  Sub-­‐Milankovitch  climatic  cycles  in  Holocene  stalagmites  from  Sauerland,  Germany,  Earth  and  Planetary  Science  Letters,  216,  539-­‐547,  2003.  

Perette,  Y.:  Etude  de  la  structure  interne  des  stalagmites:  contribution  a  la  connaisance  geographique  des  evolutions  environnementales  du  Vercors  (France),  2000.  Geographie,  Univ.  de  Savoie,  France,  277  pp.,  2000.  

Page 139: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium      

Paper  in  review  in  Climate  of  the  Past   129  

Quinif,  Y.:  Une  nouvelle  topographie  de  la  Grotte  de  Han,  Lapiaz  hors  serie  "Special  Han",  1988.  15-­‐18,  1988.  

Schmidt,  G.  A.,  LeGrande,  A.  N.,  and  Hoffmann,  G.:  Water  isotope  expressions  of  intrinsic  and  forced  variability  in  a  coupled  ocean-­‐atmosphere  model,  Journal  of  Geophysical  Research-­‐Atmospheres,  112,  2007.  

Scholz,  D.  and  Hoffmann,  D.  L.:  StalAge  -­‐  An  algorithm  designed  for  construction  of  speleothem  age  models,  Quaternary  Geochronology,  6,  369-­‐382,  2011.  

Scholz,  D.,  Muehlinghaus,  C.,  and  Mangini,  A.:  Modelling  delta  C-­‐13  and  delta  O-­‐18  in  the  solution  layer  on  stalagmite  surfaces,  Geochimica  Et  Cosmochimica  Acta,  73,  2592-­‐2602,  2009.  

Spötl,  C.  and  Mangini,  A.:  Stalagmite  from  the  Austrian  Alps  reveals  Dansgaard-­‐Oeschger  events  during  isotope  stage  3:  Implications  for  the  absolute  chronology  of  Greenland  ice  cores,  Earth  and  Planetary  Science  Letters,  203,  507-­‐518,  2002.  

Trouet,  V.,  Esper,  J.,  Graham,  N.  E.,  Baker,  A.,  Scourse,  J.  D.,  and  Frank,  D.  C.:  Persistent  Positive  North  Atlantic  Oscillation  Mode  Dominated  the  Medieval  Climate  Anomaly,  Science,  324,  78-­‐80,  2009.  

Van  Engelen,  A.  F.  V.,  Buisman,  J.,  and  Ijnsen,  F.:  A  millennium  of  weather,  winds  and  water  in  the  Low  Countries.  In:  History  and  Climate,  memories  of  the  future?,  Kluwer  Academics,  2001.  2001.  

Van  Rampelbergh,  M.,  Verheyden,  S.,  Allan,  M.,  Quinif,  Y.,  Keppens,  E.,  and  Claeys,  P.:  Seasonal  variations  recorded  in  cave  monitoring  results  and  a  10-­‐year  monthly  resolved  speleothem  δ18O  and  δ13C  record  from  the  Han-­‐sur-­‐Lesse  cave,  Belgium.,  Climate  of  the  Past,  10,  1-­‐15,  2014.  

Verheyden,  S.:  Speleothems  as  palaeoclimatic  archives.  Isotopic  and  geochemical  study  of  the  cave  environment  and  its  Late  Quaternary  records.,  Unpubl.  PhD  Thesis,  Vrije  Universiteit  Brussel,  Belgium,  132p.,  2001.  

Verheyden,  S.,  Baele,  J.-­‐M.,  Keppens,  E.,  Genty,  D.,  Cattani,  O.,  Hai,  C.,  Edwards,  L.,  Hucai,  Z.,  Van  Strijdonck,  M.,  and  Quinif,  Y.:  The  proserpine  stalagmite  (Han-­‐sur-­‐Lesse  cave,  Belgium):  Preliminary  environmental  interpretation  of  the  last  1000  years  as  recorded  in  a  layered  speleothem,  Geologica  Belgica,  9,  245-­‐256,  2006.  

Verheyden,  S.,  Genty,  D.,  Deflandre,  G.,  Quinif,  Y.,  and  Keppens,  E.:  Monitoring  climatological,  hydrological  and  geochemical  parameters  in  the  Pere  Noel  cave  (Belgium):  implication  for  the  interpretation  of  speleothem  isotopic  and  geochemical  time-­‐series,  International  Journal  of  Speleology,  37,  221-­‐234,  2008.  

Verheyden,  S.,  Keppens,  E.,  Quinif,  Y.,  Cheng,  H.  J.,  and  Edwards,  L.  R.:  Late-­‐glacial  and  Holocene  climate  reconstruction  as  inferred  from  a  stalagmite  -­‐  Grotte  du  Pere  Noel,  Han-­‐sur-­‐Lesse,  Belgium,  Geologica  Belgica,  17,  83-­‐89,  2014.  

Page 140: Maïté Van Rampelbergh

Chapter  5:  A  500  year  speleothem  record  from  Belgium    

Paper  in  review  in  Climate  of  the  Past    

130  

Wackerbarth,  A.,  Scholz,  D.,  Fohlmeister,  J.,  and  Mangini,  A.:  Modelling  the  delta  O-­‐18  value  of  cave  drip  water  and  speleothem  calcite,  Earth  and  Planetary  Science  Letters,  299,  387-­‐397,  2010.    

Page 141: Maïté Van Rampelbergh

Chapter  6:  General  Conclusions  and  Perspectives  

  131  

Chapter  6        

General  Conclusions  and  Perspectives      In   this   thesis,   the   use   of   speleothems   as   tools   to   reconstruct   the   paleoclimate  during  the  Holocene  is  evaluated  for  Socotra  (Yemen)  and  Belgium.  The  results  from   this   work   provided   better   insights   in   the   evolution   of   oceanic   and  atmospheric   phenomena   such   as   the   Indian   Ocean   Monsoon   (IOM)   and   the  Intertropical   Convergence   Zone   (ITCZ)   around   the   northern   Indian   Ocean.  Results   of   both   study   sites   clarified   how   speleothem   proxies   relate   to   climate  variations   in   semi-­‐arid   (Socotra)   and   temperate   (Belgium)   climates,   with  resolutions  up  to  seasonal  scale.      In  Chapter  3  a  multi-­‐cave,  multi-­‐speleothem  and  multi-­‐proxy  (δ18O,  δ13C,  Mg/Ca  and  Sr/Ca)  approach  was  applied  to  investigate  the  IOM  system  in  the  northern  Indian   Ocean,   which   extends   the   knowledge   already   established   by   previous  studies   in   that   region   (Burns   et   al.,   2003;   Gupta   et   al.,   2003;   Fleitmann   et   al.,  2007;   Lezine   et   al.,   2010).   The   speleothems   collected   on   the   eastern   side   of  Socotra  Island,  provided  the  first  successful  record  of  northeast  winter  monsoon  precipitation  variations  since  6  000  a  BP.  A  major  conclusion  of  the  study  is  the  different  evolution  of  the  northeast  winter  monsoon  compared  to  the  southwest  summer  monsoon,  with  no  link  between  the  winter  monsoon  variations  and  the  North   Atlantic   climate   variations.   To   understand   the   exact   factors   driving   the  northeast  winter  monsoon,  better  geographically  spread,  higher-­‐resolution  and  longer   time   records   are   necessary.   Of   major   interest   would   be   to   obtain   an  Eastern  Socotran  speleothem  record  that  covers  the  last  15  ka  to  investigate  how  larger   amplitude   climate   variations   such   as   the   Last   Glacial  Maximum,   Bølling  Allerød  and  Younger  Dryas   influenced   the  northeast  winter  monsoon.  A   longer  record  should  also  contain  the  Holocene  humid  period,  lasting  between  ±10  and  ±6   ka,   and   would   help   clarifying   its   timing   and   intensity   in   the   northwestern  Indian   Ocean   (Lezine   et   al.,   2014).   High-­‐resolution   speleothem   data   from  regions,   that   have   not   been   studied,   could   provide   the   necessary   links   to   help  understand  the  IOM  winter  monsoon  forcing  (Fig.  1).  Speleothem  data  from  such  

Page 142: Maïté Van Rampelbergh

Chapter  6:  General  Conclusions  and  Perspectives    

 132  

region  like  Iraq,  Iran  and  Pakistan  could  indicate  the  link  between  IOM  variations  and  Mediterranean  climate  variations  such  as  observed  in  Israel  (Bar-­‐Matthews  et  al.,  1997;  1999;  2000;  2003;  2011)  and  Lebanon  (Verheyden  et  al.,  2008).  The  eastern  African  coast,  a  region  that  has  hardly  been  studied,  could  possibly  house  numerous   caves,   the   study   of   which   could   help   to   connect   the   variations  observed  between  the  northern  and  southern  Indian  Ocean.        

 Figure   1.   Wind   pattern   of   the   winter   monsoon   above   the   Indian   Ocean   in  January.   The   southern   edge   of   the   ITCZ   migration   path   is   located   around   the  latitude   of   Madagascar.   The   red   question   marks   indicate   the   regions   where  currently  no  speleothem  based  paleoclimate  reconstructions  are  available.  High-­‐resolution   long-­‐term   speleothem   records   from   these   regions   could   provide  important   links   between   climate   variations   and   highlight   their   connection  (adapted  after  Van  Rampelbergh  et  al.  2013).    Of   particular   interest   is   the   island   of   Madagascar   that   displays   two   different  precipitation   regimes   in   the   north   and   the   south   of   the   island   separated   by   a  mountain  range  (similar  to  the  situation  on  Socotra  Island).  Speleothem  records  from   both   sides   of   the   Island   could   bring   new   insights   in   the   evolution   of   the  summer  and  winter  monsoons  at  the  southern  limit  of  the  ITCZ.  A  recent  ocean  core   study   from   the   Mauritian   lowlands   (east   of   Madagascar)   indicated   that  Mauritian  rainfall  and  the  Indian  and  Asian  summer  monsoons  are  linked  since  they  both  receive  moisture  from  the  southern  equatorial  Indian  Ocean  (de  Boer  et  al.,  2014).  This   study  highlights   the  complex  climate   system   in   the  Southern  Indian   Ocean   where   speleothems   could   help   providing   very   valuable  complementary   information.   Apart   from   refining   the   IOM  mechanics,   the   trace  elemental  records  in  the  Socotran  stalagmites  also  helped  to  gain  insights  in  how  

Page 143: Maïté Van Rampelbergh

Chapter  6:  General  Conclusions  and  Perspectives  

  133  

trace  elements  are   inked  to  climate   in  semi-­‐arid  regions  (Rutlidge  et  al.,  2014).  Analyzing  other  trace  elements,  such  as  Ba,  Na,  Fe,  Pb,  U,  Cd,  etc.  in  the  Socotran  speleothems  may  further  provide  insights  in  how  they  can  be  used  to  reconstruct  climate  in  semi-­‐arid  regions.        In  Chapter  4  and  5,   the   potential   of   a   Belgian   fast   growing   and   visibly   layered  speleothem  (Proserpine)  from  the  Han-­‐sur-­‐Lesse  cave  in  recording  the  climate  at  seasonal   scale   was   evaluated.   Results   from   the   cave   monitoring   (Chapter   4)  indicated   that   the   cave   parameters   (cave   air   and   drip   water   temperature,   the  water  discharge  amount,  pCO2  and  δ13CCO2  values  of  cave  air,  drip  water  pH,  δ18O,  δD  and  δ13CDIC  values)  displayed  seasonal  variations  in  response  to  the  external  climate.   Freshly   farmed   calcite   and   detailed   measurements   in   the   speleothem  confirmed  that  the  seasonal  patterns  are  recorded  in  the  calcite  fabric  and  in  its  δ18O   and   δ13C   values.   A   500-­‐year   time   series,   which   is   seasonally   resolved  between  1810  and  1479  AD,  indicated  that  the  calcite  fabric,  layer  thickness,  and  δ18O   and   δ13C   values   can   successfully   be   used   to   reconstruct   variations   in   the  past   climate   (Chapter   5).   Due   to   the   multitude   of   factors   that   can   affect   the  different   proxies,   a   multiproxy   approach   is   necessary   to   identify   the   factors  influencing   the   proxy   variations   in   Belgian   speleothems.   This   multiproxy  approach  revealed  that  the  last  part  of  the  LIA  was  generally  drier  (and  colder)  with   two   relatively  wetter   (and  warmer)  periods  between  1479  and  1565  and  between  1730  and  1770  AD.  The  seasonally  resolved  series  revealed  a  decrease  in   temperature   seasonality   of   1°C   between   1610   and   1730,   when   conditions  were  generally  drier  and  colder.        The  results  from  the  work  on  the  Proserpine  indicate  that  the  measured  proxies  (δ18O,   δ13C,   layer   thickness   and   calcite   fabric)   successfully   record   climate  variations  up  to  seasonal  scale.  In  a  following  phase,  the  measured  proxies  can  be  extended  further  back  in  time,  up  to  2000  AD  as  is  indicated  by  the  bottom  U/Th-­‐age   of   the   Proserpine   core   (Verheyden   et   al.,   2006),   and   cover   the   Medieval  Warm  Period  (±950-­‐1250  AD)  and  its  transition  into  the  Little  Ice  Age  (±1300-­‐1850).  Studying  the  MWP  and  LIA  climatic  periods  up  to  seasonal   level  has  not  been   done   yet   and  may   throw   a   light   on   how   the   seasonal   variations   evolved  during   these   two   climatic  periods.  Apart   from  extending   the   time   series   of   the  measured  proxies,  other  proxies  such  as  elemental  concentrations  of  Mg,  Sr,  Ca,  U,  Ba,  Pb,  Cd,  Fe,  Zn,…  can  be  measured  and  used  to  investigate  their  link  with  the  environment.   Monitoring   how   these   elements   are   transported   from   the  atmosphere,  the  soil  and  the  epikarst  into  the  drip  water  and  finally  in  the  calcite  will   allow   understanding   which   factors   cause   them   to   vary   over   time.   These  conclusions   can   then  be   applied   on   trace   element   time   series  measured   in   the  core   and   be   used   as   additional   proxies   to   retrieve   past   climatic   and  environmental  changes.  

Page 144: Maïté Van Rampelbergh

Chapter  6:  General  Conclusions  and  Perspectives    

 134  

 The   conclusions   and   perspective   from   this   work   are   based   on   the   currently  available  methods  to  extract  climate  information  from  speleothems.  However,  all  studied   samples   and   cave   sites   most   probably   contain   more   and   other  information   about   past   climates   and   environments,  which   cannot   be   extracted  due   to   methodological   constrains.   Higher   resolution   sampling   and   better  analytical  methods  and  tools  may  not  only  allow  to  increase  the  time  resolution  of   the   known  proxies   but  will   also  help   in   developing  new  proxies.  One   of   the  most  promising  perspectives  is  the  U/Pb-­‐dating  technique,  which  will  allow  the  speleothem   community   to   reconstruct   climates   past   the  U/Th-­‐age   limit   of   600  ka.   Also,   further   work   on   the   clumped   isotope   technique   or   fluid   inclusion  measurements  form  an  important  perspective  for  deriving  absolute  temperature  variations  from  speleothems.  Finally,  the  large  collection  of  speleothem  samples  and   records  allows  comparing   them   to   retrieve  patterns  and  gradients   such  as  has   been   done   by   McDermott   et   al.   (2011).   The   development   of   a   global  speleothem   database   and   network   should   become   a   priority   for   speleothem-­‐based  paleoclimate  research.    I  would  like  to  conclude  this  thesis  with  a  quote  from  one  of  my  favorite  authors,  Roald  Dahl,  in  his  book  ‘The  Minpins’.  These  words  perfectly  describe  the  science  of  speleothems,  as  I  have  experienced  it.  At  first  sight,  no  one  would  expect  the  dark  caves  and  their  calcite  formations  to  be  such  great  environments  to  record  climate  variations.  Since   the  start  of   the  speleothem  science,   together,  we  have  been  able  to  read  these  climate  archives  and  understand  how  the  past  terrestrial  climates  evolved.  Even   if   speleothems  already   taught  us  a   lot,  we  have   to  keep  watching  because  they  still  contain  a  lot  of  more  secrets,  which  we  are  not  able  to  see  yet.      “And  above  all,  watch  with  glittering  eyes  the  whole  world  around  you  because  the  greatest  secrets  are  always  hidden  in  the  most  unlikely  places.”      

Roald  Dahl                    

Page 145: Maïté Van Rampelbergh

Chapter  6:  General  Conclusions  and  Perspectives  

  135  

 REFERENCES    Bar-­‐Matthews,  M.   and  Ayalon,   A.:  Mid-­‐Holocene   climate   variations   revealed   by  high-­‐resolution  speleothem  records  from  Soreq  Cave,  Israel  and  their  correlation  with  cultural  changes,  Holocene,  21,  163-­‐171,  2011.  

Bar-­‐Matthews,  M.,  Ayalon,  A.,  Gilmour,  M.,  Matthews,  A.,  and  Hawkesworth,  C.  J.:  Sea-­‐land   oxygen   isotopic   relationships   from   planktonic   foraminifera   and  speleothems   in   the   Eastern   Mediterranean   region   and   their   implication   for  paleorainfall  during  interglacial  intervals,  Geochimica  Et  Cosmochimica  Acta,  67,  3181-­‐3199,  2003.  

Bar-­‐Matthews,  M.,  Ayalon,  A.,  and  Kaufman,  A.:  Late  quaternary  paleoclimate   in  the  eastern  Mediterranean  region  from  stable  isotope  analysis  of  speleothems  at  Soreq  Cave,  Israel,  Quaternary  Research,  47,  155-­‐168,  1997.  

Bar-­‐Matthews,   M.,   Ayalon,   A.,   and   Kaufman,   A.:   Timing   and   hydrological  conditions   of   Sapropel   events   in   the   Eastern   Mediterranean,   as   evident   from  speleothems,  Soreq  cave,  Israel,  Chemical  Geology,  169,  145-­‐156,  2000.  

Bar-­‐Matthews,   M.,   Ayalon,   A.,   Kaufman,   A.,   and  Wasserburg,   G.   J.:   The   Eastern  Mediterranean  paleoclimate  as  a  reflection  of  regional  events:  Soreq  cave,  Israel,  Earth  and  Planetary  Science  Letters,  166,  85-­‐95,  1999.  

Burns,   S.   J.,   Fleitmann,   D.,  Matter,   A.,   Kramers,   J.,   and  Al-­‐Subbary,   A.   A.:   Indian  Ocean  climate  and  an  absolute  chronology  over  Dansgaard/Oeschger  events  9  to  13,  Science,  301,  1365-­‐1367,  2003.  

de   Boer,   E.   J.,   Tjallingii,   R.,   Velez,   M.   I.,   Rijsdijk,   K.   F.,   Vlug,   A.,   Reichart,   G.   J.,  Prendergast,   A.   L.,   de   Louw,   P.   G.   B.,   Florens,   F.   B.   V.,   Baider,   C.,   and  Hooghiemstra,  H.:   Climate   variability   in   the   SW   Indian  Ocean   from  an  8000-­‐yr  long   multi-­‐proxy   record   in   the   Mauritian   lowlands   shows   a   middle   to   late  Holocene   shift   from   negative   IOD-­‐state   to   ENSO-­‐state,   Quaternary   Science  Reviews,  86,  175-­‐189,  2014.  

Fleitmann,  D.,  Burns,  S.  J.,  Mangini,  A.,  Mudelsee,  M.,  Kramers,  J.,  Villa,  I.,  Neff,  U.,  Al-­‐Subbary,   A.   A.,   Buettner,   A.,   Hippler,   D.,   and   Matter,   A.:   Holocene   ITCZ   and  Indian   monsoon   dynamics   recorded   in   stalagmites   from   Oman   and   Yemen  (Socotra),  Quaternary  Science  Reviews,  26,  170-­‐188,  2007.  

Gupta,   A.   K.,   Anderson,  D.  M.,   and  Overpeck,   J.   T.:   Abrupt   changes   in   the  Asian  southwest  monsoon   during   the   Holocene   and   their   links   to   the   North   Atlantic  Ocean,  Nature,  421,  354-­‐357,  2003.  

Lezine,   A.-­‐M.,   Bassinot,   F.,   and   Peterschmitt,   J.-­‐Y.:   Orbitally-­‐induced   changes   of  the   Atlantic   and   Indian   monsoons   over   the   past   20,000   years:   New   insights  based   on   the   comparison   of   continental   and   marine   records,   Bulletin   De   La  Societe  Geologique  De  France,  185,  3-­‐12,  2014.  

Page 146: Maïté Van Rampelbergh

Chapter  6:  General  Conclusions  and  Perspectives    

 136  

Lezine,   A.   M.,   Robert,   C.,   Cleuziou,   S.,   Inizan,   M.   L.,   Braemer,   F.,   Saliege,   J.   F.,  Sylvestre,   F.,   Tiercelin,   J.   J.,   Crassard,   R.,  Mery,   S.,   Charpentier,   V.,   and   Steimer-­‐Herbet,   T.:   Climate   change   and   human   occupation   in   the   Southern   Arabian  lowlands  during  the  last  deglaciation  and  the  Holocene,  Glob.  Planet.  Change,  72,  412-­‐428,  2010.  

McDermott,  F.,  Atkinson,  T.  C.,  Fairchild,   I.   J.,  Baldini,  L.  M.,   and  Mattey,  D.  P.:  A  first   evaluation   of   the   spatial   gradients   in   delta   O-­‐18   recorded   by   European  Holocene  speleothems,  Glob.  Planet.  Change,  79,  275-­‐287,  2011.  

Rutlidge,  H.,  Baker,  A.,  Marjo,  C.  E.,  Andersen,  M.  S.,  Graham,  P.  W.,  Cuthbert,  M.  O.,  Rau,   G.   C.,   Roshan,   H.,   Markowska,   M.,   Mariethoz,   G.,   and   Jex,   C.   N.:   Dripwater  organic   matter   and   trace   element   geochemistry   in   a   semi-­‐arid   karst  environment:   Implications   for   speleothem   paleoclimatology,   Geochimica   Et  Cosmochimica  Acta,  135,  217-­‐230,  2014.  

Van   Rampelbergh,   M.,   Fleitmann,   D.,   Verheyden,   S.,   Cheng,   H.,   Edwards,   L.,   De  Geest,  P.,  De  Vleeschouwer,  D.,  Burns,  S.  J.,  Matter,  A.,  Claeys,  P.,  and  Keppens,  E.:  Mid-­‐   to   late   Holocene   Indian   Ocean   Monsoon   variability   recorded   in   four  speleothems  from  Socotra  Island,  Yemen,  Quaternary  Science  Reviews,  65,  129-­‐142,  2013.  

Verheyden,  S.,  Baele,  J.-­‐M.,  Keppens,  E.,  Genty,  D.,  Cattani,  O.,  Hai,  C.,  Edwards,  L.,  Hucai,  Z.,  Van  Strijdonck,  M.,  and  Quinif,  Y.:  The  proserpine  stalagmite  (Han-­‐sur-­‐Lesse  cave,  Belgium):  Preliminary  environmental  interpretation  of  the  last  1000  years  as  recorded  in  a  layered  speleothem,  Geologica  Belgica,  9,  245-­‐256,  2006.  

Verheyden,   S.,   Nader,   F.   H.,   Cheng,   H.   J.,   Edwards,   L.   R.,   and   Swennen,   R.:  Paleoclimate   reconstruction   in   the   Levant   region   from   the   geochemistry   of   a  Holocene  stalagmite  from  the  Jeita  cave,  Lebanon,  Quaternary  Research,  70,  368-­‐381,  2008.