Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon isotopes of diatom frustules Armand Hernández Hernández Aquesta tesi doctoral està subjecta a la llicència Reconeixement- NoComercial 4.0. Espanya de Creative Commons. Esta tesis doctoral está sujeta a la licencia Reconocimiento - NoComercial 4.0. España de Creative Commons. This doctoral thesis is licensed under the Creative Commons Attribution-NonCommercial 4.0. Spain License. ULTRA-HIGH RESOLUTION ENVIRONMENTAL AND CLIMATIC RECONSTRUCTION USING OXYGEN AND CARBON ISOTOPES OF DIATOM FRUSTULES The Late Glacial and Early Holocene laminated sediments from Lago Chungará (Andean Altiplano, northern Chile) Armand Hernández PhD thesis Universitat de Barcelona June 2010 UNIVERSIDADE DA CORUÑA Consejo Superior de Investigaciones Científicas CSI ULTRA-HIGH RESOLUTION ENVIRONMENTAL AND CLIMATIC RECONSTRUCTION USING OXYGEN AND CARBON ISOTOPES OF DIATOM FRUSTULES The Late Glacial and Early Holocene laminated sediments from Lago Chungará (Andean Altiplano, northern Chile) Memòria presentada per l’Armand Hernández Hernández per optar al grau de Doctor en Geologia. Aquesta memòria ha estat realitzada dins del Programa de Doctorat «Exploració, Anàlisi i Modelització de Conques i Sistemes Orogènics (Bienni 2004-2006)» sota la direcció del Dr. Santiago Giralt Romeu i del Dr. Roberto Bao Casal, i la tutoria del Dr. Alberto Sáez Ruiz. Armand Hernández Hernández Barcelona, Juny del 2010 Dr. Santiago Giralt Romeu Dr. Roberto Bao Casal UNIVERSITAT DE BARCELONA U B Al meu pare Agraïments Ara mateix no estaries llegint aquesta tesi si no fos per totes les persones que m’envolten. Aquest és el moment d’agrair-los a tots la seva ajuda, el seu recolzament, la seva comprensió, la seva paciència (sobretot paciència), els seus ànims, la seva amistat, les seves rialles, i l’immens afecte que m’han donat. Si una cosa tinc molt clara és que sense tots ells mai hauria arribat fins aquí. Per això, i encara que sigui d’una forma molt discreta, m’agradaria donar-los les gràcies. En primer lloc agrair l’esforç fet pels meus directors de tesi, ells saben millor que ningú com ha costat arribar fins aquí. Però ells també saben que durant tot aquest temps ha sorgit alguna cosa més que una relació director-doctorand. En Santi i el Roberto han estat amics, col·legues, pares, i alguna vegada «jefes» (sí, alguna vegada sí). Així només puc dir-vos: gràcies, gràcies i gràcies per tot. Aínda que tamén podería dicirvos: grazas, grazas e grazas por todo. I wish to express my sincere thanks to Melanie J Leng and Philip A Barker, my English advisors, who have made a major contribution to this thesis. They went to a lot of trouble to make me feel comfortable in my new surroundings in England. Despite the language difficulties, they showed kindness and forbearance and made me very welcome. They believed in me, and if I may be considered a researcher, it is thanks to their support and encouragement. They were a source of inspiration and their advice proved invaluable in helping me to get to grips with isotope and diatom research. It was through Melanie and Phil that I became a member of the IBiS group. I am especially grateful to Hillary J Sloane, who played a very important part in this study. Without her isotope analysis, this thesis would not have been written. In addition to being very supportive, she willingly took the time to teach me about the secrets of the fluorination line during my stay at NIGL (UK). I am also indebted to Christopher P Kendrick for undertaking the carbon isotope analysis. També vull agrair l’ajuda, l’amistat i la dedicació del «compañero y gurú» Alberto Saéz, i el recolzament, les rialles, els savis consells i el suport econòmic d’en Juan José Pueyo. Tots nosaltres formem part d’un grup de recerca compacte on el que està per sobre de tot és el bon ambient i les ganes de fer les coses ben fetes. Aquí em venen al cap reunions de projecte, campanyes de camp, sopars... Per això, vull donar les gràcies a tot el grup de recerca: a Blas L Valero-Garcés padre de todo esto y gran maestro del Altiplano, a Ana Moreno por su actitud incansable y por estar siempre dispuesta a ayudar, v al Christian Herrera, por su ayuda desde Chile. També als que han anat arribant: en Jordi Catalán, en Sergi Pla, l’Olga Margalef, la Núria Cañellas i en Valentí Rull. Per últim, cal recordar als que ja no formen part del grup: la Conxita Taberner va ser qui em va donar l’oportunitat de començar aquest camí, Bogumila Klosowska, who not only helped me with the sampling but also with my English, i en Roger O Gibert i la Penélope González-Sampériz que van fer una part important del treball inicial al Lago Chungará. És la meva obligació agrair el suport econòmic que he rebut mitjançant una beca FPI del Ministerio de Ciencia e Innovación (BES-2005-6971) i un contracte d’ajudant de recerca amb la Fundació Bosch i Gimpera (CENITE: VISION CO 2 ). Els projectes del Ministerio de Ciencia e Innovación ANDESTER (BTE2001-3225, BTE2001-5257-E), LAVOLTER (CGL2004-00683/BTE), GEOBILA (CGL2007-60932/ BTE) i CONSOLIDER-Ingenio 2010 GRACCIE (CSD2007-00067) han permès el suport econòmic d’aquesta tesi durant aquests anys. Per dur a bon port aquesta tesi s’han hagut de fer viatges, mostreigs, preparació de mostres i anàlisis. Per això no em puc oblidar de mostrar tot el meu agraïment a tothom que ho ha fet possible. Grateful acknowledgment is made to D. Schnurremberger, M. Shapley and A. Myrbo from the Limnological Research Center staff (USA) for their invaluable assistance in the field. Muchas gracias a toda la gente de la CONAF por ofrecerme su ayuda y apoyo. No olvido que estoy en deuda con Jimena Saavedra (CONAF), quien ayudó a que mi estancia en el Altiplano fuera una de las mejores experiencias que uno puede vivir. También agradecer a la familia Riroroko su hospitalidad durante la campaña de campo en la Isla de Pascua ya que, aunque el material obtenido allí no forme parte de esta tesis, sí forma parte del proyecto al que me debo. Durante los muestreos, la gente del IPE (Mayte, Mario, JuanPi,…) siempre ha estado dispuesta a ayudar, y en el laboratorio del ICTJA siempre he tenido la colaboración de Graciela Monzón. La actitud y predisposición para ayudar en todo lo posible por parte de Chelo Palacios, de la administración del ICTJA, ha sido de gran valor. Por tanto, gracias a todos ellos. I am also indebted to Michael Köhler (GFZ-Potsdam) for preparing the thin sections and to Andy Quin for helping me in the lab at LEC (UK). Special thanks are due to Elizabeth Hurrell for making my stay at LEC such a nice experience and for compiling a vocabulary of useful words in English and Spanish. Her «Catalan almond cookies» were unforgettable. Thanks are due to Alice Chang for the many interesting and stimulating discussions about rhythmites and the self-sedimentation process. This thesis benefited substantially from the critical reviews of Antje Schwalb, Sarah Metcalfe and two anonymous persons. I wish to thank George von Knorring for improving the English of the final version of this thesis. No vull oblidar els meus inicis, i per això també li estic agraït a l’Emilio Ramos i al Mariano Marzo per donar-me l’oportunitat de fer recerca. Amb ells vaig poder començar a conèixer aquest món i fer els no sempre engrescadors cursos de doctorat. vi També vull tenir un record especial per a tota aquella gent amb qui he compartit despatx durant aquest temps. Sóc perfectament conscient que no és fàcil compartir un espai així amb mi, i ells en lloc de posar males cares sempre s’han mostrat disposats a ajudar-me. Així doncs, moltes gràcies a l’Alfonso Muñoz, la Gemma Labraña, l’Almudena Lorenzo i la Marta Rejas, que han compartit amb mi el «zulo» del ICTJA; e a Manel Leira, Tânia Ferreira e Laia Rovira, que compartiron comigo «o laboratorio» no anexo da Facultade de Ciencias da UDC. Tamén agradecer a amizade, boa vontade e ganas de axudar da xente da UDC. Quizá eles tamén teñan parte de culpa de que decida acabar a tese en Galicia. Eles axudaron, pero tamén as marabillosas praias que alberga esta terra, a súa natureza, a súa rica gastronomía, a súa calidade de vida, a súa tranquilidade, os seus paxaros, a súa xente, o seu tempo (dacordo, o seu tempo non), e o seu “Mesón del Hockey” (o noso santuario), un claro exemplo de biodiversidade cunha tortilla marabillosa. Alí, xurdiron moitas ideas desta tese. Una agraïment molt especial per la Patricia Cabello, ella va ser qui em va fer creure que jo podia fer una tesi, i qui em va ensenyar a no rendir-me mai i lluitar pel que realment vull. Sempre ha estat molt exigent, però alhora pacient, amb tot el que li he ensenyat perquè em donés la seva opinió. Per tant, ella també té gran part de culpa del resultat d’aquesta tesi. Vull agrair a la Clara Prats tots els consells que m’ha donat des de la seva «experiència doctoril», hi ha una llista interminable de recomanacions, idees i discussions que m’han ajudat a arribar fins aquí. També vull donar les gràcies a la Núria Carrera per haver-me recolzat i escoltat en molts moments, per haver-me fet ser més crític amb mi mateix i per haver-me ajudat amb els seus coneixements de la geologia dels Andes. Estic molt content, orgullós, i fins i tot sorprès, de la quantitat d’amics i companys que tinc. Aquí, estan inclosos tots els «Tragapans», la majoria dels quals han estat estudiants de geologia. Alguns no en van tenir prou amb la llicenciatura i han fet una tesi, un màster o simplement han allargat la carrera. Això ens ha permès compartir dinars, cafès i birres, fent més agradables les llargues jornades de feina a la Facultat. També vull donar les gràcies a tots els amics de Sant Feliu, que potser no han intervingut directament en la realització d’aquesta tesi però que són una part molt important de la meva vida. Ells són la gent de la «penya espardenya» o «Qtales», amb qui a vegades m’he sentit com un «bitxo raro» explicant les meves històries de diatomees, i la «gent de l’escola», «del MotoGP» i «de les calçotades», que m’han fet adonar que sóc un privilegiat cada cop que els he explicat les meves aventures de tesi (només els hi explicava les coses bones). També vull donar les gràcies als «amics de la piscina» que encara em queden i als «geòlegs» que ja no estan a la facultat. Amb tots ells comparteixo una part important de la meva vida i una gran amistat. When I embarked upon my thesis, my supervisors told me that I would have to do a training period in the UK. I was apprehensive at first as I did not know anybody in England and as I spoke very vii little English. But now I am very glad that I went. I met many kind and interesting people, I improved my English and I decided to return the following year. I should like to take this opportunity to thank all my friends and colleagues in England for their help and encouragement. Per últim, gràcies a tota la meva família. Al meu pare per haver-me fet veure que aquest camí no seria fàcil, però que lluitant algun dia arribaria al final. Malauradament ell no ha pogut arribar a veure aquest final, però tal com parlava i em mirava sé que va marxar convençut de que ho aconseguiria. Papa, aquesta tesi és per tu. A la meva mare, per haver-me recolzat sempre i per haver tirat amb tota la seva força i energia d’aquest carro que anomenem família. Gracias mama. Al meu germà per ser el millor que m’ha passat a la vida, sempre he pensat així i tot aquest temps de tesi només ha fet que refermar aquesta idea. Gràcies Ferran per ser com ets, per fer-me sentir admirat, per dir-me sempre que té molt de mèrit el que faig, per fer-me riure, i pel teu suport lingüístic. A les meves iaies per haver aconseguit que estigui convençut i segur de fer el que faig i de ser com sóc. Ambdues, des de la seva experiència, sempre m’han dit el mateix: «El millor que pots fer és veure món i omplir-te d’experiències». I això és el que he fet i vull seguir fent. I als meus cosins, tiets i cosinets, per ser família i amics al mateix temps, ells sempre han admirat la meva feina, el meu valor i la meva determinació, i això m’ha fet avançar amb força durant tot aquest temps de tesi. Gemma, no et pensis que m’oblido de tu! Junts estem compartint els millors anys de la nostra vida, i junts hem aconseguit acabar aquesta tesi. Gràcies per la teva paciència, gràcies pel teu amor, gràcies pel teu afecte, gràcies per la teva comprensió, gràcies per fer-me veure que no sempre tinc raó, gràcies per acompanyar-me en les meves bogeries, gràcies per fer-me feliç, gràcies per fer-me sentir estimat, gràcies pel teu somriure, gràcies als teus pares, gràcies per la teva lasagna d’espinacs i, evidentment, gràcies per la teva feina a l’Altiplano. Gràcies a tots, gracias a todos, grazas a todos, thanks to everybody. viii Resum Introducció Els isòtops estables de la sílice de les diatomees sovint s’han utilitzat amb èxit per dur a terme reconstruccions paleoambientals, especialment paleoclimàtiques. Els isòtops més emprats són els d’oxigen (δ18O diatom ) tot i que, darrerament, l’ús d’isòtops de carboni (δ13C diatom ), silici (δ30Si diatom ) i fins i tot nitrogen (δ15N diatom ) ha experimentat un increment important. Tot i això, molts camps romanen oberts i estan pendents de ser explorats amb aquests isòtops. Els isòtops estables de la sílice de les diatomees són especialment útils per poder obtenir informació en aquells registres sedimentaris on la presència de carbonats és escassa o nul·la, i en alguns casos (δ13C diatom i δ15N diatom ) per evitar efectes diagnètics en la senyal isotòpica obtinguda. L’anàlisi de δ18O diatom s’ha aplicat tant en sediments marins com en sediments lacustres. Per contra, la resta d’isòtops pràcticament no han estat utilitzats en sediments lacustres degut a la gran quantitat de factors que intervenen en la seva incorporació a les diatomees. Per altra banda, els registres lacustres tropicals són molt importants per a dur a terme reconstruccions climàtiques, ja que es troben situats en una posició geogràfica clau per entendre els canvis climàtics del passat i, així, poder-nos donar informació fonamental per entendre els canvis climàtics del futur. Els llacs són uns excel·lents sensors dels canvis ambientals i, per tant, el seu rebliment sedimentari conté molta informació que ens pot ajudar a entendre millor els canvis ambientals que van succeir en el passat. Així doncs, l’estudi dels isòtops estables de les diatomees de sediments lacustres situats en llocs estratègics pot ser una font d’informació molt important per entendre els canvis ambientals del passat i, així, obtenir una visió privilegiada del context actual de canvi global. Un d’aquests llocs estratègics és l’Altiplà Andí i, per tant, els llacs situats en aquesta regió són uns bons candidats per dur a terme aquest tipus de reconstruccions. Fins a l’actualitat, però, en els seus registres sedimentaris només s’han portat a terme anàlisis d’isòtops en carbonats i en matèria orgànica total. Tot i que les restes de diatomees acostumen a estar ben preservades en aquests registres, encara no s’han aplicat anàlisis de δ18O diatom i δ13C diatom en aquesta regió. xi Objectius Els objectius d’aquesta Tesi Doctoral són: 1) explorar les diferents possibilitats que pot oferir l’estudi de δ18O diatom i δ13C diatom de sediments lacustres en reconstruccions paleoambientals, i 2) dur a terme reconstruccions ambientals i climàtiques, tant a alta com a molt alta resolució temporal, de la regió del Lago Chungará (Altiplà Andí) durant el Tardiglacial i l’Holocè inicial, fent ús dels isòtops estables abans esmentats. Característiques del Lago Chungará El Lago Chungará (18º15’S, 69º10’W, 4520 m s.n.m.) es troba situat a l’Altiplà Andí (Andes Centrals, Xile), en un context tectònic i volcànic molt actiu. El llac es va formar com a conseqüència d’una esllavissada, durant un col·lapse parcial del volcà Parinacota, la qual va tallar el curs del riu Chungará donant lloc quasi d’immediat a la formació del llac. La data de formació del llac encara aixeca controvèrsia, però sembla clar que es va formar entre els 13000 i els 20000 anys BP. El Lago Chungará presenta una morfologia irregular amb una longitud màxima de 8,75 km, una profunditat màxima de 40 m, una superfície total de 21,5 km2 i un volum aproximat de 400 hm3. Els marges nord i oest del llac són abruptes i estan formats per les vessants dels Volcans Parinacota i Ajoya, respectivament. Els marges sud i est, per contra, són de pendent suau, formats per les parts distals de diversos cons al·luvials i la vall del riu Chungará. Actualment, la principal entrada d’aigües al llac ve donada pel riu Chungará i la principal sortida d’aigua és per evaporació. El llac es pot considerar polimíctic i de oligomesoeutròfic cap a mesoeutròfic, amb un contingut d’1,2 g l-1 de sòlids dissolts. La conductivitat varia entre 1500 i 3000 μS cm-1, és un llac alcalí i el seu quimisme és del tipus Na+-Mg2+-HCO3--SO42-. L’aigua del llac està enriquida isotòpicament respecte l’aigua de precipitació per mitjà de l’evaporació i els seus valors mitjans de δ18O i δD són –1,1‰ SMOW i –39,2‰ SMOW, respectivament. La regió del Lago Chungará està dominada per un clima majoritàriament àrid, amb una època humida anomenada «Invierno Boliviano». Aquesta època es concentra en els mesos de l’estiu austral i ve determinada per la migració cap al sud del cinturó de convergència intertropical, encarregat de portar les masses d’aire humides des de l’Oceà Atlàntic cap a l’Altiplà. En aquesta regió, un altre component important de la variabilitat climàtica entre anys diferents és el fenomen d’El Niño-Oscil·lació del Sud (ENSO). Per això, els anys considerats «Niño» presenten valors de precipitació més baixos i, en canvi, durant els anys considerats «Niña» els valors de precipitació són més elevats. xii Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles Treballs previs realitzats al Lago Chungará Aquesta Tesi ha estat realitzada dins d’un grup de recerca multidisciplinar i, per tant, la seva idea i naixement sorgeixen dins d’aquest àmbit. Així, la feina realitzada per aquest grup és imprescindible per entendre el context en que s’emmarca aquesta tesi. Al mes de novembre del 2002 es va portar a terme la campanya de sondeig del Lago Chungará, on es van recuperar quinze testimonis de sondatge. Tots aquests testimonis, de fins a 8 metres de longitud, van ser tallats en seccions de 1,5 metres de llargada i transportats al laboratori. Allà, i mitjançant un GEOTEKTM Multi-Sensor Core Logger (MSCL), es van mesurar a cada centímetre certes propietats físiques com la densitat, la velocitat de les ones p i la susceptibilitat magnètica. Amb posterioritat, els testimonis van ser tallats en dues meitats i es van descriure les seves textures, colors i estructures sedimentàries. Tot seguit es van preparar frotis segons les diverses facies descrites per estudiar la composició dels sediments. Totes aquestes dades, amb l’ajut de dades sísmiques obtingudes amb anterioritat, van permetre elaborar una reconstrucció 3D del rebliment sedimentari del Lago Chungará. Els sondatges 10 i 11, situats a la zona central del llac, van ser seleccionats per dur a terme la reconstrucció paleoambiental de la zona. Amb aquests dos testimonis de sondatges es va construir un sondatge compost que representa deu metres del rebliment sedimentari. De base a sostre, el sondatge està format per dues unitats sedimentàries (unitats 1 i 2), que van ser dividides en dues subunitats (subunitats 1a, 1b, 2a, 2b). La subunitat basal 1a està formada per una alternança de làmines fines de color blanc i verd, molt riques en diatomees. La subunitat 1b està composta per intercalacions d’intervals laminats i massius de color marró i rics en diatomees. Alguns intervals centimètrics també són rics en carbonats. La subunitat 2a està formada per sediments massius marrons rics en diatomees i amb intercalacions de cendres volcàniques i carbonats. Per últim, la subunitat 2b està composta per sediments massius de color gris i negre rics en diatomees amb abundants intercalacions centimètriques de cendres volcàniques. Els testimonis de sondatge també van ser analitzats per fluorescència de raigs-X, difracció de raigs-X, carboni total i carboni orgànic total, pol·len, associacions de diatomees i sílice biogènica total. El model cronològic de la seqüència sedimentària del Lago Chungará està basat en 17 datacions radiocarbòniques realitzades en matèria orgànica total i macrofòssils de plantes aquàtiques, i en una datació feta mitjançant les sèries de desintegració de l’urani (238U/230Th) d’un nivell de carbonats. L’efecte reservori actual es va calcular a partir de datar el carboni orgànic dissolt a l’aigua del llac i corregir aquesta datació pels efectes dels tests termonuclears realitzats en superfície durant els anys 50 i 60. Totes les datacions es van calibrar mitjançant el software CALIB 5.02 per poder construir el model cronològic final. xiii Resum Mètodes Les metodologies emprades han estat basades en l’estudi de les aigües actuals del Lago Chungará, l’estudi petrogràfic dels sediments i l’estudi dels δ18O diatom i δ13C diatom . Les dades obtingudes han estat tractades estadísticament per tal d’objectivar els resultats. Al desembre del 2009 es van recollir mostres d’aigua del Lago Chungará i cossos d’aigua propers per tal de completar els mostratges realitzats amb anterioritat (2002 i 2004) pel propi grup de recerca. Les mostres d’aigua del Lago Chungará es van recollir cada 2 metres, en 2 perfils verticals de fins a 8 i 20 metres de fondària. In situ es van mesurar la conductivitat, la concentració d’oxigen, el pH, la temperatura i la salinitat. Per a l’anàlisi d’isòtops d’hidrogen i oxigen es van recollir 24 mostres que van ser analitzades per espectrometria de masses (ICP-IRMS) als Serveis Científico-Tècnics de la Universitat de Barcelona. Als mateixos laboratoris també es va dur a terme l’anàlisi per cromatografia de les concentracions iòniques dels elements principals. Els sediments del Lago Chungará van ser mostrejats seguint tres estratègies: 1) Es van seleccionar tres intervals de sediments amb alternança mil·limètrica de làmines clares i fosques riques en diatomees, per a caracteritzar els canvis ambientals i climàtics que van tenir lloc durant la transició del Tardiglacial a l’Holocè inicial (12000-9400 anys cal BP). Aquests intervals van ser mostrejats làmina a làmina amb l’objectiu d’obtenir el màxim d’informació possible dels processos ambientals que van tenir lloc. Una primera selecció de mostres de làmines fosquesprocedents dels tres intervals, les quals indicarien les condicions ambientals de fons, van ser analitzades per a determinar els δ18O diatom amb la intenció de fer una reconstrucció de la evolució hidrològica del Lago Chungará durant aquest període. Un segon estudi isotòpic amb totes les mostres fosques d’un dels intervals va permetre reconstruir a molt alta resolució temporal (entre 4 i 24 anys) la influència de l’ENSO i de l’activitat solar durant el període de temps que inclou aquest interval. 2) Es van analitzar totes les mostres (clares i fosques) de l’anterior interval per a la determinació de δ18O diatom , i una selecció d’11 mostres per a la determinació de δ13C diatom i %C diatom , per tal de caracteritzar a molt alta freqüència els canvis biogeoquímics que van tenir lloc durant la transició del Tardiglacial a l’Holocè inicial. 3) Es van determinar els valors de δ18O diatom i δ13C diatom de 51 mostres de diatomees per caracteritzar els canvis ambientals a alta freqüència registrats des del Tardiglacial fins al final de l’Holocè inicial (unitat 1). Els intervals seleccionats també van ésser mostrejats per a fer observacions i descripcions petrològiques mitjançant microscopi òptic a partir de l’estudi de làmines primes. Alhora, un nombre representatiu de mostres també van ésser estudiades mitjançant el microscopi electrònic de rastreig (SEM). xiv Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles L’anàlisi d’isòtops del frústuls de les diatomees requereix que les mostres estiguin formades quasi exclusivament per diatomees, ja que altres components com poden ser argiles, carbonats o cendres volcàniques, poden alliberar isòtops que distorsionin la senyal isotòpica real. Per tant, totes les mostres van ser tractades mitjançant agents químics, per a eliminar les substàncies amb composició química diferent de les diatomees, i mitjançant processos mecànics per a separar les substàncies no diatomítiques però amb igual composició que aquestes. Un cop les mostres estaven formades per més d’un 90% de diatomees, entre 5 i 10 mg de mostra van ser processats per alliberar els isòtops d’oxigen amb el mètode de fluorinització clàssica amb passos esglaonats i, així, poder eliminar amb garanties la capa superficial hidratada que presenta l’òpal de les diatomees. Un cop l’oxigen estructural de la sílice va ser alliberat, aquest va ésser convertit en CO 2 i analitzat mitjançant un espectròmetre de masses Finnigan MAT 253 dual inlet. Per altra banda, els isòtops de carboni van ser analitzats mitjançant la combustió de mostres d’entre 1 i 2 mg en un analitzador elemental Costech ECS4010 en fase amb un espectròmetre de masses VG dual inlet. Totes les anàlisis isotòpiques realitzades sobre els sediments es van dur a terme al NERC Isotope Geosciences Laboratory (Regne Unit), a càrrec de la Prof. Melanie J Leng. La corba de grisos va ser utilitzada per obtenir, de manera objectiva, les diferències d’intensitat de colors de les làmines dels intervals estudiats. Per a la construcció de la corba de grisos es va utilitzar el paquet de software ImageJ. Finalment, per obtenir els components periòdics que podien presentar els resultats dels δ18O diatom es van realitzar anàlisis freqüencials amb els mètodes Multi-Taper (MTM) i de Temps-Freqüència. Tots aquests tractaments estadístics de les dades es van dur a terme mitjançant el paquet de software R. Evolució paleohidrològica del Lago Chungará (Altiplà Andí, nord de Xile) durant el Tardiglacial i l’Holocè inicial a partir dels isòtops d’oxigen de les diatomees En aquest capítol es presenten els resultats obtinguts a partir de la determinació de δ18O diatom i la caracterització petrogràfica dels sediments. Per això, s’han seleccionat tres intervals de sediments laminats del registre sedimentari del Lago Chungará. L’objectiu ha estat establir l’evolució paleohidrològica del llac durant el Tardiglacial i l’Holocè inicial (ca 12000-9400 anys cal BP), i alhora mostrar com aquesta evolució pot tenir un paper clau en la interpretació dels isòtops d’oxigen. L’estudi petrogràfic mitjançant làmines primes ha mostrat que els sediments laminats del Lago Chungará estan formats per ritmites compostes de làmines mil·limètriques clares i fosques. Les làmines clares estan formades quasi exclusivament per la diatomea planctònica Cylostephanos andinus, mentre que les làmines fosques són riques en matèria orgànica i contenen una mescla de diverses associacions xv Resum de diatomees. La formació de les làmines clares està relacionada amb «blooms» de diatomees de curta durada (dies o setmanes). Per contra, les làmines fosques representen les condicions limnològiques de base al llarg de diversos anys de deposició. L’anàlisi de δ18O diatom ha estat realitzada únicament a les làmines fosques, que mostren una tendència general d’enriquiment isotòpic al llarg del període estudiat. La comparació dels valors de δ18O diatom amb la reconstrucció de l’evolució del nivell del llac duta a terme en treballs previs suggereix que, a més de canvis en la relació precipitació/evaporació (P/E), l’evolució d’altres factors hidrològics locals podrien haver dominat les variacions del registre de δ18O diatom . Aquests canvis podrien ésser tant variacions de la pèrdua d’aigua subterrània com variacions de la relació superfície/volum de l’aigua del llac. La reconstrucció hidrològica del llac mostra que durant la pujada més important del nivell del llac, la qual va succeir cap els 10400 anys cal BP, es van inundar marges molt més soms, donant com a resultat canvis en la morfologia del llac. Aquests canvis van provocar un increment de la relació superfície/ volum del llac i, per tant, va augmentar l’evaporació, causant un enriquiment isotòpic de l’aigua del llac. D’altra banda, el llac, en trobar-se en els estadis inicials de la seva formació, segurament va patir modificacions en la seva hidrologia, com per exemple que la deposició de sediments va anar segellant el fons del llac, dificultant així la sortida d’aigua subterrània i disminuint el seu flux. Així, el temps de residència de l’aigua va augmentar i, alhora, va augmentar l’evaporació potencial. Per tant, ambdós factors són possibles causes de l’enriquiment isotòpic, més enllà dels factors climàtics. Els treballs previs sobre δ18O diatom han fet especial èmfasi en qüestions com la contaminació, la diagènesi i les interaccions de l’aigua intersticial amb l’òpal de les diatomees com a factors determinants en les variacions de δ18O diatom més enllà de factors climàtics i, en canvi, els factors hidrològics locals, en gran mesura, han estat negligits. Els resultats d’aquest capítol demostren la complexa interacció que existeix entre els diversos factors que intervenen en els registres de δ18O diatom dels llacs hidrològicament tancats i com la seva interpretació necessita ser adaptada als diferents estadis d’evolució del llac. Les senyals de l’ENSO i de l’activitat solar a partir dels isòtops d’oxigen de la sílice de les diatomees durant la transició Tardiglacial-Holocè inicial als Andes Centrals (18ºS) Aquest capítol mostra, per primer cop, la reconstrucció del balanç hídric a l’Altiplà Andí a escala de desenes i centenes d’anys basada en δ18O diatom durant la transició Tardiglacial-Holocè inicial (11900- 11450 anys cal BP). Les característiques texturals de les làmines estan explicades als capítols 4 i 6. Per a realitzar aquest estudi es van analitzar 40 mostres de làmines fosques consecutives de l’interval estudiat. xvi Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles Els resultats obtinguts mostren una sèrie d’esdeveniments secs i humits a escala de desenes i centenars d’anys. Els episodis àrids corresponen a valors d’isotopia alts, mentre que els episodis humits estan indicats per valors isotòpics més baixos, degut a que el senyal isotòpic esta directament relacionat amb les variacions de la P/E a l’Altiplà. Les dades d’isotopia s’han comparat amb reconstruccions prèvies realitzades en el mateix sondatge amb la intenció de confirmar aquesta interpretació. Aquestes reconstruccions prèvies reflecteixen l’entrada d’elements terrígens i el balanç hídric al Lago Chungará. El resultat de la comparació ha estat satisfactori, mostrant una bona correlació entre les diferents reconstruccions. Tot i això, existeix un desfasament temporal sistemàtic entre elles. Aquest desfasament ha estat interpretat com que és degut al temps necessari perquè s’observin canvis significatius en els valors d’isòtops d’oxigen de l’aigua del llac i la seva posterior incorporació als frústuls de les diatomees, el qual és diferent al temps de resposta de les altres dues reconstruccions. Aquest fet permet destacar la resposta no lineal que sovint mostren els ecosistemes lacustres respecte als forçaments ambientals. A l’interval estudiat s’han pogut identificar dues caigudes principals dels valors d’isotopia (11800 i 11550 anys cal BP) que estarien indicant unes majors condicions humides a escala de centenars d’anys. Al mateix temps, també s’ha pogut identificar un enriquiment isotòpic principal, per sobre dels nivells de base, que es localitza entre els 11990 i els 11550 anys cal BP i que indicarien una fase d’aridesa breu durant el Tardiglacial. El trànsit Tardiglacial-Holocè inicial també conté diverses caigudes de menor magnitud dels valors d’isotopia i que estarien associades a esdeveniments de major humitat a una escala temporal inferior (desenes d’anys). Per altra banda, també s’han realitzat anàlisis espectrals dels valors obtinguts de δ18O diatom , els quals mostren que, durant els esdeveniments amb més humitat, els canvis en les condicions atmosfèriques a l’Altiplà Andí estarien relacionats tant amb l’ENSO com amb l’activitat solar. Amb l’ajuda de les anàlisis espectrals s’han identificat freqüències de 7-9 anys i 15-17 anys, les quals es corresponen amb freqüències significatives del fenomen ENSO. Al mateix temps, també han estat identificades periodicitats de 11, 13 i 35 anys corresponents als cicles d’activitat solar Schwabe, Hale i Brückner, respectivament. El treball realitzat ha permès establir una relació entre l’activitat solar i l’ENSO a escala de desenes d’anys i superiors, i per tant és molt probable que el forçament, a causa de l’activitat solar al registre del Lago Chungará, sigui transmès mitjançant la modulació ENSO dels monsons d’Amèrica del Sud. Per altra banda, l’anàlisi de Temps-Freqüència duta a terme mostra que, encara que els forçaments per l’activitat solar i per l’ENSO van estar presents durant l’inici de l’Holocè, van ésser més intensos durant el Tardiglacial. Com s’observa a les anàlisis espectrals, probablement l’Holocè inicial va estar dominat per condicions de La Niña, que corresponen amb esdeveniments humits sobre l’Altiplà Andí. Molts estudis han demostrat forçaments ENSO durant la transició Glacial-Interglacial, però els resultats presentats xvii Resum aquí mostren que aquest registre és una de les poques seqüències terrestres que preserva freqüències clau de l’ENSO i, per tant, està demostrant que aquests processos climàtics estan dominant la transició cap a l’Holocè. Processos biogeoquímics que controlen els isòtops d’oxigen i carboni de la sílice de les diatomees presents en ritmites lacustres L’objectiu d’aquest capítol ha estat la reconstrucció dels processos biològics, químics i sedimentaris que van donar lloc a la formació de les làmines del registre sedimentari del Lago Chungará. Al mateix temps, amb aquesta reconstrucció s’ha pogut demostrar com els cicles biogeoquímics i els registres sedimentaris dels llacs estan directa o indirectament relacionats amb el control climàtic de la hidrologia i amb els processos que tenen lloc a la conca de captació d’un llac. Per aconseguir aquest objectiu s’han utilitzat: 1) els valors de δ18O diatom a cadascuna de les làmines de l’interval seleccionat (11990-11530 anys cal BP) i 2) els valors de δ13C diatom en 11 mostres, escollides segons la seva representativitat. D’altra banda, també s’ha dut a terme un estudi petrogràfic més detallat que al capítol 4. Com es comenta al capítol 4, el registre laminat del Lago Chungará està format principalment per ritmites mil·limètriques clares i fosques de caràcter multianual. No obstant, l’estudi petrogràfic més detallat ha permès identificar una làmina intermèdia de color verd clar localitzada sempre entre una làmina clara i una fosca. Les làmines clares es van formar durant «blooms» extraordinaris de curta durada i estan bàsicament formades per Cyclostephanos andinus de mida gran (>50 μm). Aquests «blooms» extraordinaris de diatomees es van produir durant esdeveniments de gran mescla de la columna d’aigua i/o per episodis d’erosió excepcional de la conca de drenatge. En el primer cas l’ascens de les aigües riques en nutrients procedents de l’hipolimnion van permetre més disponibilitat de nutrients, mentre que en el segon cas, la font de nutrients seria al·lòctona. Les làmines de color fosc es van acumular al llarg de diversos anys sota diferents condicions de la columna d’aigua. Aquestes làmines estan formades per una barreja de diatomees, principalment teques més petites de Cyclostephanos andinus i per Discostella stelligera, i matèria orgànica. Aquestes làmines fosques reflecteixen el pas progressiu des de les condicions favorables per a que tinguin lloc els «blooms» extraordinaris de diatomees cap a les condicions de base del llac. Per últim, les làmines de color verd clar estan formades per components de les làmines clares que van canviant progressivament cap a la part superior a components de les làmines fosques. Aquestes làmines de color verd clar sempre es troben immediatament després d’una làmina clara, i podrien ser conseqüència d’un fenomen anomenat «auto-sedimentació» que indicaria la fi dels «blooms» excepcionals. Per tant, s’ha establert un nou model deposicional que, en aquest cas, estaria xviii Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles format per triplets compostos d’una làmina clara a la part inferior, una làmina verda clara en posició intermèdia i una làmina fosca a la part superior. Els valors de δ18O diatom i δ13C diatom a les ritmites s’han utilitzat per detectar oscil·lacions paleoclimàtiques i paleoambientals, ja que els valors d’isotopia han estat interpretats com variacions en la relació P/E i en la concentració de carboni dissolt a l’aigua del llac, respectivament. Els valors de δ18O diatom mostren que, majoritàriament, la formació de làmines clares va ser induïda per caigudes del nivell del llac, mentre que la formació de làmines fosques es va veure afavorida per pujades del nivell del llac. Aquestes dades d’isotopia, junt amb les dades cronoestratigràfiques, suggereixen que la deposició de les làmines clares va ésser provocada per esdeveniments ambientals extrems. L’ENSO i l’activitat solar són els principals forçaments climàtics que podrien estar desencadenant aquests esdeveniments, ja que afavoreixen els canvis hidrològics a escala interanual i de desenes d’anys a l’Altiplà Andí. Les pertorbacions ambientals d’alta freqüència, induïdes per esdeveniments climàtics d’escala interanual o de desenes d’anys, difícilment queden registrades als sediments lacustres de l’Altiplà Andí. Tanmateix, els sediments laminats del Lago Chungará s’han mostrat com un bon registre per reflectir la intensitat de diversos fenòmens climàtics d’alta freqüència i els seus efectes en el cicle del carboni dels llacs. Els registres isòtopics de δ18O diatom i δ13C diatom de la unitat laminada del Lago Chungará (dels 12400 als 8400 anys cal BP) En aquest capítol es presenten els registres isotòpics de δ18O diatom i δ13C diatom de la unitat sedimentària laminada i rica en diatomees del Lago Chungará, la qual abasta dels 12400 als 8400 anys cal BP. Les dades de l’anàlisi provenen de 51 mostres. L’objectiu ha estat caracteritzar els canvis ambientals i climàtics que van succeir als Andes Centrals durant el període de temps estudiat. El registre de δ18O diatom s’ha interpretat com indicador de la hidrologia del llac i del balanç d’humitat regional a escala de centenars i milers d’anys, mentre que les variacions de δ13C diatom és molt probable que estiguessin condicionades per canvis en la productivitat del fitoplàncton i en la concentració i origen del diòxid de carboni dissolt a l’aigua (CO 2(aq) ). Han estat identificades tres fases climàtiques principals durant el Tardiglacial i l’Holocè inicial: 1) una fase humida durant la transició Tardiglacial-Holocè inicial (12400-10100 anys cal BP), 2) una fase seca a l’Holocè inicial (10100-9100 anys cal BP), i 3) una fase progressivament més humida durant l’última part de l’Holocè inicial (9100-8400 anys cal BP). Encara que la fase humida de la transició Tardiglacial-Holocè inicial coincideix amb una tendència cap a un mínim en la insolació, l’establiment de condicions climàtiques semblants a les de La Niña a la zona tropical de l’Oceà Pacífic s’imposarien al forçament per precessió. Per altra banda, durant aquest xix Resum període no es va registrar al Lago Chungará cap esdeveniment equivalent al Younger Dryas de l’hemisferi nord. El mínim d’insolació als ca. 10000 anys cal BP va tenir un paper clau durant l’Holocè inicial (10100- 9100 anys cal BP), provocant la migració cap al Nord de la zona de convergència intertropical (ITCZ). Aquesta, junt amb la debilitació de les condicions ENSO, va sotmetre els Andes tropicals a un període de sequera extrema. El retorn cap a condicions més humides va tenir lloc al voltant dels 9000 anys cal BP seguint un increment en la insolació estival austral. Aquest darrer període va haver de ser breu, ja que molts registres climàtics de l’Altiplà Andí mostren l’inici d’un important període sec al voltant dels 8000 anys cal BP com a conseqüència d’un afebliment de les condicions de l’ENSO. Tot i això, es requereixen més anàlisis en tota la unitat no laminada (unitat 2) del registre del Lago Chungará per poder confirmar aquesta hipòtesi. L’anàlisi isotòpica de la matèria orgànica inclosa dins de les parets dels frústuls de les diatomees en sediments lacustres ha demostrat que diversos factors, com la productivitat fitoplanctònica o l’origen del carboni i llur concentració, interaccionen donant com a resultat oscil·lacions en els valors de δ13C diatom . Aquestes oscil·lacions també revelen interaccions entre la reserva de carboni de l’aigua del llac i la conca de drenatge. Durant els períodes humits els valors de δ13C diatom mostren que la contribució de materials externs, a més de la productivitat fitoplanctònica del llac, van incrementar significativament la reserva de carboni de l’aigua del llac, mentre que els períodes secs van afavorir l’acumulació de carboni format al mateix llac i el posterior enriquiment dels valors de δ13C diatom . L’ alliberament del CO 2 de l’hipolímnion durant els períodes de mescla també va poder causar una baixada dels valors de δ13C diatom . Així doncs, les mesures de δ18O diatom i δ13C diatom s’han mostrat com a eines molt útils per entendre els patrons climàtics regionals i els processos d’interacció entre el llac i la seva conca de drenatge. Aquest capítol posa èmfasi en que l’anàlisi conjunta de δ18O diatom i δ13C diatom poden ajudar a entendre millor el paper dels llacs en el cicle del carboni a escala global, sobretot dins del context del canvi global actual. Conclusions En aquesta tesi s’han obtingut conclusions de caire metodològic, limnològic i climàtic: - Diversos factors ambientals, més enllà dels forçaments climàtics, poden influir en els valors de δ18O diatom . Els registres de δ18O diatom en sistemes lacustres tancats no poden ser simplement interpretats en termes de sec o humit, sinó que és imperatiu entendre la hidrologia de cada sistema. Per la seva part, l’anàlisi de δ13C diatom ha demostrat que aquesta tècnica és una eina vàlida per a fer reconstruccions del cicle del carboni als llacs, així com per donar un millor punt de vista del cicle del carboni a nivell global. - La unitat sedimentària laminada del Lago Chungará està formada per ritmites multianuals compostes per triplets de làmines de color clar, verd clar i fosc. Aquestes làmines són riques en diatomees xx Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles i són el resultat de processos lacustres diferents. Les làmines de color clar es van formar com a conseqüència de «blooms» extraordinaris de molt curta durada (dies o setmanes) que van tenir lloc durant episodis de turbulència extrema i/o episodis d’erosió excepcional de la conca de recepció. Les làmines de color verd clar es van formar com a resultat del final dels «blooms» extraordinaris, i per últim, les làmines de color fosc es van dipositar al llarg de diversos anys sota diferents condicions de la columna d’aigua i, per tant, representen les condicions de base del llac. - Els valors de δ18O diatom ens mostren que tant les làmines clares com les làmines fosques es poden formar en períodes secs i en períodes humits. Tot i així, aquest valors també mostren que els «blooms» extraordinaris van ésser més intensos amb condicions de baix nivell del llac. En la majoria de casos la formació de làmines clares es va veure afavorida per baixades del nivell del llac, mentre que la formació de làmines fosques es va veure especialment induïda per pujades del nivell del llac. Al mateix temps, els valors de δ13C diatom indiquen que la disponibilitat de carboni va ser superior durant el «blooms» extraordinaris de diatomees. La combinació d’ambdós registres ha destacat les complexes relacions entre els processos limnològics, els processos de la conca de drenatge, la hidrologia i els forçaments climàtics. Durant els períodes amb més precipitació indicats pels valors baixos de δ18O diatom , la contribució relativa del carboni extern, a més de la productivitat fitoplanctònica, van incrementar les reserves de carboni total al llac, tal i com indiquen els valors de δ13C diatom . Per contra, els períodes secs (valors elevats de δ18O diatom ) van afavorir els processos d’acumulació de carboni format al propi llac, com demostren els valors elevats de δ13C diatom . - El registre d’isotopia de la unitat laminada del Lago Chungará exposa clarament que, segons l’escala temporal, un tipus de forçament pot dominar sobre els altres en la interpretació del valors de δ18O diatom i δ13C diatom . - A escales temporals de centenars i milers d’anys, factors hidrològics com són canvis en la relació de la pèrdua d’aigua pel flux subterrani o per evaporació i/o en l’extensió del llac poden jugar un paper molt important. Tot i això, l’evolució temporal a més baixa freqüència de δ18O diatom a la unitat laminada del Lago Chungará posa de manifest els principals canvis en el balanç hídric regional, els quals estarien relacionats amb forçaments orbitals i amb condicions similars a les que es donen durant els fenòmens ENSO. - Per altra banda, a escales de temps de desenes d’anys, els valors de δ18O diatom estarien induïts per fenòmens climàtics d’altra freqüència. Les oscil·lacions en els valors de δ18O diatom , que indiquen canvis en el balanç de precipitació/evaporació a la regió del Lago Chungará, serien el resultat de canvis en les condicions atmosfèriques sobre l’Altiplà Andí, provocats tant per l’ENSO com per l’activitat solar. Per tant, la transició Glacial-Interglacial, va estar dominada per canvis abruptes en les condicions climàtiques i no pas per un canvi progressiu d’aquestes condicions. xxi Resum Contents Agraïments ................................................................................................................... v Resum .......................................................................................................................... x Chapter 1: Introduction ................................................................................................. 1 1.1.- Lakes ........................................................................................................................................................ 3 1.1.1.- Lake types .................................................................................................................................. 3 1.1.2.- Physical, chemical and biological characteristics of lakes ...................................................... 4 1.1.3.- Lacustrine sediments ............................................................................................................... 8 1.2.- Diatoms .................................................................................................................................................. 13 1.2.1.- The diatom frustule .................................................................................................................. 14 1.2.2.- Diatom ecology ........................................................................................................................ 15 1.3.- Isotopes .................................................................................................................................................. 17 1.3.1.- Stable isotopes in palaeoenvironmental research ................................................................... 18 1.3.2.- Stable oxygen isotopes ............................................................................................................ 22 1.3.3.- Stable carbon isotopes ............................................................................................................ 24 1.4.- Aims ...................................................................................................................................................... 26 1.5.- Thesis structure ..................................................................................................................................... 26 Chapter 2: Geological, geographical, limnological and climate framework ................. 27 2.1.- Geographical and geological setting ..................................................................................................... 27 2.1.1.- The Central Andes ................................................................................................................... 28 2.2.- Climate framework ............................................................................................................................... 30 2.2.1.- Modern climate ....................................................................................................................... 31 2.2.2.- Last Glacial-Holocene climate variability in the Central Andes ............................................ 34 2.3.- Limnological features of Lago Chungará .............................................................................................. 37 2.4.- Earlier work undertaken at Lago Chungará ......................................................................................... 40 2.4.1.- Sedimentary record ................................................................................................................ 40 2.4.2.- Chronological framework ...................................................................................................... 44 xxiii Chapter 3: Methods ..................................................................................................... 49 3.1.- Hydrochemical and isotopic water analyses ......................................................................................... 49 3.2.- Sediment sampling ............................................................................................................................... 49 3.3.- Light and Scanning Electron Microscope sample preparation ............................................................. 51 3.4.- Isotope analyses in sediments ............................................................................................................... 51 3.4.1.- Cleaning of diatom frustules .................................................................................................... 51 3.4.2.- Oxygen iostope extraction ...................................................................................................... 54 3.4.3.- Analyses of δ13C diatom and %C diatom ............................................................................................ 56 3.5.- Statistical analyses and Grey-colour curve ........................................................................................... 57 Chapter 4: The palaeohydrological evolution of Lago Chungará (Andean Altiplano, northern Chile) during the Late Glacial and Early Holocene using oxygen isotopes in diatom silica ........................ 59 4.1.- Introduction .......................................................................................................................................... 59 4.2.- Results: Petrography and isotope composition of diatoms .................................................................. 60 4.2.1.- Interval 1 (11,990 - 11,530 cal years BP) ................................................................................. 60 4.2.2.- Interval 2 (10,430 - 10,260 cal years BP) ............................................................................... 61 4.2.3.- Interval 3 (9,890 - 9430 cal years BP) ................................................................................... 62 4.3.- Discussion ............................................................................................................................................. 64 4.3.1.- The sedimentary model of diatom rhythmites ....................................................................... 64 4.3.2.- Lake level and δ18O diatom changes ............................................................................................. 64 4.4.- Conclusions ........................................................................................................................................... 67 Chapter 5: ENSO and solar activity signals from oxygen isotopes in diatom silica during the Late Glacial-Holocene transition in Central Andes (18ºS) .............................................................................. 69 5.1.- Introduction .......................................................................................................................................... 69 5.2.- Results .................................................................................................................................................. 70 5.2.1.- Oxygen isotopes ...................................................................................................................... 70 5.2.2.- Spectral analyses of the diatom oxygen isotope record .......................................................... 71 5.3.- Discussion ............................................................................................................................................. 73 5.3.1.- Controlling factors of δ18O diatom in Lago Chungará ................................................................... 73 5.3.2.- Variation of the Precipitation/Evaporation balance in the lake ............................................ 75 5.3.3.- Long-term, centennial- to millennial-scale palaeoclimatic implications ............................... 77 5.3.4.- Short-term, decadal- to centennial-scale palaeoclimatic implications .................................. 78 5.4.- Conclusions ........................................................................................................................................... 79 xxiv Chapter 6: Biogeochemical processes controlling oxygen and carbon isotopes of diatom silica in lacustrine rhythmites ...................................... 81 6.1.- Introduction ...........................................................................................................................................81 6.2.- Results .................................................................................................................................................. 82 6.2.1.- Laminae biogenic composition .............................................................................................. 82 6.2.2.- Laminae isotopic composition ............................................................................................... 83 6.3.- Discussion ............................................................................................................................................. 88 6.3.1.- Biological and sedimentary processes forming rhythmites ................................................... 88 6.3.2.- δ18O diatom and δ13C diatom interpretation ..................................................................................... 89 6.3.3.- δ18O diatom inter-cycle relationships (white laminae formation) ............................................... 91 6.3.4.- δ18O diatom and δ13C diatom intra-cycle relationships (green laminae formation) .......................... 91 6.3.5.- Climatic forcing of the laminae formation ............................................................................. 93 6.4.- Conclusions ........................................................................................................................................... 94 Chapter 7: Oxygen and carbon diatom isotope records from the Lago Chungará laminated unit (12,400 to 8,400 cal years BP) .............. 95 7.1.- Introduction .......................................................................................................................................... 95 7.2.- Results ................................................................................................................................................... 96 7.3.- Discussion ............................................................................................................................................. 97 7.3.1.- Controlling factors on δ18O diatom and δ13C diatom ......................................................................... 97 7.3.2.- Palaeoenvironmental reconstructions ................................................................................. 100 7.4.- Conclusions ......................................................................................................................................... 104 Chapter 8: Conclusions .............................................................................................. 107 8.1.- Concluding remarks ............................................................................................................................. 107 8.1.1.- Methodological conclusions .................................................................................................. 107 8.1.2.- Limnological conclusions ..................................................................................................... 108 8.1.3.- Climate conclusions ............................................................................................................. 109 8.2.- Perspectives and future work .............................................................................................................. 110 Bibliography .............................................................................................................. 111 Appendices ................................................................................................................ 125 Glossary ........................................................................................................................................................ 125 Final isotope data ......................................................................................................................................... 131 Papers .......................................................................................................................................................... 141 xxv 1Chapter 1 Introduction Stable isotopes from marine and lacustrine sedimentary records have been used to improve our understanding of the evolution of the Earth since the end of World War II (e.g. McCrea, 1950; Urey et al. 1951, Emilliani, 1955). Oxygen and carbon isotopes in microfossil carbonates have been widely used to carry out palaeoenvironmental reconstructions (see reviews in Ito, 2001; Richardson, 2001; Schwalb, 2003; Leng and Marshall, 2004; Grottoli and Eakin, 2007; Ravelo and Hillarie-Marcel, 2007). However, calcareous microfossils (mainly foraminifera, molluscs and ostracods) are not always present in marine or lacustrine sediments owing to unfavourable ecological or post-depositional conditions. These non- carbonated sediments sometimes contain abundant biogenic silica (mainly diatoms), rendering them suitable for studies of stable isotopes (e.g. Shemesh et al. 1992, 1995, 2001; Rietti-Shati et al. 1998; Rosqvist et al. 1999, 2004, Barker et al. 2001, 2007; Rioual et al. 2001; Leng et al. 2001, 2005a; Hu and Shemesh, 2003; Jones et al. 2004; Lamb et al. 2005; Polissar et al. 2006; Tyler et al. 2008; Mackay, 2007; Mackay et al. 2008; Swann et al. 2008, 2010; Jonsson et al. 2010). For this reason, considerable progress has been made in the study of biogenic silica using isotopes in recent years (see Leng and Barker, 2006; Crosta and Koç, 2007, and Swann and Leng, 2009 for reviews). In lacustrine sedimentary materials changes in isotope oxygen values (from both carbonates and biogenic silica) are usually related to changes in temperature or isotope composition of the lake water (δ18O lakewater ) (Leng and Marshall, 2004; Leng and Barker, 2006). On the other hand, carbon isotopes in bulk organic matter (δ13C bulk ) are commonly used to demonstrate changes in carbon cycle and productivity which are often climatically induced (Meyers and Teranes, 2001; Leng et al. 2005b). Using oxygen and carbon isotope ratios in palaeoenvironmental reconstruction is, however, not easy, given that the sedimentary record can be influenced by a wide range of interlinked environmental processes ranging from local hydrology to regional climate change. Diatom oxygen isotopes (δ18O diatom ) are increasingly being used for palaeoenvironmental reconstructions in lacustrine sedimentary records (see Leng and Barker, 2006 for a detailed review). Although they have been applied in long-scale palaeoclimatic reconstructions (e.g. Rietti-Shati et al. 1998; Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles Rosqvist et al. 1999; Rioual et al. 2001; Barker et al. 2007; Mackay, 2007; Swann et al. 2010), they can also be used in many other fields. On the other hand, considerable progress has been made in the use of organic inclusions within the diatom frustule from marine sedimentary records in recent years (Singer and Shemesh, 1995; Crosta and Shemesh, 2002; Des Combes et al. 2008). However, very little systematic work has been undertaken on the analysis of the diatom carbon isotope composition (δ13C diatom ) in lacustrine sediments (Hurrell et al. submitted). Tropical proxy records offer valuable insights into past climate and environmental changes of the Earth and into possible future climate change scenarios (Thompson et al. 1998; Barker et al. 2001). The tropics are implicated in forcing global climate shifts at interannual to orbital time scales (Gasse, 2000; Timmermann et al. 2007; Wanner et al. 2008; Chiang, 2009). The global impact of the El Niño–Southern Oscillation (ENSO) variability is one example of how climate phenomena can propagate rapid changes worldwide (Placzek et al. 2006, Merkel et al. 2010). Furthermore, high-frequency climate variability often exerts a considerable influence on human activities (i.e. monsoons, El Niño) in low latitudes. Thus, it is essential to investigate the past tropical climate variability in order to better understand the regional and global climate changes (Thompson et al. 2005). Research into tropical regions has therefore become a key issue among palaeoclimatologists. The tropics of South America host the world’s largest river, and have higher total terrestrial primary productivity and biological diversity than any other continental region of comparable size (Rigsby et al. 2005). Influenced by the tropical circulation in the north, and by the mid-latitude westerlies in the south, the Central Andes are an ideal site to study past variations of atmospheric circulation systems. Thus, the Andean Altiplano has become a key region for the study of late Quaternary climate change in South America. Although this region is very arid today, it has shown significant moisture changes during the Late Glacial and Holocene. Lake sediments, ice cores, pollen profiles, tree rings, glacial deposits together with modeling studies are important sources of information about past environmental conditions in this high-altitude mountain region (e.g. Thompson et al. 1998; Baker et al. 2001a; Garreaud et al. 2003; Goslin et al. 2008; Kull et al. 2008; Solíz et al. 2009; Villalba et al., 2009). Sedimentary records of high-altitude Andean Altiplano lakes are good candidates for carrying out oxygen isotope studies to reconstruct the late Quaternary climatology of the region (Valero-Garcés et al. 2000). They usually preserve an excellent centennial- to millennial-scale record of effective moisture fluctuations and source changes during the Late Glacial and Holocene despite the fact that the interpretation is not always straightforward (e.g. Baker et al. 2001a; Grosjean et al. 2001; Fritz et al. 2004). Oxygen and carbon isotope analyses in carbonates (δ18O carbonate and δ13C carbonate ) and δ13C bulk have been successfully used to reconstruct the hydrological responses to climate change in different Andean lacustrine systems to date (e.g. Schwalb et al. 1999; Valero-Garcés et al. 1999; Seltzer et al. 2000; Wolfe 2 3et al. 2001; Abbott et al. 2003; Rowe et al. 2003). No attempt, however, has been made to use δ18O diatom and δ13C diatom despite the fact that they are usually the best preserved fossils in the sedimentary record of the Andean Altiplano lakes. 1.1 Lakes Although lakes constitute at present about 1% of the Earth’s continental surface, containing less than 0.02% of the water in the hydrosphere, their importance is far greater than these meagre figures suggest (Talbot and Allen, 1996; Wetzel, 2001). Lakes and lake deposits have been the subject of many studies owing to their economic importance (e.g. Robins, 1983; Tiercelin, 1991). Many modern lakes are a vital source of food and water, and the sediments of some of them constitute a source of valuable minerals (i.e. borax) (Swihart et al. 1996). Thus, the question ‘What is a lake?’ may be posed. According to the Encyclopedia Brittanica (1962), a lake is a mass of still water situated in a depression of the ground without direct communication with the sea. Natural lakes are located in depressions such as ponds and include those that have not resulted from the construction of dams such as impoundments and reservoirs (Meybeck, 1995). 1.1.1 Lake types Lakes can be classified according to their origin (Wetzel, 2001), chemistry (Eugster and Hardie, 1995), mixing periods (Lewis, 1983), temperature (Hutchinson and Löffler, 1956) and trophic state (Hutchinson, 1969). A satisfactory classification would be based on their origin (Margalef, 1983). Although lakes may be due to a variety of natural processes such as landslides, rivers, dissolutions, or even meteoritic impacts, the most usual origins are volcanic, tectonic and glacial (Wetzel, 2001). Lakes of volcanic origin may be formed by lava flows that obstruct a river valley, forming a new lacustrine basin (e.g. Lake Yojoa, Honduras) (Devevey et al. 1993) or by crater explosion and collapse forming a depression that fills up with rainwater (e.g. Holzmaar, Germany) (Zolitschka, 1992). Large lakes primarily tectonic in origin can be grouped into: a) lakes that are formed in extensional rift valleys (e.g Lake Baikal, Russia) (Colman et al. 1995), b) associated with pull-apart basins (e.g. Lake Issyk-Kul, Rep Kyrgyzstan) (De Batist et al. 2001), or c) formed on slowly subsiding sags in cratonic areas (e.g Lake Chad, Rep Chad) (Durand, 1982). Finally, lakes in glaciated regions are generally small and may be proglacial (e.g Lake Malaspina, USA) (Gustavson, 1975) caused by ice-damming, by barriers composed of moraine (e.g. Imja Lake, Nepal) (Chikita et al. 2000), or by ice scour, freeze-thaw and by valley glaciation (e.g. Pitt Lake, Canada) (Ashley and Moritz, 1979) (Fig. 1.1). Introduction Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 1.1.2 Physical, Chemical and biological characteristics of lakes Physical features Physical processes in lakes are important because they are linked to external environmental forcing events such as climate (Cohen, 2003). For instance, lake level oscillations are mainly attributed to water outputs and inputs that are commonly controlled by climate. In addition, light and heat penetration regulate the distribution of organisms and the mixing of the water column leaving their signal in the sediment characteristics (Imboden and Wüest, 1995). Inputs and outputs of lake waters in combination with the morphometry of the lake basin determine the lake level. Water inputs to the lake include rainfall, surface runoff and groundwater inflow whereas outflows include surface outflow, evaporation and groundwater loss (Talbot and Allen, 1996; Cohen, Figure 1.1. Images of different types of lakes according to their origin. A) Lake Yojoa (Honduras) is an example of a lake formed by a lava flow that obstructed a river valley; B) Lake Holzmaar (Germany) resulted from volcanic activity. After a crater explosion and collapse, the lake filled up with rainwater; C) Lake Baikal (Russia), lake of tectonic origin formed in an ancient rift valley. It is the world’s deepest lake; D) Lake Issyk-Kul (Rep Kyrgyzstan) is an example of a lake of tectonic origin associated with a pull-apart basin; E) other lakes of tectonic origin are those formed on slowly subsiding sags in cratonic areas, such as Lake Chad (Rep Chad); F) Lake Malaspina (USA) is an example of a lake of proglacial origin; G) glacial lakes can also be formed by ice-damming, i.e. barriers composed of a moraine (e.g. Imja Lake, Nepal); and H) Pitt Lake (Canada) is placed in a valley of glacial origin. 4 A D G H E F B C 52003; Darling et al. 2005; Leng and Barker, 2006). If water inputs and outputs of the lake are equal over a short time span the lake water level remains constant and can then be considered as an open lake. By contrast if significant variations in the lake level occur as the ratio of inputs and outputs changes, the lake is closed (Street-Perrott and Harrison, 1985). The depth of light penetration into the lake water and the manner in which it is absorbed determine the distribution of the organisms and heat in the lake (Dehaan, 2003). It has recently been suggested that productivity in a large proportion of the world’s unproductive lakes is limited by light and not by nutrients as is commonly believed (Karlsson et al. 2009). Lakes can be divided into two zones: those that are penetrated by sufficient light to allow photosynthesis (photic zone), and those that lie below these (aphotic zone). Between these two zones is a zone where photosynthesis equals respiration, which is known as the compensation depth (Brönmark and Hanson, 2005). Light penetration varies because of differences in suspended sediment (turbidity), plankton and dissolved solid content (Cohen, 2003). On the other hand, heat enters lakes as solar radiation and from geothermal sources. Solar radiation penetrates the lake and heats the water, but the penetration is exponentially reduced with depth (Ragotzkie, 1978). Heat is irregularly distributed in the water column, and is redistributed by mixing. Mixing can be induced by wind, density inestabilities and/or turbulent water inflows when the external forcings are strong enough to counteract the density differences (Imboden and Wüest, 1995). When turbulent mixing or density instabilities are not strong enough to mix the water column vertically, this can become stratified. Stratification in a lake is also influenced by morphometry, depth, solute concentrations, the temperature of the atmosphere, solar irradiation and wind speed (Lewis, 1983). Stratification results from the high temperature differences between surface and bottom waters. This stratification separates the water masses by a thermocline. The surface water mass is termed the epilimnion and the bottom water mass is known as the hypolimnion (Fig. 1.2). Vertical mixing and stratification in the water column of the lake regulate the distribution of dissolved gases and nutrients in various temporal cycles (Cohen, 2003). Lake mixing can be seasonal or occur at longer temporal scales throughout changes in the thermal structure. Lakes can be classified according to their mixing regimes as amictic (never mix), monomictic cold and warm (mix once per year), dimictic (mix twice per year) or polymictic cold and warm (mix repeatedly throughout year) (Hutchinson and Löffler, 1956; Lewis, 1983). Lakes that are only partially mixed are termed meromictic. Chemical features Chemical processes in lakes are closely related to external climate and watershed factors. Solute concentrations regulate the distribution of organisms, and the precipitation or dissolution of mineral Introduction Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles phases (Chalie and Gasse, 2002). Non-marine waters are dominated by four cations, calcium, magnesium, sodium and potassium, and three anions, bicarbonate, sulphate and chloride. Silica can also be a significant dissolved component in alkaline lakes. The relative proportions of these and other ions are largely determined by the geology of the catchment (Eugster and Hardie, 1995). Dissolved oxygen is a very important component of the water chemistry since it is a product of the photosynthesis and is responsible for oxidation processes. Oxygen concentrations in the water column are a function of productivity, respiration, temperature and mixing, which are all linked to climate and morphometry (Martin et al. 1998). Reduction and oxidation reactions depend on the capacity of the water to oxidise or reduce, and oxygen is one of the most prominent oxidising agents present in lakes. On the other hand, for reduction to occur the absence of oxygen is as necessary as the availabilty of a reducing agent (Davison, 1993; Hamilton-Taylor and Davison, 1995; Hamilton-Taylor et al. 2007). Both oxygen content and redox conditions affect the distribution of organisms, bioturbation and the precipitation or dissolution of minerals (Cohen, 2003). Nutrients are chemical elements or simple compounds that are essential for organisms. C, which is an important nutrient in lake waters, occurs in various organic and inorganic forms. The lacustrine carbon cycle is related to the fixation of these carbon forms in the water column according to productivity or pH conditions (Cole et al. 2007). C can enter the lake as atmospheric carbon dioxide (CO 2 ), which is fixed by photosynthesisers (authocthonous carbon), or as a result of the degradation of terrestrial organic matter (allochthonous carbon). It can also enter as HCO 3 - through ground and surface waters from the THERMAL STRATIFICATION TEMPERATURE (ºC) EPILIMNION METALIMNION HYPOLIMNION 0 4 10 20 30 Figure 1.2. Thermal stratification in lakes. Deep lakes generally become physically stratified into three identifiable layers, known as the epilimnion, metalimnion, and hypolimnion. The epilimnion is the upper, warm layer, and is typically well mixed. Below the epilimnion is the metalimnion or thermocline region, a layer of water in which the temperature declines rapidly with depth. The hypolimnion is the bottom layer of colder water that is separated from the epilimnion by the metalimnion. The change of density in the metalimnion acts as a physical barrier that prevents mixing of the upper and lower layers for several months in summer (modified from Horne and Goldman, 1994) 6 7catchment (Brönmark and Hanson, 2005). CO 2 , which plays a key role in photosynthesis and in other areas of the lacustrine carbon cycle, is one of the most important drivers of climate change (Cole et al. 1994; Dean and Gorham, 1998). Other nutrients such as P, N and Si are present in small concentrations in lake waters and may therefore limit the growth of lacustrine organisms (Horne and Goldman, 1994; Wetzel, 2001; Cohen, 2003). The atomic relationship C:N:P is considered to be 106:6:1 in plankton, conforming to the so- called Redfield ratio (Reynolds, 2006). Departure from this ratio imposes nutrient limitations on plankton growth. If the N:P in the water is higher than the Redfield ratio, phytoplankton will be phosphorous limited. The opposite is true if N:P is lower. C, N and P as well as the Si needed for diatom growth are thus considered the major nutrient elements for primary productivity in lakes (Margalef, 1983). P has traditionally been considered the main growth-limiting nutrient for algae in most lakes, and therefore the main determinant of their productivity levels (e.g. Wetzel, 2001). Given that the «phosphorous limitation paradigm» is controversial, lake productivity co-limitation by multiple nutrients is today considered to be more the rule than the exception (Sterken, 2008). P enters the lake as a result of weathering of the catchment via sediment release or by atmospheric deposition (Brönmak and Hansson, 2005). Sediments are in general richer in P than lake water. Depending on pH, redox conditions, and Fe concentrations, P can be released to the epilimnion, prompting primary productivity (Cohen, 2003). N enters the lake by precipitation, by input from surface and groundwater drainage, but also by fixation of atmospheric hydrogen (Brönmak and Hansson, 2005). NO 3 - is an abundant source of N for phytoplankton, whereas NH 4 + and NO 2 - are present in much lower quantities. When P is not a limiting factor, such as in eutrophic lakes, the lake can be limited by N. Under these circumstances, organisms capable of N fixation, such as cyanobacteria, become dominant in the plankton. Si, which is required by diatoms for wall silicification, is mainly taking up by diatoms as dissolved Si(OH) 4 to form silica (Willén, 1991). Although a large amount of silica in lakes is supplemented with the inflowing waters, the rate of uptake by diatoms is so high that it can fall to unmeasurable values during the period of summer stratification (Harris, 1986). The availability of these nutrients is linked to watershed characteristics and internal mixing processes, which in turn regulates the primary productivity of lakes. The trophic status is detemined by lake productivity (Hutchinson, 1969). Infertile soils release relatively little N and P leading to less productive lakes, which are classified as oligotrophic or mesotrophic. Watersheds with rich organic soils or agricultural regions that are usually enriched with fertilizers yield much higher nutrient loads, resulting in more productive, eutrophic (even hyper-eutrophic) lakes (Horne and Goldman, 1994). Fractionation processes of stable isotopes are related to regional climate parameters, organic matter sources and/or photosynthesis. These processes are discussed below in section 1.4. Introduction Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles Biological features Biotic systems are the most complex components of lake systems, involving numerous species and the interaction between them and/or the external environment (Margalef, 1983). The different communities of organisms (planktonic, nektonic and benthic) are distributed as a function of their habitats, which are influenced by light, turbulence and proximity to the lake floor. Planktonic organisms are drifting organisms that inhabit the pelagic zone of the water column. Nektonic organisms also inhabit the pelagic zone but are able to swim against the water flow and so control their position. Benthic organisms inhabit the lowest level of a body of water, including the sediment surface and some sub- surface layers (Emiliani, 1991). Planktonic organisms are grouped into phytoplankton (e.g. cyanobacteria, dinoflagellates, diatoms) and zooplankton (e.g. rotifers, arthropods, copepods, insects). Nektonic organisms include fish, large crustaceans and tetrapods. Benthic organisms can be various algae, vascular plants, sponges, bryozoans, annelids, crustaceans, insects and molluscs. All these organisms are conditioned by abiotic factors such as light (Hill, 1996), dissolved oxygen (Dinsmore et al. 1999), pH (Battarbee, 1984), salinity (Bos et al. 1999), nutrient availability (Leland and Berkas, 1998), water turbulence (Johnson and Wiederholm, 1989), and substrate (Burkholder, 1996). Additionally, biotic effects, such as grazing, predation, competition and symbiotic interactions are also key processes in the distribution of freshwater organisms (Kitchell and Carpenter, 1993). 1.1.3 Lacustrine sediments Lake sediments constitute accurate sources of information about the lake history (Cohen, 2003, Smol, 2008, Cabrera et al. 2009). On the other hand, ancient lacustrine basins contain extensive evaporitic and oil shale deposits which have provided sites for the economic exploitation of resources such as uranium and hydrocarbons (Robbins, 1983). The material deposition at the bottom of a lake involves both horizontal and vertical transport processes. The former process is mainly due to clastic deposition from rivers and streams at the bottom of the lake. Depending on the geographical location of a particular lake, wind-blown, ice-rafted and volcanic material may also be locally important. The nature and size of the surrounding drainage basins exert a major influence on the input of terrigenous sediment. The sedimentation rates of this clastic material are usually irregular over time and are mainly associated with river discharge and catchment runoff, attaining occasionally very high values (up to several cm per year) (Dussart, 1961). Shorelines in lacustrine environments may be marked not by beaches but by stands of macrophytes (Livingstone and Melack, 1984). Siliciclastic sediments along lake margins are generally concentrated 8 9around river mouths given that the absence of tides means that wave attack may be limited to a narrow zone along the shores of hydrologically open lakes. In closed lakes, seasonal or longer-term variations in lake level may distribute the effects of littoral processes over a large area. Abandoned beach ridges, which are preserved as staircase-like features around the margin of modern closed basins, provide some of the most striking evidence for former high lake levels (Talbot and Allen, 1996). The offshore deposition is mainly regulated by dispersion and sedimentation of fine-grained suspended matter determined by lake circulation patterns. When lakes are stratified, the distribution of suspended sediment may be greatly influenced by density contrasts within the water column. Bottom-hugging density underflows are also effective in distributing fine-grained sediments over large areas of deep lake basins (Cohen, 1990). Ultimately, when an adequate clastic supply exists, turbidity currents can be a major source of sediment to the offshore areas of large lakes (Lambert and Giovanoli, 1988; Scholz et al. 1993). On the other hand, vertical deposition mainly consists of suspended materials of chemical and biological origin formed within the lake. In diluted lakes, these are typically carbonate, siliceous or iron mineral accumulations whereas saline waterbodies can produce a wide range of carbonate, evaporate and silicate minerals. According to Kelts and Hsü (1978) calcareous sediments are formed by one or more of the four following processes: a) inorganic precipitation generally associated with the photosynthetic activities of plants or, less commonly, induced by evaporation, changes in temperature, or mixing of water masses; b) production of calcareous shells, surface encrustations or skeletal elements of living organisms; c) clastic input of allochthonous carbonate particles derived from the drainage basin; and by d) postdepositional or early diagenetic precipitation. Large amounts of biogenic silica accumulate in some modern lakes (Talbot and Allen, 1996). Diatoms are the principal source of this silica, although sponges may locally make a contribution. Where silica supplies are not limiting, diatoms can outcompete other autotrophic species to become extremely abundant and major sediment contributors. Lakes or areas of the lake floor starved of clastic sediment may accumulate considerable thickness of diatomaceous ooze. Moreover, seasonal plankton blooms can produce diatom-rich laminae interbedded in sediments (Owen and Crossley, 1992; Ishihara et al. 2003; Chu et al. 2005; Simola, 2007; Jacques et al. 2009; Lindqvist and Lee, 2009). Although not widespread, sediments rich in iron minerals occur in some lakes. The Fe is believed to be derived from regolith and transported in colloidal or adsorbed form as part of the suspended load of incoming rivers. In lacustrine environments, this chemical element coprecipitates with silica (Lemoalle and Dupont, 1976). Saline minerals form after Ca-Mg carbonates precipitate from the water body (Talbot and Allen, 1996). They are usually produced by evaporative concentration, but temperature changes can also cause their precipitation. Saline minerals are found in three principal environments (Eugster and Hardie, Introduction Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 1978): a) in perennial brine bodies or saline lakes, b) as efflorescent crusts on and around ephemeral salt pans, and c) as cements within sediment deposited in and around these water bodies. Gypsum is commonly the first mineral to form after Ca-Mg carbonates, but it precipitates only if the alkaline earth elements have not previously been depleted by carbonate production. Thereafter, the saline minerals that precipitate depend upon the initial composition of the inflow, which is in turn a function of the geology of the catchment (Eugster and Hardie, 1978; Talbot and Allen, 1996). Lakes can also be important sites for the accumulation of organic matter. Lacustrine deposits often have contents of organic matter that are significantly above the average of sediments and sedimentary rocks. The organic matter in lakes has three principal sources: terrestrial vegetation, marginal macrophyte swamps, and phytoplankton. The remains of phytoplankton are most abundant in open-water, offshore sediments of large lakes, waterbodies receiving little clastic supply, in the middle of large crustal sag basins, and in areas where low slope gradients do not favour the widespread distribution of clastic material by density currents (Talbot, 1988). In contrast, where the lake margin is steep, density flows can transport terrestrial and marginal plant debris far offshore (Le Fournier et al. 1990). The preservation of large amounts of organic matter in lakes is dependent upon rates of organic productivity, anoxic bottom conditions, or a combination of both (Talbot, 1988). This sedimentation in lakes is therefore highly variable and is affected by six major factors (Fig. 1.3): watershed geology, watershed climate, ontogeny of the lake, inflow and outflow hydrology, internal water circulation, and organic productivity (Cohen, 2003). Hence, many lake sequences should be studied, at least, centimetre by centimetre in order to document the full range of sedimentary environments (Tallbot and Allen, 1996). Rhythmites Rhythmites are defined as «sequences of finely laminated, regular alternations of two or more contrasting sediment types» (Talbot and Allen, 1996). These repetitive clusters are called couplets when formed by two facies, but the pattern of repetition may be more complex, involving three (triplets) or even more repetitive sediment types (Cohen, 2003). They are particularly useful for reconstructing high resolution records of terrestrial climate because their individual laminae can provide valuable information about the variable processes responsible for generating each component defining the rhythmite (Simola, 2007; Shunk et al. 2009). The component sediments of a rhythmite may be a combination of clastic, chemical and/or organic deposits rich in organic matter (Fig. 1.4). When diatoms are preserved in laminated sediments, they often provide environmental records of high quality. The rhythmically laminated sediments, mainly made up of diatom frustules, may reflect the regular yearly succession and the seasonal blooms of planktonic diatoms (Simola, 1992). In other cases, the absence of such seasonality may indicate a non-annual rhythmite. 10 11 The accumulation and preservation of thick rhythmite sequences requires a special combination of environmental conditions. There must be a periodic variation in the nature of the sediment reaching a lake floor and are favourable environment for rhythmite preservation (Kelts and Hsü, 1978; Zolitschka et al. 2000; Francus et al. 2008; Besonen et al. 2008). Kelts and Hsü (1978) proposed the following scenarios where a) bioturbation is minimised to preserve laminae, b) sedimentation rates are high even in oxic conditions, c) bottom current activity is minimised to preclude resuspension, d) the floor below the wave base is relatively extensive and flat to prevent incoherent sediment from creeping or sliding down- Figure 1.3. Major factors affecting the broad-scale differences observed in lake sedimentation patterns, facies development, and facies geometry. Note that even these «major» factors are not independent of one another but instead affect each other (shown in bold) through a complex web of interactions (modified from Cohen, 2003) Introduction Watershed geology Fabric/mineralogy of source sediments Weathering intensity & transport rates Sediment, solute & groundwater supply Brine evolution/authigenic mineralisation Mixing frequency/redox state of sediments Lake origin & evolution Inflow/outflow hydrology & lake level fluctuations Solute/temperature effects on diagenesis Fabric/mineralogy of sediments Particle/solute supply Watershed geology Basin morphometry & sediment delivery Internal circulation Sediment accomodation and facies geometry Hydrotermal and groundwater effects on diagenesis Lake level fluctuation and facies migration Authigenic mineralisation Sediment focussing/redistribution Internal circulation (water depth) Redox-related mineralisation Subsurface brine evolution Suspended sediment settling patterns Organic productivity Redox state & bioturbation (extent of organic-rich deposits) Redox-related mineralisation and diagenesis Sediment focussing Biogenic particle production Community and evolutionary change/effects on sediments Organic particle flux Redox-related mineralisation and diagenesis Extent of organic-rich deposits Sediment, solute & groundwater supply Erodibility of source materials Mode of lake formation Watershed climate Mode of lake formation/evolution Inflow/outflow hydrology Internal circulation Organic productivity Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles slope, e) gas bubble generation is minimised so as not to disrupt lamination, f) there is a a seasonally variable flux of mineral and organic matter from epilimnion that settles through the water column, and where g) particle settling occurs fast enough for environmental variations to be transmitted to the lake floor. 2 .5 m m c a 1 m m 0 .4 -3 0 m m 2 -5 m m1 -1 0 m m SP SP W Dry/windy Dry Wet/calm Wet W A A SU SU SP Spring Summer Autumn WinterW A SU SP-SU W Plant fragments Dinoflagellates Ostracodes Magadiite spherulites Calcite crystals Organic-rich muds and silts Laminated muds and silts Sand-size clastic grains Light Dark Diatoms A D EB C Figure 1.4. Some rhythmite types showing lithologies and relationships between compositional variations and seasonal or longer- term climate change (modified from Talbot and Allen, 1996). A) Varve from a subarctic Swedish lake (after Renberg, 1981); B) Varve from a glacial lake (after Smith and Ashley, 1985); C) non-glacial varve from a temperate hardwater lake (Lake Zürich, after Kelts and Hsü, 1978); D) non-glacial varve from Lake Malawi (after Pilskaln and Johnson, 1991); E) rhythmite from the Upper Pleistocene section in Lake Magadi, Kenya. Each couplet represents ca 2−3 years accumulation (after Damnati et al.,1992) 12 High-resolution lake records often display cyclical facies changes at different scales, indicative of fluctuations in productivity, lake level, and climate. Commonly, the seasonal sediment pulse is responsible for generating a rhythmic sedimentation pattern because the seasonal climate pulse governs the production, movement and deposition of sediment in the lacustrine system (Anderson and Dean, 1988; Brauer et al. 1999; Zolitschka et al. 2000). These rhythmites are termed varves. However, it is conceivable that rhythmites could represent a sediment cycle produced by other factors than the annual seasonal cycle. For example, they can record periodicities marking solar cycles (Kemp, 1996; Shunk et al. 2009) as well as climate cycles such as the ENSO phenomenon (Rittenour et al. 2000; Muñoz et al. 2002). A great deal of research has focused on tracking the climate signal in sedimentary rhythmites (e.g. Grosjean et al. 1995; Zolitschka and Negendank, 1996; Brauer et al. 1999; Prasad et al. 2006; Kirilova et al. 2009). Hydroclimate cyclicities at different scales are evidenced by a variety of depositional trends that provide a linkage between the rhythmite pattern and climate. Laminae thickness patterns may change as a result of decadal-to centennial-scale trends in rainfall or temperature (Kelts and Hsü, 1978; Giralt et al. 1999; Romero- Viana et al. 2008). Furthermore, mineralogical and organic contrasts are also common as a result of changing moisture balance. Overall, the main problem is to distinguish the climate signals from those of catchment processes and, especially those, from human activities (Lotter and Birks, 1997). 13 1.2 Diatoms Diatoms (Phyllum Heterokontophyta, Class Bacillariophyceae) are a widely distributed group of microscopic, unicellular and photosynthetic organisms that are characterised by cell walls composed of opalline or biogenic silica (SiO 2 · nH 2 O), which conforms to a particular type of skeleton called frustule (Round et al. 1990). Much of the uniqueness of diatoms is related to their silica composition (Battarbee et al. 2001). Diatoms live in lakes, oceans, rivers and streams, and many typically live as free-floating species, making up part of the ecological assemblage known as phytoplankton (Willén, 1991; Reynolds, 2006). Other species grow on various substrates forming part of the ecological assemblage known as periphyton (Theriot, 2001). Before the emergence of diatoms about 250 Myrs ago in accordance with the estimates of the molecular-clock (Sorhannus, 2007), the phytoplankton in the seas consisted primarily of cyanobacteria and green algae that were slightly larger than bacteria (Falkowski et al. 2004). The appearance of diatoms and the emergence of other groups such as dinoflagellates and coccolithophorids led to a major shift in global organic carbon cycling. This ushered in an era of declining atmospheric CO 2 concentrations and increasing atmospheric O 2 (Armbrust, 2009). At present, the main importance of diatoms is that they contribute to 20-25% of global primary productivity (Saade and Bowler, 2009), which is equivalent to the photosynthetic activity of all the rainforests combined (Field et al. 1998; Saade and Boular, 2009). The kind of diatoms found in lakes depends on the range and extent of habitats available for growth, and also on the combination of physical, chemical and biological conditions that prevail in the water column (Battarbee et al. 2001). Diatoms require dissolved inorganic nutrients, including CO 2 , N, P, Si and a variety of metals and vitamins, to photosynthesise and reproduce (Werner, 1977). Diatom shape, form and physiology vary in ways that suggest that they have adapted to the different environments in which they occur (Theriot, 2001). They can survive in a wide gradient of pH values, concentrations of solutes, nutrients, and organic and inorganic contaminants, and across a range of water temperatures. Individual species are often restricted to specific ecological conditions which make them extremely valuable for palaeoenvironmental reconstructions (Stoermer and Smol, 1999; Battarbee et al. 2001; Crosta and Koç, 2007; Jones, 2007). Most diatom silica is recycled through dissolution within the water column, but a fraction, typically 5–20%, is incorporated into lake sediments (Bootsma et al. 2003) and preserved as a record of environmental change. Good examples are the records of eutrophication (Hall and Smol, 1999), acidification (Birks et al. 1990) or climate change (Mackay, 2007). Diatoms do not, however, respond directly to the temperature of the atmosphere and the amount of rain (Fritz et al. 1999; Anderson, 2000). The water budget of endorheic lakes is mainly driven by changes in precipitation to evaporation balance (P/E) that bring about changes in ionic concentrations, causing a rise or a decrease in salinity and/or alkalinity, and in nutrient availability. In this way, diatoms can be used as indirect indicators of P/E in closed lake systems (e.g. Fritz et al. 1991, 1999; Gasse et al. 1997; Bao et al. 1999; Barker et al. 2002). Introduction Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles More recently, diatom frustules have also proved useful for chemical analysis (Swann and Leng, 2009). Stable isotopes such as δ13C, δ15N, δ18O and δ30Si can now be extracted from diatom silica or from the organic matter enclosed within, and used for palaenvironmental reconstructions (e.g. De La Rocha, 2002; Crosta and Shemesh, 2002; Robinson et al. 2004; Hurrell et al. 2009). 1.2.1 The diatom frustule The cell wall of a diatom is made up of a silica-based scaffold and specific organic macromolecules (proteins and polysaccharides) (Kröger and Poulsen, 2008). Proteins located in the cell wall are believed to control the morphogenesis of the species-specific silica structure (Hecky et al. 1973; Kroger et al. 1994). Silica formation takes place in intracellular compartiments termed silica deposition vesicles (SDVs). The SDVs grow during cell division until valve formation is complete and the SDVs fuse with the cell membrane (Sumper and Kröger, 2004). Suggested functions for the siliceous cell wall include acting as an ultraviolet filter, as armour to protect against grazing by zooplankton, as ballast to control water column position, and as a buffer in the conversion of HCO 3 - to CO 2 . It has also been proposed that the construction of a cell wall with silica is energetically cheaper than with organic carbon (Milligan and Morel, 2002; Hurrell, 2009). The opalline diatom cell wall consists of two halves that are very similar in structure with one half (epitheca) that is slightly larger and overlaps the other half (hypotheca). The epitheca and hypotheca completely enclose the protoplast (Round et al. 1990). Diatoms can be divided into three major groups according to the shape and symmetry of their cell walls: radial centrics, polar centrics, and pennates (Kröger and Poulsen, 2008) (Fig. 1.5). Radial centrics have a Petri dish-like shape with a circular centre of symmetry (annulus) in the middle of the valve, and rows of pores (striae) radiating from the annulus. Polar centrics have bi- or multipolar valves with an elongated or distorted annulus. Pennate diatoms are bilaterally symmetrical and instead of an annulus they possess a rib (sternum) running along the longitudinal axis of each valve. In raphid pennates, the sternum contains a slit (raphe) that is obstructed by the central nodule in the middle of the valve (Edgar and Pickett-Heaps, 1984; Round et al. 1990). The rigidity and architecture of the silica cell wall imposes restrictions on the mechanism of cell division and growth (Round et al. 1990). Diatom valves can only be formed during cell division, but new valves are usually smaller than the parental cell. Accordingly, the average cell size in a diatom population gradually decreases with continued vegetative growth (the MacDonald-Pfitzer rule) (Macdonald, 1869; Pfitzer, 1869; Battarbee et al. 2001). Ultimately, this reduction in size would result in cells that are too small to be viable with the result that the diatom population would die. The only way to escape this fate is by sexual reproduction (Chepurnov et al. 2004). During this process meiotic cell division takes place and the resulting gametes slough off their cell walls. Immediately after gamete fusion, a specialized zygote (auxospore) is formed, leading to a considerable increase in volume within a relatively short time (hours to a few days) (Fig. 1.6). 14 15 The silica frustules enclose the content of the cell (protoplast) which includes chloroplasts, cytoplasm, vacuoles, mitochondriae and SDVs (Sicko-Goad et al. 1984). The cell wall itself is enclosed in a brown/yellow organic biofilm (mucilage layer) that is not closely associated with the siliceous structure (Crawford et al. 2001). In a sedimentary setting, some or all of these components may be subject to diagenesis, thus altering the elemental composition and isotopic ratios required for environmental reconstructions (Hurrell, 2009). Alternatively, amino acids and long chain polyamines entombed within the structure of the silica cell wall contain carbon and nitrogen which may be independently preserved (Hecky et al. 1973; Volcani, 1981). 1.2.2 Ecology of diatoms Despite the fact that diatoms display a wide species-specific range in their ecological preferences, they prefer aquatic environments with a turbulent regime or environments at the onset of stratification in condi- tions of low light, low temperature and with suficient nutrients (Willén, 1991). Although diatoms have been considered cosmopolitan (e. g. Kociolek and Spaulding, 2000), their distribution can also be due to historical constraints within geographical regions. Recent studies suggest that this distribution is regulated by the same processes that operate in macro-organisms, involving the existence of endemisms attaining unexpected high levels (Vanormelingen et al. 2008). Therefore, the study of the ecological preferences and historical distributions would help us to better understand diatom occurrences (Stoermer and Julius, 2003). Diatoms are good competitors during periods of low light and comparatively low temperatures (Margalef, 1983). Most diatoms have temperature optima in the laboratory well above 10°C. However, they Introduction Figure 1.5. Scanning electron microscopy (SEM) images of the cell walls of four different diatom genera: A) Cyclostephanos, B) Triceratium, C) Staurosira, D) Navicula. A C D B Diatoms require dissolved inorganic nutrients including CO 2 , N (NO 2 -, NO 3 - and/or NH 4 +), P (as PO 4 3- although uptake of organic phosphates may occur), Si (as dissolved Si(OH) 4 ) and a variety of metals (such as Fe) and vitamins (for instance vitamin B12) to photosynthesise and reproduce (Werner, 1977; Willén, 1991). As stated above, C, N and P follow a more or less closely stochiometric relationship of 106:16:1 (by atoms) in healthy algae (Harris, 1986; Reynolds, 2006). Whether or not CO 2 is a limiting nutrient for dia- toms in the ocean (Riebesell et al. 1993) remains a moot point (Vaulot, 2006). Si is essential for diatoms since it is necessary for the construction of the siliceous cell wall. Silica is less available in oceans than in lakes, and marine diatoms have four times less silica per cell than freshwater diatoms on average (Conley et al. 1989). In Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles a b c d e f g h i Figure 1.6. Life cycle of a diatom based on Stephanodiscus except that the position of the auxospore (c) and the formation of motile gametes (a) have not been seen in this particular genus; d) auxospore wall breaking open to reveal initial cell; e) first division of initial cell to form two new normal hypovalves back to back; f) one of the cells from stage (e) with a normal valve and an initial cell valve; g) a cell formed following several divisions of f) (the other valve is not illustrated but will be a replica of f); h) and i) vegetative size reduction; i) small cell which will give rise to either male or female gametes (Modified from Round et al., 1990). 16 are more abundant in lakes in spring when nutrients are in high concentration and temperatures are as low as 4°C (Werner, 1977; Theriot, 2001). Diatoms that develop during spring blooms have higher growth rates at lower light levels and lower temperatures than diatoms growing later in the year when temperature and light conditions are higher (Guillard and Kilham, 1977). Planktonic diatoms are as dense, or denser, than water, and lack all propulsion with the result that they will quickly sink in still water (Round et al. 1990). They therefore depend on the existence of a turbulent regime to avoid sinking (Margalef, 1978). 17 Introduction a series of culture experiments, Tilman et al. (1982) reported the competition between different freshwater diatom specie as a function of the Si:P ratio. These authors found that large pennate diatoms were better competitors at high Si:P ratios and that small centrics required more P. Fe, which is a micronutrient, has also been found limiting in certain oceanic environments (Martin, 1992). The complex interplay between light, temperature, nutrient concentration, turbulence, and biotic interactions throughout the annual cycle in aquatic ecosystems gives rise to a phytoplankton succession where diatoms are present (Margalef, 1983; Harris, 1986; Reynolds, 2006). A fundamental model of how phytoplankton succession proceeds is summarised in Theriot (2001). Phytoplankton generally occur in low numbers in the winter. The temperature of the water column is more or less uniform with the result that mixing is facilitated. On the other hand, there is little light for growth. Nutrient-rich deep waters circulate and are well mixed until spring, when thermal stratication commences and diatoms and other phytoplankton begin to grow. Phytoplankton reproduce at the same rate as they consume the nutrient supply. This is particularly true of silicate, which, being relatively insoluble, is transported to the sediments as diatoms die or are eaten. The freshwater and marine systems must await autumn turnover (during which nutrient-rich bottom waters are again mixed to the surface) and/or heavy runoff to return nutri- ents to the well-lit surface waters. In this dynamic system, nutrients are utilised and regenerated at dif- ferent rates and from different sources. It may well be that absolute levels of nutrients and light do not determine diatom species. Rather, diatoms are determined by the rate at which nutrients are supplied (through regeneration by dissolution, runoff, rainfall, bacterial degradation, etc.), the ratios of nutrient concentrations and supply rates. In other words, a system with high N and high P may have the same dominant diatom as a lake with a smaller supply of these nutrients provided that the nutrient ratios are similar. This could account for the otherwise enigmatic distribution of some diatom species (Theriot, 2001). 1.3 Isotopes Isotopes are atoms whose nuclei contain the same number of protons but a different number of neutrons (Urey, 1947). They can be divided into two main kinds: stable and radioactive species. The total number of known stable isotopes is about 300 although over 1,200 radioactive species have been discovered to date (Hoefs, 2004). Isotope exchange reactions and kinetic processes are the main phenomena that produce isotope fractionation, which is the partitioning of isotopes into two substances or two phases of the same substance (Bigeleisen and Wolfsberg, 1958; Melander and Saunders, 1980). Differences in chemical and physical properties arising from variations in atomic mass of an element are called «isotope effects» (Bigeleisen and Mayer, 1947; Melander, 1960; Richet et al. 1977; Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles O’Neil, 1986; Fritz and Fontes, 1986; Clark and Fritz, 1997; Chacko et al. 2001; Schauble, 2004). These isotope effects give rise to certain differences in the physicochemical isotope properties because of mass differences. The replacement of any atom in a molecule by one of its isotopes leads to a very small change in chemical behaviour. In nature these slight differences result in isotopic fractionation according to biological, chemical and physical processes. This fractionation, which is indicated by the fractionation factor (α), is recorded in the sediments and may facilitate palaeoenvironmental reconstructions (Criss, 1999; Hoefs, 2004). This α is defined as the ratio of the numbers of any two isotopes in one chemical compound divided by the corresponding ratio of another chemical compound (Ito, 2001). However, in isotope geochemistry it is common to present isotope composition in terms of delta (δ) values and this is expressed as parts per thousand (per mil or ‰). Thus, isotopes are expressed as ratios relative to an internationally recognised standard (equation 1): δ = [(R compound /R standard ) −1]·103 (‰) (1) where R compound is the absolute isotope ratio of the sample and R standard is the absolute isotope ratio of the laboratory standard (e.g. Ito, 2001; Schwalb, 2003; Leng and Barker, 2006). The stable isotopic composition of any compound reflects the relative proportions of the two stable isotopes. This composition shows the degree of depletion in the heavier isotope with respect to a specific standard (Fritz and Fontes, 1986; Clark and Fritz, 1997). 1.3.1 Stable isotopes in palaeoenvironmental research Stable isotopes, such as δD, δ13C, δ15N, and δ18O can be employed to better understand present and past environmental and climate evolutions of natural systems. There are several natural archives where the stable isotope analyses may be applied, such as water, marine and lacustrine sediments, bones and teeth, speleothems and tree rings (e.g. Gat, 1996; Eronen et al. 2002; Drucker et al. 2003; Leng and Marshall, 2004; Maslin and Swann, 2005; McDermott et al. 2005). δD, δ13C and δ18O of the studied compound, organism or sediment reflect specific boundaries and/or environmental conditions of the water and carbon cycles at the time of their formation (Fig. 1.7). Therefore, the present isotopic behaviour of the chemical elements within these cycles must be well understood and characterised in order to interpret past isotopic oscillations recorded in the sedimentary sequences (Darling et al. 2005). The measurement of δ18O and δ13C isotopes in marine sediments has played a major part in establishing stable isotopes as one of the most important sets of proxies within palaeoceanography. δ18O 18 19 from marine sediments allow the reconstruction of past global ice volume, ocean temperatures, relative sea level, ocean circulation and structure, surface water salinity, iceberg melting, river discharge, and monsoonal intensity (e.g. Urey, 1947; Emilliani, 1955; Shackleton and Opdyke, 1973; Tiedemann et al. 1994; Shemesh et al. 1995; Shackleton et al. 1995; Niebler et al. 1999; Maslin and Burns, 2000; Zachos et al. 2001; Skinner et al. 2003; Maslin and Swann, 2005). δ13C carbonates , δ13C bulk and δ13C diatom from marine sediments can provide information on the past carbon cycle over a range of time-scales, offering a useful insight into past marine productivity, ocean circulation, past surface water, atmospheric CO 2 , and storage and exchange of carbon at both local and global scales (Berger et al. 1978; Sarnthein et al. 1994; Andersen et al. 1999; Ruhlemann et al. 1999; Rosenthal et al. 2000; Kennett et al. 2003). Stable isotopes in palaeolimnology Stable isotope composition of lacustrine sedimentary materials can yield a wide range of useful palaeoclimate information (Leng et al. 2005b). Stable isotope studies have become an essential part of palaeolimnology since McCrea (1950) and Urey et al. (1951) highlighted the potential of oxygen isotope composition for palaeotemperature reconstruction. δ18O are the main isotopes used in palaeolimnology although other palaeoenvironmental information can be obtained from δD, δ13C and δ15N in lacustrine biogenic materials (Leng et al. 2005b). It is possible to measure several stable isotope ratios (e.g. δD, δ13C, δ15N, δ18O) from either bulk lake sediments or any organic and/or inorganic compound depending on the kind of material incorporated into the lake deposits (Fig. 1.8). * ** ** * * * * * vapour transport evaporation evaporation transpiration precipitaton precipitaton surface runoff percolation groundwater ocean-atmosphere exchange land-use changes photosynthesis & respiration fossil fuel combustion & cement manufacture ocean circulation sinking particles river runoff photosynthesis & respiration atmosphere surface ocean phytoplankton deep ocean terrestrial biosphere lithosphere A B Figure 1.7. A) Block-diagram of the global water cycle. Evaporation of seawater leads to cloud formation with progressive rainout as temperature decreases. The moisture will eventually return to the sea mainly via surface runoff or groundwater flow though it may be delayed by recycling via evaporation from lakes or transpiration from vegetation. B) Block-diagram depiction of the Earth’s carbon cycle, highlighting various sources and sinks for carbon. Note the arrows indicating the main relationships between the different sources and sinks Introduction Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles The δ13C bulk in lake sediments are important for assessing organic matter sources, reconstructing past productivity rates and for identifying changes in the availability of nutrients in surface waters. Increases in the accumulation rates of organic matter and its δ13C have been widely used as an indicator of enhanced aquatic productivity in lakes (Hollander and McKenzie, 1991; Brenner et al. 1999; Meyers, 2003). The δ13C bulk are proxies for palaeoenvironmental changes in lake watersheds as well as in the lakes themselves (Meyers and Teranes, 2001). Although the δ15N bulk are less used, they can also help to distinguish the sources of this material and reconstruct past productivity rates (Gu et al. 1996, Talbot and Laerdal, 2000). This proxy is particularly useful in identifying changes in the past availability of nitrogen used by aquatic primary producers. However, the dynamics of nitrogen biogeochemical cycling is very complicated. Changes in nitrogen cycling that accompany anthropogenic eutrophication of lakes can also exert a strong influence on the δ15N bulk (Brenner et al. 1999; Teranes and Bernasconi, 2000). This does not facilitate δ15N interpretations of sedimentary records (Fogel and Cifuentes, 1993; Brenner et al. 1999; Talbot, 2001; Meyers, 2003). Despite their great potential for providing palaeolimnological information, oxygen and hydrogen isotopic compositions of organic matter have rarely been determined in sedimentary records because of practical difficulties in their analysis (Meyers and Teranes, 2001). diatom silica chemical removal of organics and carbonates sieve at above and below mean diatom size (50-250 m) various separation techniques to provide pure diatom (settling, heavy liquids, splitt flow techniques)  18 O chemical removal of organics sieve at 80 m shell material (>80 m) identification (microscopy and hand selection) gastropods molluscs ostracods separation into individual species  13 C  18 O marl (<80 m) fraction  mineral identification (microscopy, SEM and XRD) mixed mineralogy high % calcite (partial reaction techniques)   13 18 C O calcite   13 18 C O chemical removal of carbonates   13 15 C N chemical separation of organic compounds (cellulose, lipids, kerogens)     13 15 18 C N O D sieve &/or hand select for specific fractions (chironomid, pollen)     13 18 C O 15 N D carbonates organics Figure 1.8. Flow diagram showing the most common materials analysed by isotope analysis in lake sediments, and the laboratory pretreatments required (Leng et al. 2005b). 20 21 Isotope analyses in carbonates constitute an ideal tool for identifying the mechanisms that cause their accumulation in lake sediments, enabling the reconstruction of the history of a lake body (e.g. Schelske and Hodell, 1991; Thompson et al. 1997; Hodell et al. 1998; Ricketts and Anderson, 1998; Filippi et al. 1999). The ability to distinguish the carbonate sources archived in lacustrine carbonates through stable isotope analyses demand a thorough understanding of the local lake dynamics and the biology of the lake system (Gierlowski-Kordesch, 2010). Detailed studies of the climate, hydrology and the provenance of a lake are necessary for a reliable reconstruction and interpretation of the isotope signals archived in its sedimentary carbonates (Ito, 2001; Leng and Marshall, 2004). δ13C of lake waters is influenced by the geochemical composition of input waters, atmospheric CO 2 exchange, gas mixing through bacterial methanogenesis in the degradation of organic matter, and pelagial photosynthesis (McKenzie, 1985; Talbot and Kelts, 1990). On the other hand, δ18O is influenced by the isotopic composition of waters supplied to the lake, including rainfall, surface inflow, and groundwater inflow (Leng and Marshall, 2004). Temperature, and thus evaporation, also controls the output of the lighter oxygen (16O), affecting the isotope ratio. Changes in temperature, rainfall sources (especially through seasonal changes), P/E, riverine influx, and even groundwater input are postulated to have been preserved as δ18O carbonates precipitating in lake waters (Teranes et al. 1999; Lamb et al. 2000; Schwalb and Dean, 2002; Leng and Marshall, 2004; Yansa et al. 2007). Biogenic silica in lacustrine sediments is deposited by a variety of aquatic organisms, including diatoms and sponges, and its study is especially useful in lakes with no endogenic carbonates (Leng and Marshall, 2004; Leng and Barker, 2006). Measurements of δ18O diatom provide a potentially important source of palaeolimnological information equivalent to that from δ18O carbonate . δ18O diatom has been used as a proxy for temperature change in some studies (e.g. Rietti-Shati et al. 1998; Rosqvist et al. 1999; Leng et al. 2001; Hu and Shemesh, 2003) although the temperature dependence of oxygen isotope fractionation between diatom silica and water continues to be controversial. The use of δ18O diatom in terms of variations in the isotope composition of the lake water due to changes in P/E or in the source of precipitation has become more rational (e.g. Barker et al. 2001, 2007; Shemesh et al. 2001; Jonsson et al. 2010). A new approach involving a combined methodology for analysing δ18O diatom and δ30Si diatom in the same sample has recently been implemented (Leng and Sloane, 2008; Swann et al. 2010). In conjunction with separate analyses on δ13C diatom and δ15N diatom , this will permit a more detailed, isotope based reconstruction of carbon and nitrogen cycles (Hurrell et al. submitted). To date, few studies have focused on the combined analysis of δ18O diatom , δ30Si diatom , δ13C diatom and/or δ15N diatom at the same site (Hurrell, 2009; Swann et al. 2010; Hernández et al. submitted). Introduction Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 1.3.2 Stable oxygen isotopes δ18O can be obtained from a large range of organic and inorganic materials from different sources. It is essential to obtain accurate information about how rainfall isotope compositions are determined by climate in order to better understand the environmental signal contained within the isotope composition from different materials (Leng et al. 2005b). The water cycle is especially important, since it is the precursor back to which most δ18O studies are attempting to relate (Clark and Fritz, 1997). While the concept of the water cycle may be simple in essence −water evaporates from the sea, falls as rain over land, and eventually returns to the sea mainly via river and groundwater discharge (Fig. 1.7)− there are a number of complicating factors at least where δ18O are concerned, such as seasonal freeze-thaw, plant transpiration, soil evaporation and/or evaporation from surface waters (Dincer et al. 1974; Allison et al. 1984; Berner and Berner, 1987). In particular, lake sediments are an important archive for climate change, and it is therefore vital to improve our understanding of the often complex isotopic relationship between individual lakes and regional precipitation (Darling et al. 2005). δ18O carb/diatom is controlled by water temperature and by the isotope composition of the lake water from which the mineral is formed (Stuiver, 1968)(Fig. 1.9). If the mineral is precipitated in isotope equilibrium with the lake water, then mineral–water fractionation equations may be used to estimate variations in past temperatures providing there is no change in the δ18O lakewater (Leng and Marshall, 2004). A thorough understanding of the factors that may have influenced the isotope composition of the lake water (δ18O lakewater on Fig. 1.9) is necessary to correctly interpret the δ18O signal from the analysed mineral  18 Odiatom/carb  18 Owater  18 Oprecipitation Temperature ‘Disequilibrium’ (+vital) effects What Temp? Season? Time of day? Water depth? Weighted mean? P/E variation Temperature Seasonality Air mass source Changes in lake water input and output Figure 1.9. Controls on the oxygen isotope composition of lacustrine diatom silica and carbonates (δ18O diatom/carb ). If biogenic material is precipitated in isotopic equilibrium, δ18O diatom/carb depends entirely on temperature and on the isotopic composition of the lake water (δ18O lakewater ). Disequilibrium effects, commonly known as ‘vital effects’ in biogenic precipitates caused by local changes in microenvironment or by the rate of precipitation can induce systematic or non-systematic offsets in the δ18O diatom/carb signal (modified from Leng and Marshall, 2004 and Leng and Barker, 2006) 22 23 or organic compound. The δ18O lakewater in hydrologically open lakes usually reflects the isotope composition of precipitation (δ18O precipitation ), both rain and snowfall, received by the lake, whereas oscillations of δ18O lakewater in closed lakes respond predominantly to the P/E (Ito, 2001; Leng and Barker, 2006). Moreover, disequilibrium effects (vital effects in biogenic materials) (Fig. 1.9) should also be taken into account when the material analysed is composed by more than one taxon. Stable oxygen isotopes from diatoms All silicates, including diatom silica, are composed of silica tetrahedrons. Following the uptake and fractionation of oxygen, covalent Si–O–Si bonds are formed in the diatom frustule through the condensation of two Si–OH groups to form (SiO 2 ) n (Labeyre and Juillet, 1982) (Fig. 1.10). Within the centre of the diatom frustule the –Si–O–Si bonds form an isotopically homogeneous dense layer of silica, the δ18O diatom of which is assumed to reflect δ18O from the water in which the silica precipitated at a given temperature (Juillet, 1980a,b). Research on δ18O diatom was first developed in marine sedimentary records (e.g. Labeyrie, 1974; Labeyrie and Juillet, 1982, Labeyrie et al. 1984; Shemesh et al. 1993, 1995). Nonetheless, in the last decade, the δ18O diatom in lacustrine sediments has increased in number since carbonates may be rare (or absent) in non-alkaline, dilute, and/or open lakes (Leng and Marshall, 2004). Such lakes are common at high-altitudes and are ideal for investigating climate change using δ18O because the isotope composition of the lake water is often similar to that of meteoric water (on either an annual or seasonal basis) (Leng and Barker, 2006). Introduction Figure 1.10. Structure of biogenic silica showing a 2nm particle. The core of the particle is the only part predominantly made up of Q4 and subsequent layers contain a mixture of Qn species. Q represents a silicon atom surrounded by four oxygen atoms and suffix n gives the number of surrounding oxygen atoms (out of 4) that are bonded to another silicon atom. Typically, the amount of Q1 and Q2 is small and therefore the level of hydration is studied as a ratio of Q4/Q3. A higher ratio implies lower hydration and a higher atomic organisation of biogenic silica (Leng et al. 2009).Oxygen Silicon Hydrogen Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles The δ18O diatom record results from a) changes in P/E (e.g. Barker et al. 2007), b) changes in lake temperature assuming that the δ18O diatom does not vary (e.g. Hu and Shemesh, 2003), c) oscillations in δ18O or temperature at the source of the precipitation (e.g. Swann et al. 2010) or d) changes in isotopic composition of δ18O lakewater related to fluctuations in the contribution of precipitation from different sources (e.g. Rosqvist et al. 2004). Naturally, more than one effect is possible, and different effects may predominate at different times with the result that identification is not easy (Jones et al. 2004; Morley et al. 2005; Leng and Barker, 2006). 1.3.3 Stable carbon isotopes The carbon cycle is much more complex than the water cycle even if our attention is restricted to dissolved organic (DOC) and inorganic (DIC) carbon in water (Fig. 1.7). This is because there are several sources of carbon in contrast to the single source of water (precipitation). δ13C are crucial for palaeoenvironmental reconstructions because they are affected by many factors that may be more or less related to climate. They have considerable potential for the interpretation of past environmental conditions where the overall geochemical conceptual model can be suitably constrained (Darling et al. 2005). In particular, the lake carbon pool is a complex mixture of DIC and DOC derived from internal biological and chemical processes and inputs from the catchment (Fig. 1.11) (Meyers and Teranes, 2001; Duarte and Praire, 2005; Cole et al. 2007). Inorganic carbon isotope ratios from DIC in lake waters (δ13C DIC ) are useful as tracers of environmentally determined processes which are often attributed to climate change (Leng and Marshall, 2004). In general, there are three predominant processes that control δ13C DIC : a) annual turnovers that mix low δ13C DIC from oxidised or respired organics from the lake bottom throughout the water column (Ito, 2001; Houser et al. 2003); b) extraction of low δ13C DIC from CO 2 by photosynthesis, atmosphere-water exchange or evasion of CO 2 from water (Cole et al. 1994); and c) input of DIC or DOC into the lake from surface runoff and groundwater inflow (Ito, 2001). Additionally, the release of methane from surface sediments by methanotrophic bacteria can also create an isotopically depleted DIC pool (Boschker and Middelburg, 2002) (Fig. 1.11). Lacustrine DOC is derived from allochthonous organic material or through autochthonous productivity. Lake derived organic matter, such as phytoplankton and macrophytes, is added to the DOC pool (Meyers and Teranes, 2001). The δ13C of the DOC (δ13C DOC ) in the lake water body reflects the isotopic composition of the source materials. Although DOC inputs from the catchment are large, they are often in a form not readily available to aquatic organisms (Pace et al. 2004; Brönmark and Hansson, 2005). However, oxidation of terrestrial DOC and respiration of bacteria that feed on DOC can contribute a considerable amount of DIC to lakes (Tranvik, 1988; Jonsson et al. 2001; Cole et al. 2002) (Fig. 1.11). 24 25 Stable carbon isotopes from diatoms Although diatoms require carbon, usually in the form of CO 2 , for photosynthesis (Werner, 1977), they are also able to utilise HCO 3 - through biophysical concentrating mechanisms (Johnston et al.2001; Milligan and Morel, 2002). A complex set of biological and physical factors control δ13C diatom . The literature suggests primary productivity and CO 2(aq) concentration as the main factors that may determine the δ13C diatom (Shemesh et al. 1993, 2002; Singer and Shemesh, 1995, Crosta and Shemesh, 2002; Schneider-Mor et al. 2005). However, other factors such as diatom growth rate, carbon source, metabolic pathway, diatom shape and/or taxonomic composition can also influence the isotopic signal of organic matter (Fry, 1996; Laws et al. 1997; Gervais and Riebesell, 2001). Furthermore, this organic matter, which is enclosed in the diatom cell wall, is not altered by diagenesis with the result that it is useful in palaeoenvironmental reconstructions (Des Combes et al. 2008). The interpretation δ13C diatom is not simple. The most common interpretation which relates higher δ13C diatom values to high productivity events in marine sediments is in apparent contradiction to the recent results found at other lacustrine sites (Hurrell et al. submitted). Alternatively, δ13C diatom in lacustrine sediments can also be used to interpret lake-catchment interactions (Hurrell et al. submitted). These findings have important implications for the understanding of the global carbon budget. They show that  13 C = -20 to -33‰org  13 C = 0 to -10‰calcite Anoxic respiration Diagenesis CH4 Hypolimnion: Anaerobic Epilimnion: AerobicOxic respiration Organic Matter Terrestrial Organic Matter C plants ( 3  13 C= -22 to -33‰) C plants ( 4  13 C= -8 to -22‰) ( ~-20‰) ( ~-8.8 to - 11.5‰) ( ~-1.2‰) ( ~-3.4‰)  13 C= -7 to -8‰ CO2(g) CO2(aq) CO3 4- Assimilation (mainly C algae)3 HCO3 - CaCO3 Ca 2- Watershed DIC DissolutionOxidation Figure 1.11. Idealised carbon isotope cycle in a small stratified lake. The isotopic composition of organic matter buried in sediments is determined by the proportions of different terrestrial and lacustrine organic matter, the carbon isotopic composition of dissolved inorganic carbon (DIC), and the rates of primary production and respiration within the water column. Isotope enrichment factors (ε), listed here as the difference between the product and the substrate, vary with the form of DIC, which is assimilated by lake algae (e.g. CO 2(aq) or HCO 3 -). Inorganic carbonate (CaCO 3 ) typically forms in isotopic equilibrium with the DIC pool and, as such, is indirectly affected by organic matter sources and primary production and respiration rates (modified from Meyers and Teranes, 2001). Introduction Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles the most productive lakes emit more CO 2 into the atmosphere than can be sequestered. This is because carbon inputs prevent carbon limitation (Talling, 1976). 1.4 Aims The aims of this PhD Thesis are twofold: a) to explore the possibilities that the study of δ18O diatom and δ13C diatom can offer in palaeoenvironmental reconstructions, and b) to carry out high- and ultra-high resolution environmental and climate reconstructions in the Andean Altiplano during the Late Glacial- Early Holocene transition using these stable isotopes. 1.5 Thesis structure This PhD thesis is concerned with an environmental reconstruction based on the study of stable isotopes in diatom silica from lake sediments. Chapter 1 provides a state-of-the-art introduction to lakes, diatoms and isotopes. The regional geographical and geological settings as well as the modern and ancient climates of the Central Andes and the studied site are presented in Chapter 2. This thesis forms part of a major work carried out by a multidisciplinary research group with scientists belonging to the Institut de Ciències de la Terra, ‘Jaume Almera’- Consejo Superior de Investigaciones Científicas (ICTJA-CSIC), the Universities of A Coruña (UDC), Barcelona (UB) and Católica del Norte (UCN), the Instituto Pirenaico de Ecologia (IPE-CSIC), the NERC Isotopes Geosciences Laboratory (NIGL) and the Lancaster Environmental Center (LEC). Chapter 2 sets out the scientific context of the thesis. The methodological procedures employed are descibed in Chapter 3. A detailed petrographical study is a prerequisite for verifying the reliability of the sedimentary record for a robust palaeoenvironmental reconstruction. This is developed in Chapters 4 and 6. The thesis also focuses on new and poorly documented fields where δ18O diatom and δ13C diatom can successfully be applied to lacustrine sediments. It shows how stable isotopes from diatom silica may be used a) to highlight the importance of reconstructing the different evolutionary stages of lake ontogeny given that climate derived palaeohydrological signals can be distorted by changes in lake morphology (Chapter 4); b) as a main proxy in ultra-high resolution moisture balance reconstructions forced by fluctuations in the intensity of the ENSO and solar activities (Chapter 5); c) to reveal the major biogeochemical processes that give rise to the formation of rhythmites (Chapter 6), and finally d) to reconstruct the regional environmental evolution at centennial-to-millenial time scales (Chapter 7). The conclusions are grouped according to their methodological, limnological or climate implications and are presented in Chapter 8. Finally, the raw data obtained during the experimental processes, and the original published papers are contained in the appendices. 26 27 Chapter 2 Geological, geographical, limnological and climate framework 2.1 Geographical and geological setting The Andes (8,000 km long and up to 750 km wide), the dominant landform of South America, extends along the entire western coast (Fig. 2.1). They are the result of the oblique subduction of the Nazca Plate (oceanic plate) beneath the South-American Plate (continental plate) (Dewey and Bird, 1970; James, 1971; Allmendinger and Jordan, 1997). Subduction began soon after the breakup of Rodinia in Late Proterozoic times, and since then it has been intermittently active up to the present (Ramos, 2009). The highest peaks in the southern tropical and subtropical Andes exceed altitudes of 6,000 m; ice caps and glaciers are present where the atmospheric circulation systems provide sufficient moisture for snow and ice accumulation (Zech et al. 2009) (Fig. 2.1). The Andes constitute a unique physiogeographical setting with vertical gradients ranging from sea level up to peaks that exceed 6,000 m in less than 150 km. This mountain range resulted from a Cenozoic tectonic uplift in the forearc region of the active tectonic convergence zone (Stecker et al. 2007; Grosjean et al. 2007). A longitudinal tectonic segmentation that has been ascribed to dip variations in the flat subduction of the Nazca Plate is observed in the Andes. Structures involving the basement (thick-skinned) with no volcanism are developed in the low-angle subduction zones, whereas thin-skinned structures associated with volcanism are present in high-angle subduction zones (Carrera, 2009). The geological features of this active orogen resulted from a succession of orogenic cycles that probably started during the Late Precambrian and continued during Phanerozoic times (Ramos, 1994, 2009). Two major tectono-sedimentary super-cycles are commonly distinguished in the Andes: the Pre- Andean Cycle (mostly Paleozoic-Early Triassic in age) and the Andean Cycle (mostly Permian-Early Triassic to Cenozoic), which is responsible for the present Andean uplift (Coira et al. 1982). According to the main features of the subducted plate, the Andes has been sub-divided into different geological regions: Northern Andes (5-15º S), Central Andes (15-33.5º S), Southern Andes (33.5-47º S) and Austral Andes (47-56º S) (Tassara and Yañez, 2003). Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 2.1.1 The Central Andes The geological region of the Central Andes is where the Andean orogen attains its maximum height and width. Tectonic, sedimentary and magmatic processes closely interplayed in the Central Andes, which can be sub-divided into different morphostructural units. From West to East, these include Cordillera de la Costa, Valle Longitudinal, Precordillera Andina, Cordillera Occidental, Altiplano or Puna, Cordillera Oriental, and in the easternmost part before reaching the Chaco, from North to South, the Sierras Subandinas, Sistema de Santa Bárbara or Sierras Pampeanas (Coutand et al. 2001) (Fig. 2.2). The Andean Altiplano is a segment of the Andean orogenic belt between 14º and 27º S, which consists of Northern Chile and Argentina, Western Bolivia and Southern Perú (Isacks, 1988; Strecker et al. 2007) (Fig. 2.1). At present, this high-elevation Plateau (~3,500 – 4,100 m) is an area of inland drainage (endorheic) formed during an uplift ~30 Myrs ago (Gregory-Wodzicki, 2000; Orme, 2007) owing to the compression of the western rim of the South American Plate due to the subduction of the Nazca Plate Figure 2.1. Satellite images of A) South America, B) Andean Altiplano, C) Lago Chungará. Square on the image depicts the position of the following image. Numbers indicate the position of the main geographical features shown in the pictures: 1) Lago Titicaca (Perú-Bolivia), 2) Lago Popoó (Bolivia), 3) Nevado Sajama (Bolivia), 4) Salar de Uyuni (Bolivia), 5) Volcan Ajoya (Chile), 6) Bofedal de Parinacota (Chile), 7) Vocan Quisiquisini (Chile-Bolivia), 8) Volcan Parinacota (Chile), 9) Lago Chungará (Chile), 10) Lagunas de Cotacotani (Chile). 28 Lago Titicaca Lago Chungará Bofedal de Parinacota Lagunas de Cotacotani Volcán Parinacota Volcán Ajoya Volcán Quisiquisini Salar de Uyuni Lago Poopó Nevado Sajama 9 9 5 8 6 7 10 3 4 1 2 5 86 7 21 43 10 A B C 29 C o rd ille ra O c c id e n ta l A ltip la n o P u n a Sierras Pampeanas P re c o rd ille ra Sistema de Sta. Bárbara Volcanic arc C o d ille ra d e la c o s ta Chaco P e rú -C h il e T re n c h V a lle lo n g itu d in a l S ie rra s S u b a n d in a s Cordillera Oriental C o rd ille ra O c c id e n ta l 25º 20º 15º 65º70º75º NAZCA PLATE Cordillera de la costa Precordillera Cordillera Occidental Puna (Altiplano) Cordillera oriental Sistema de Sta. Bárbara Valle longitudinal A B Figure 2.2. A) Main morphostructural units from the Central Andes (from West to East): Cordillera de la Costa, Valle Longitudinal, Precordillera Andina, Cordillera Occidental, Altiplano or Puna, Cordillera Oriental, and, from North to South, Sierras Subandinas, Sistema de Santa Bárbara or Sierras Pampeanas, and the Chaco. Note the position of Lago Chungará. Modified from Coutand et al. (2001) and Amilibia (2002). B) General cross-section of the Central Andes at 25.5º S. Modified from Muñoz et al. (2004). and the Antarctic Plate (Allmendinger et al. 1997). It attains a width of 350–400 km and is dominated by massive active volcanoes (Wörner et al. 2000). The most prominent features that define this non- collisional orogen are the subducting Nazca plate and its associated trench, the present day active volcanic arc and the fold-and-thrust belt in the foreland (Whitman et al. 1996). The modern lithospheric thickness is roughly 150 km beneath the Altiplano (Whitman et al. 1996). The Andean Altiplano is characterised by the presence of thick syn-orogenic sediments (up to 3 km) lying directly over the Precambrian or Paleozoic basement (Carrera, 2009). These syn-orogenic sediments were deposited and preserved in intra-mountain basins. Their growth geometries and disconformities show that the development of the present thrusts began during the Paleogene and continued developing until the Present (Ramos, 1999; Coutand et al. 2001). Geological, geographical, limnological and climate framework Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles Lago Chungará region Lago Chungará (18º15’S, 69º10’W, 4,520 m asl) is located on the Andean Altiplano (Central Andes, Chile) in a highly active tectonic and volcanic context (Clavero et al. 2002; Hora et al. 2007). It lies on the northeastern edge of the Lauca basin, which is the westernmost and also the highest fluvio-lacustrine basin of the Altiplano. Lago Chungará is filled with an Upper Miocene to Pliocene volcaniclastic alluvial and lacustrine sedimentary succession (> 120m thick) that rests unconformably on an Upper Cretaceous– Lower Miocene volcanic substrate (Kött et al. 1995; Gaupp et al. 1999). In particular, the lake forms part of the small hydrologically closed Chungará Sub-Basin. It came into existence as a result of a debris avalanche during the partial collapse of the Parinacota Volcano, which dammed the former Lauca River. Lago Chungará and Lagunas Cotacotani were formed almost immediately (Fig. 2.1). However, the age of this collapse is not well constrained, with estimates ranging from 13,000 to 20,000 years BP (Wörner et al. 2000; Clavero et al. 2002, 2004; Hora et al. 2007). The area surrounding Lago Chungará and Lagunas Cotacotani has been characterised by intermittent volcanic activity since the Miocene until the present (Katsui and González-Ferrán, 1968; Wörner et al. 1988; Clavero et al. 2002). There are three main volcanoes surrounding the lake: Parinacota, Ajoya and Quisiquisini (Fig. 2.1). Parinacota volcano is a conic and symmetric stratovolcano (6,342 m asl) with excellent lava flows and dome preservation. The most distal lava flows of the southern volcano flank coincide with the northern boundary of Lago Chungará. Lagunas Cotacotani, which are located on debris avalanche deposits on the western volcano flank, formed during the collapse of Parinacota volcano (Clavero et al. 2002) (Fig. 2.1). The most recent lava flows of the Parinacota volcano are located on the southern and western flanks and are composed of dark grey-colour basaltic-andesitic lavas. Parinacota has been the only active volcano in the Lago Chungará watershed during the Holocene (Wörner et al. 1988; de Silva and Francis, 1991). The Miocene-age eroded Ajoya volcano is situated on the southern boundary of Lago Chungará (Fig. 2.1) and its grey-colour lavas are mainly andesitic with porphyritic textures of plagioclase, hornblende and pyroxene phenocrystals (Katsui and González-Ferrán, 1968). Despite the presence of an ancient crater at the top, it is not possible to identify the lava flows of this volcano. The Pleistocene-age Quisiquisini volcano is located along the eastern boundary of Lago Chungará. It is also very eroded and is largely made up of andesitic lavas and tuffs (Katsui and González-Ferrán, 1968). 2.2 Climate Framework Given its vast extension across the equator from about 10ºN to 55ºS, South America displays a large distribution of tropical to extratropical climates. Superimposed on the mean north-to-south 30 31 variations are significant east–west asymmetries across the continent due to a) the presence of the Andes, b) changes in the continental width (broad at low- latitudes, narrow at mid-latitudes), and c) the boundary conditions imposed by the cold south-eastern Pacific and the warm south-western Atlantic oceans (Sylvestre, 2009). The sea surface temperature (SST) anomalies in association with the ENSO, the Pacific Decadal Oscillation (PDO), the Antarctic Oscillation (AAO), or the North Atlantic oscillation (NAO) significantly contribute to the climate variability of the continent (Garreaud et al. 2009). 2.2.1 Modern Climate The InterTropical Convergence Zone (ITCZ) is a belt of low pressure created by the convergence of northeast and southeast trade winds over the equatorial oceans (Barry and Chorley, 1992; McGregor and Nieuwolt, 1998). This zone is east–west oriented over the tropical oceans (Fig. 2.3). The rainfall in South America exhibits the highest values along the ITCZ over the continent (Garreaud et al. 2009). This rainfall is mainly convective and is produced by deep cumulus–nimbus (Mitchell and Wallace, 1992). During summer months (December to February), strong surface heating in the Southern Hemisphere results in the southerly location of the ITCZ, whilst strong surface heating in the Northern Hemisphere between July and August causes the ITCZ to migrate north of the Equator (McGregor and Nieuwolt, 1998). Thus, during the austral winter the north of the equator experiences maximum rainfall in accordance with the oceanic ITCZ, whereas the central part of the continent (including southern Amazonia) undergoes its dry season (Garreaud et al. 2009). During the austral fall, the rainfall maximum gradually returns to northern South America. Such migration has led many scientists to describe the climate of the central part of South America as monsoon-like (Zhou and Lau, 1998; Vera et al. 2006). The climate is not fully monsoonal, however, because the low-level winds never reverse their direction. In contrast to the copious precipitation near the ITCZ, rainfall is almost absent over broad areas of the subtropical oceans owing to the large-scale mid-tropospheric subsidence (Garreaud et al. 2009). At present, the seasonal rainfall cycle over South America is dominated in tropical latitudes by the so-called South American Summer Monsoon (SASM) (Zhou and Lau, 1998). Rainfall over the Andean Altiplano originates primarily from the Atlantic Ocean, and humidity is advected into the Amazon Basin by the north-easterly SASM, which is associated with the ITCZ (Zhou and Lau, 1998; Vuille and Werner, 2005). Humidity is then transported further south along the eastern slopes of the Andes by the South American Low Level Jet (Saulo et al. 2000; Marengo et al. 2004). The deep moist convection, especially during the austral summer, is ascribed to the distinctive atmospheric pressure systems known as the near-surface Chaco Low and the tropospheric Bolivian High (Aceituno and Montecions, 1993; Lenters and Cook, 1997). Ultimately, upper-tropospheric easterly anomalies, which are modulated by the Pacific Geological, geographical, limnological and climate framework Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles SSTs, transport the humidity to the Cordilleras and onto the Altiplano (e.g. Bradley et al. 2003; Garreaud et al. 2003; Vuille and Keimig, 2004; Falvey and Garreaud, 2005). El Niño-Southern Oscillation-related climate variability Among the many forcings that determine the interannual climate variability in South America (e.g. PDO, AAO), the ENSO plays a major role in many regions. The ENSO is a coupled ocean-atmosphere phenomenon rooted in the equatorial tropical Pacific Ocean and is characterised by irregular fluctuations between its warm (El Niño) and cold (La Niña) phases with a periodicity ranging from 2 to 7 years (Diaz and Markgraf, 1992; Cane, 2005; McPhaden et al. 2006) (Fig. 2.4). Rainfall and temperature anomalies associated with the occurrence of the El Niño and La Niña events are the major source of interannual variability over much of South America (e.g., Aceituno, 1988; Marengo, 1992; Antico, 2009; Hill et al. 2009). The overall pattern is that the El Niño episodes are associated with a) below normal rainfall over tropical South America, b) above normal precipitation over the southeastern part of the continent and central Chile, and c) warmer than normal conditions over tropical and subtropical latitudes. Contrasting rainfall and temperature anomalies are observed during the La Niña episodes (Garreaud et al. 2009). Otherwise, the most plausible forcing behind the low-frequency fluctuations is the PDO, an enduring pattern of Pacific climate variability (e.g. Mantua et al. 1997; Garreaud and Battisti, 1999; Villalba et al. 2001). The PDO is often described as ENSO-like because the spatial climate fingerprints of its warm and Figure 2.3. Location of Lago Chungará on a rainfall rate map (mm/year) of South America simplified from Negri et al. (2004). The main atmospheric systems are indicated. ICTZ: Intertropical Convergence Zone, and SPCZ: South Pacific Convergence Zone. 32 30ºW 60ºS 45ºS 30ºS 15ºS 15ºN 0º 60ºW South Pacific Anticyclone ITCZ ITCZ SPCZ South American Summer Monsoon Lago Chungará 90ºW <120 1000 1000 4800 1000 1000 4800 4800 33 cold phase bear a strong resemblance to those of the El Niño and the La Niña events, respectively (e.g. Garreaud and Battisti, 1999; Barichivich et al. 2009). However, the causes of the PDO and its links with the ENSO are not yet fully understood (Newman et al. 2003; Schneider and Cornuelle, 2005). Present-day climate in the Lago Chungará region The climate in the Chungará-Cotacotani lake district is dominated by arid conditions due to the influence of the South Pacific Anticyclone (Fig. 2.3). The modern mean annual temperature at Lago Chungará is +4.2ºC (Figur3 2.5). The annual rainfall ranges from 100 to 750 mm year-1 (mean 411 mm year-1) whereas the potential evaporation has been estimated at 1200 mm year-1, which exceeds the precipitation (Valero-Garcés et al. 2000). The lake region shows pronounced seasonal contrasts owing to the dominance of the SASM (Garreaud et al. 2003; Vuille and Werner, 2005). This climate situation defines the wet season (December-March) in the Altiplano, accounting for more than 70% of the annual precipitation. Instrumental data from the Chungará area show a reduction of precipitation during moderate to intense El Niño years (Fig. 2.5). However, there is A C El Niño conditions La Niña conditions Convective loop Convective loop Convective loop Equator Equator Thermocline Thermocline 120ºE 120ºE 80ºW 80ºW Convective loop Equator 120ºE 80ºW Thermocline Neutral conditions B Figure 2.4. Atmospheric circulation and thermocline position in the Pacific Ocean zone between America and Australia during A) El Niño events; B) La Niña events; and C) Neutral conditions. Red and green colours represent warm and cold SST, respectively. Modified from http://www.pmel.noaa.gov/tao/elnino/nino-home.html Geological, geographical, limnological and climate framework Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles no direct relationship between the relative El Niño strength and the amount of rainfall reduction (Valero- Garcés et al. 2003). Furthermore, at longer timescales, it is postulated that changes in the tropical-Atlantic meridional SST gradients also govern rainfall variability on the Altiplano (Baker et al. 2001b). The rainfall isotope composition in the Central Andes is characterised by a large variability in δ18O (between +1.2 and –21.1‰) and of δD (between +22.5 and –160.1‰). The origin of the lightest oxygen isotope values is the strong fractionation in the air masses from the Amazon and is directly related to higher rainfall intensity (‘amount effect’) (Herrera et al. 2006). However, the rainfall oxygen isotope composition in the Chungará-Cotacotani lake district only shows a variation of 6‰, between – 14‰ and –20‰, with a mean value of –14.3‰ (Fig.2.5). R a in fa ll (m m ) T e m p e ra tu re (ºC ) Rainfall Temperature 20 1 40 2 60 3 80 4 100 5 120 6 140 7 Months 0 0 1 2 3 4 5 6 7 8 9 10 11 12 Year 800 700 600 500 400 300 200 100 0 A n n u a l R a in fa ll (m m ) 60 65 70 75 80 85 90 95 Chucuyo (18º15’S, 69º20’W, 4,200 m a.s.l.) Chungará (18º17’S, 69º08’W, 4,500 m a.s.l.) Cotacotani (18º12’S, 69º14’W, 4,500 m a.s.l.) -20 -16 -12 -8 -4 0 -140 -120 -100 -80 -60 -40 -20 0 Lago Chungará Springs Río Chungará Rainfall Hail  D ( S M O W ) ‰ 1  8 O ( SMOW)‰ LML GMWL EL A B C Figure 2.5. A) Isotope values (δ18O/δD) from rainfall, Lago Chungará, Río Chungará and springs. GMWL: Global Meteoric Water Line; LML: Local Meteoric Line; EL: Evaporation Line. Note the isotopic enrichment of the lake water by evaporation with regard to rainfall and springwater. Modified from Herrera et al. (2006). B) Mean monthly rainfall (mm) and temperature (ºC) at the Chungará meteorological station (18.17ºS, 69.08ºW, 4,500 m asl). Note the seasonality of both parameters. C) Annual rainfall in the Chungará region from 1962 to 1994. The arrows indicate strong El Niño years. Modified from Valero-Garcés et al. (2003). 2.2.2 Last Glacial-Holocene climate variability in the Central Andes Until recently, few palaeoclimatic studies have been carried out in the Central Andes, and the basic character of climate change during the Late Quaternary was poorly documented. In the last two decades, important studies have been published (e.g. Thompson et al. 1995; Betancourt et al. 2000; Barker et al. 2001a,b) without reaching a consensus on some key questions (e.g. Rigsby et al. 2005; Villalba et al. 2009). 34 35 Geological, geographical, limnological and climate framework The Last Glacial Maximum (LGM) appears in several Altiplano records and is dated at approx. 21,000 cal years BP (e.g. Baker et al. 2001a; Placzek et al. 2006), but whether it was arid (e.g. Heine, 2000; Mourguiart and Ledru, 2003) or wet (e.g. Baker et al. 2001a,b; Tapia et al. 2003) remains unclear. A number of authors suggest a warmer and wet situation throughout the succeeding Late Glacial (e.g. Thompson et al. 1998; Baker et al. 2001a; Rigsby et al. 2005; Hyllier et al. 2009) (Fig. 2.6). Changes in the tropical circulation system are responsible for the massive precipitation increase in Late Glacial times. They are probably linked to a southward displacement of the ITCZ, which gives rise to an enhanced moisture transport onto the Altiplano (Rowe et al. 2002). The Late Glacial-Early Holocene transition is generally presented as a period of change towards drier conditions with minor wet events (e.g. Abbot et al. 2003; Tapia et al. 2003; Rigsby et al. 2005). However, there is no clear consensus on whether climate periods such us the Northern Hemisphere’s Bølling-Allerød-like or the Younger Dryas-like could have co-existed in the Central Andes. These short millennial events are reflected in Lake Titicaca (Perú-Bolivia), Salar de Uyuni (Bolivia) and Nevado Sajama (Bolivia) (e.g. Thompson et al. 1998; Baker et. al 2001a; Placzek et al. 2006) but there are no records of these events in other places (e.g. Abbott et al. 2003; Hillyer et al. 2009). By contrast, there is a general agreement that a strong arid period throughout the last 30,000 cal years BP occurred during the Early-to-Mid Holocene (e.g. Wolfe et al. 2001; Grosjean et al. 2007; Giralt et al. 2008), but this period does not occur at the same time on the Andean Altiplano (Betancourt et al. 2000; Grosjean et al. 2003). Abbott et al. (2003) found an approximately 2000-year lag in the timing of this dry event, with more northerly lakes responding earlier. In the Lago Pachuca region (Perú), the driest period began earlier (at ~10,000 cal years BP) than in the southern region (Lago Chungará, Lago Poopó (Bolivia) and Salar de Uyuni), which did not begin earlier than 8,000 cal years BP (Fig. 2.6). The amplitude of this period also differs depending on the record (Abbott et al. 2003; Grosjean et al. 2003; Placzek et al. 2006). The most plausible explanation of the Mid-Holocene aridity does not seem to be an insolation minimum, which occurred earlier (Fig. 2.6), but the establishment of a weakening in the ENSO variability between ca 8,000 and 5,000 cal years BP (Rodbell et al. 1999; Rein, 2007). The additive effect of a weakened ENSO and the slowly increasing insolation during the wet season may have thrown the region into a drought-prone stage (Baker et al. 2001a; Abbott et al. 2003; Paduano et al. 2003). Finally, the Late Holocene period showed a general return to moister conditions probably attributed to enhanced precipitation, decreased evaporation, a shorter dry season or a combination of all these factors (Marchant and Hooghiemstra, 2004). This period is however commonly considered highly oscillating because the lake water level increases are punctuated by minor lake level drops (e.g. Valero- Garcés et al. 1996; Grosjean et al. 1997; Baker et al. 2001a; Rigsby et al. 2005; Hyllier et al. 2009). These changes would be ascribed to orbital forcing, which resulted in a strengthening of wet-season convection Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles as summer insolation increased during the Late Holocene (Abbot et al. 2003). Overprinting these centennial- to millennial-scale climate shifts are higher-resolution changes that are not directly attributed to insolation forcing. These fluctuations in the regional water balance are due to changes in the tropical Pacific SST and are therefore linked to the ENSO activity. The pacific SST changes give rise in the Altiplano to dry conditions during the El Niño phase and to wet conditions during the La Niña phase (Vuille, 1999; Vuille et al. 2000; Valero-Garcés et al. 2003). In summary, the timing of wet and dry periods over the Andean Altiplano is broadly consistent with the insolation inducing precipitation (e.g. Martin et al. 1997; Betancourt et al. 2000; Seltzer et al. 30 28 26 24 22 20 18 16 14 12 10 8 6 4 2 0 C O L D C O L D C O L D W A R M W A R M W A R M H O L O C E N E C L IM A T E P E R IO D S L G M B -A Y D G L A C I A L L ag o P ac u ch a (1 3º S ) H ill y e r e t a l. 2 0 0 9 A b b o tt e t a l. 2 0 0 3 T a p ia e t a l. 2 0 0 3 R ig s b y e t a l. 2 0 0 5 B a k e r e t a l. 2 0 0 1 a A b b o tt e t a l. 2 0 0 3 W o lf e e t a l. 2 0 0 1 T h o m p s o n e t a l. 1 9 9 8 G ir a lt e t a l. 2 0 0 9 S y lv e s tr e e t a l. 1 9 9 9 B a k e r e t a l. 2 0 0 1 b P la c z e k e t a l. 2 0 0 6 F ri tz e t a l. 2 0 0 4 L ag o P ac o C o ch a (1 3º 50 ’S ) Lago Titicaca (16ºS) Lago Taypi Chaca Kkota (16ºS) N ev ad o S aj am a (1 8º S ) L ag o C h u n g ar á (1 8º S ) S al ar d e C o ip as a L ag o P o o p ó (1 9º S ) Salar de Uyuni (20ºS) Insolation 18ºS Dec (Wm ) -2 Wet conditions Intermediate conditions Dry conditions 485450 Figure 2.6. Compilation of palaeoclimatic records from the Andean Altiplano and insolation curve for the austral summer at 18ºS to account for the main climate patterns in terms of wet-dry conditions during the last 30,000 cal years BP. LGM: Last Glacial Maximum; B-A: Bølling-Allerød chronozone; YD: Younger Dryas chronozone 36 37 2000; Baker et al. 2001a,b; Cross et al. 2001)(Fig. 2.6). Maximum insolation during the austral summer (20,000 cal years BP and present) gave rise to the highest deep convection over southern tropical South America (maximum intensity of the SASM), which is consistent with wet conditions on the Altiplano during the LGM and the Late Holocene (Rigsby et al. 2005). Conversely, the summertime insolation minimum at 10,000 cal years BP ushered in an arid phase on the Altiplano (Seltzer et al. 2000; Tapia et al. 2003). Furthermore, the most extreme climate events over the Andean Altiplano did not occur during the terminal ice age, but during the Holocene. Hence, regardless of human influence, the rates of climate change appear to be faster in the Holocene than in Glacial or Deglacial times (Hillyer et al. 2009) 2.3 Limnological features of Lago Chungará Lago Chungará has an irregular shape, with a maximum length of 8.75 km, maximum water depth of 40 m, a surface area of 21.5 km2 and a water volume of ca 400 Hm3 (Mühlhauser et al. 1995; Herrera et al. 2006) (Fig. 2.7). The western and northern lake margins are steep, constituted by the eastern slopes of the Ajoya and Parinacota volcanoes. The eastern and southern margins are gentle, formed by the distal fringe of recent alluvial fans and the River Chungará valley (Sáez et al. 2007). The morphology of the lake floor has been determined by bathymetric data (Villwock et al. 1985) and seismic profiles (Valero-Garcés et al. 2000). Six morphological components can be differentiated along a west-to-east profile (Fig. 2.7): a) a narrow western littoral platform, ca 175–300 m wide, 0–7 m deep (slope <1º); b) a western slope, 115 m wide and 7–20 m deep, dipping 10º; c) a 2–3º rise at the base of the slope, 115– 235 m wide and 25–40 m deep; d) a central plain, 4 km wide and 25–40 m deep; e) an eastern slope of 3º, 200 m wide and between 7 and 25 m deep; and f) a subhorizontal (<1º) eastern platform, 450–850 m wide and between 0 and 7 m deep. The absence of high-level shorelines along the lake margins suggests that the current level of the lake is the highest since its formation (Sáez et al. 2007). At present, the main inlet to the lake is the Chungará River (300-460 l s-1) although small streams flow into the lake along the south-western margin. Water inputs to the lake have, on average, the following composition: 42 ppm HCO 3 - , 3 ppm Cl-, 17 ppm SO 4 2- , 7 ppm Na+, 4 ppm Mg2+, 8 ppm Ca2+, 3 ppm K+ and 22 ppm Si. The Mg:Ca ratio of water inputs ranges from 0.22 to 0.71, depending on the local lithology of the catchment (Herrera et al. 2006, Sáez et al. 2007). Evaporation constitutes the main water loss (3.107 m3 year-1). There is no surface outlet, but groundwater outflow from Lago Chungará to Lagunas Cotacotani (6-7·106 m3year-1) (Dorador et al. 2003) represents about 20% of the total outflow. The calculated residence time for the lake water is approximately 15 years (Herrera et al. 2006). The lake, which can be regarded as polymictic and oligo- to meso-eutrophic, contains 1.2 g l-1 TDS (total dissolved solids); its conductivity ranges between 1,500 and 3,000 μS cm-1 (Dorador et al. 2003), Geological, geographical, limnological and climate framework Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles and its water chemistry is of the Na+-Mg 2 +-HCO 3 -SO 4 2- type and alkaline (Table 2.1). The δ18O and δD composition of the lake water revealed that it diverges from the Global Meteoric Water Line (GMWL) and the Regional Meteoric Line (RML) (Herrera et al. 2006). This divergence can be attributed to the enrichment of the lake water due to evaporation with regard to rainfall and springwater (Herrera et al. 2006; Hernández et al. 2008). The mean lake water values of δ18O and δD are –1.1‰ SMOW and – 39.2‰ SMOW, respectively (Table 2.1 and Fig. 2.5). Present day primary productivity is mainly governed by diatoms and chlorophyceans during the cold and warm seasons, respectively (Dorador et al. 2003). Seasonal measurements of conductivity, nitrate, phosphate and chlorophyll reveal that these changes in productivity and in the composition of algal communities are mainly due to variations in water temperature and salinity (Dorador et al. 2003). Dense macrophytic vegetation patches and microbial colonies are present in the littoral zone, also contributing to primary productivity (Dorador et al. 2003). The local vegetation is dominated by tussock- like grasses, shrubs, Polylepis, dwarf trees of the Rosaceae family as well as extensive soligenous peatlands (‘‘bofedales’’) (Schwalb et al. 1999; Earle et al. 2003). Finally, the fauna includes endemic cyprinodontid fish (e.g. 19 species of Orestias) (Villwock et al. 1985). SUBUNIT 1a. Laminated diatomaceous ooze SUBUNIT 1b. Carbonate laminated diatomaceous ooze SUBUNIT 2b. Mafic minerals-rich diatomaceous ooze and tephra layers Peat, silt and diatomaceous oozes, rich in charophytes and mollusc Gravel, sand and sitl, including deformed massflows Pre-Lacustrine Quaternary Fluvial-Alluvial deposits Miocene volcanic rocks 40 m 30 m 20 m 10 m 0 m Central plain E platformRise YX SUBUNIT 2a. Diatomaceous ooze, carbonate and tephra layers 2 km 32m 27m 17m Chungará river2 Km 7m 7m 1 2 3 4 5 7 8 19939 10 13 14 15 11 12 E slope E platform W platform W slope Rise Central plain 69 10’ W o 18 15’ S o X Y Lago Chungará Parinacota volcano NW-SEA B C Figure 2.7. A) Panoramic view of Lago Chungará. B) Bathymetric map of Lago Chungará showing the main morphological units of the lake floor cited in the text, and the position of the recovered cores. Black line indicates the cross section (C) throughout the lake. C) Cross section of sediment infilling of Lago Chungará. Position of the studied core is shown. Note that the position of the core is projected in its equivalent position in the central plain. Arrows indicate major hydrological inputs and sedimentary contributions to the lake. Simplified from Sáez et al. (2007). 38 39 Geological, geographical, limnological and climate framework S a m p le T e m p a ir (º C ) T e m p w a te r (º C ) p H C o n d u c ti v it y   s /c m ) S a li n it y (m g /l ) C l (p p m ) S O 4 (p p m ) H C O 3 (p p m ) B (p p m ) N a (p p m ) K (p p m ) C a (p p m ) M g (p p m ) S i (p p m ) P (p p m ) L i (p p m )  D (S M O W )  1 8 O (S M O W ) L a k e C h u n 1 - 0 m 9 .7 1 2 9 .1 9 1 ,5 0 1 7 4 3 8 1 4 2 2 3 6 5 0 ,9 1 6 2 ,9 3 2 ,2 5 0 ,6 1 0 5 ,4 0 ,6 0 ,7 0 ,2 -4 0 ,0 5 -1 ,8 5 C h u n 1 - 2 m 9 .7 1 1 .4 9 .1 9 1 ,4 9 5 7 4 1 7 4 4 0 5 3 6 4 0 ,9 1 6 2 ,1 3 2 ,1 5 0 ,5 1 0 4 ,5 0 ,7 0 ,7 0 ,2 -3 8 ,8 -1 ,5 9 C h u n 1 - 4 m 9 .7 1 0 .5 9 .1 8 1 ,4 9 6 7 4 1 7 4 4 0 6 3 5 6 0 ,9 1 6 0 ,2 3 2 ,0 4 9 ,9 1 0 2 ,8 1 ,0 0 ,7 0 ,2 -3 9 ,4 -1 ,8 1 C h u n 1 - 6 m 9 .7 1 0 9 .1 7 1 ,4 9 0 7 3 7 7 4 3 9 6 3 6 4 0 ,9 1 5 9 ,1 3 1 ,7 4 9 ,7 1 0 2 ,2 1 ,3 0 ,7 0 ,2 -3 9 ,9 -1 ,0 3 C h u n 1 - 8 m 9 .7 9 .8 9 .1 5 1 ,4 9 0 7 3 5 7 4 4 0 5 3 6 3 0 ,9 1 5 9 ,0 3 1 ,7 5 0 ,0 1 0 2 ,0 1 ,3 0 ,7 0 ,2 -3 9 ,6 -0 ,2 9 C h u n 5 - 0 m 1 7 1 2 .1 9 .2 7 1 ,4 6 3 7 2 5 7 4 3 8 4 3 3 9 0 ,9 1 5 8 ,9 3 1 ,1 4 8 ,9 1 0 1 ,1 0 ,9 0 ,7 0 ,2 -4 1 -0 ,0 1 C h u n 9 - 0 m 1 4 .2 1 1 9 .1 8 1 ,5 0 5 7 4 6 7 9 4 0 3 3 5 9 0 ,9 1 6 3 ,4 3 1 ,9 5 0 ,6 1 0 5 ,2 0 ,6 0 ,8 0 ,2 -3 8 ,2 0 ,0 7 C h u n 9 - 2 m 1 4 .2 1 1 9 .1 6 1 ,4 9 8 7 4 2 7 9 4 0 0 3 6 4 0 ,9 1 6 3 ,7 3 2 ,1 5 0 ,9 1 0 5 ,3 0 ,6 0 ,7 0 ,2 -3 7 ,4 -1 ,6 9 C h u n 9 - 4 m 1 4 .2 1 0 .4 9 .2 0 1 ,5 0 2 7 4 5 6 9 4 0 2 3 5 9 0 ,9 1 6 4 ,1 3 2 ,1 5 0 ,9 1 0 5 ,4 0 ,6 0 ,8 0 ,2 -3 8 ,3 -1 ,9 8 C h u n 9 - 6 m 1 4 .2 1 0 .6 9 .1 8 1 ,5 0 2 7 4 4 7 5 3 9 1 3 6 4 0 ,9 1 6 4 ,4 3 2 ,2 5 0 ,6 1 0 5 ,3 0 ,6 0 ,8 0 ,2 -3 7 ,6 -1 ,6 8 C h u n 9 - 8 m 1 4 .2 1 0 .4 9 .1 9 1 ,4 9 9 7 4 3 7 6 3 9 0 3 6 2 0 ,9 1 6 3 ,4 3 2 ,0 5 0 ,5 1 0 5 ,0 0 ,6 0 ,7 0 ,2 -3 8 ,4 -1 ,6 7 C h u n 9 - 1 0 m 1 4 .2 1 0 .2 9 .1 9 1 ,5 0 5 7 4 6 7 5 3 9 6 3 6 1 0 ,9 1 6 4 ,8 3 2 ,3 5 1 ,2 1 0 6 ,4 0 ,6 0 ,8 0 ,2 -3 8 -1 ,4 8 C h u n 9 - 1 2 m 1 4 .2 1 0 9 .1 7 1 ,5 1 0 7 4 7 7 5 3 9 1 3 5 8 0 ,9 1 6 3 ,6 3 1 ,9 5 0 ,6 1 0 4 ,9 0 ,7 0 ,8 0 ,2 -3 7 ,5 0 ,1 4 C h u n 9 - 1 4 m 1 4 .2 9 .8 9 .1 7 1 ,5 1 0 7 4 6 7 3 3 8 9 3 5 7 0 ,9 1 6 4 ,6 3 2 ,2 5 0 ,9 1 0 5 ,5 0 ,7 0 ,7 0 ,2 -3 6 ,8 0 ,0 7 C h u n 9 - 1 6 m 1 4 .2 1 0 .6 9 .1 5 1 ,4 9 2 7 3 9 7 2 3 9 0 3 6 3 0 ,9 1 6 4 ,3 3 2 ,3 5 0 ,4 1 0 4 ,7 0 ,7 0 ,8 0 ,2 -3 6 ,6 -2 ,0 9 C h u n 9 - 1 8 m 1 4 .2 9 .8 9 .1 3 1 ,4 8 7 7 3 1 7 4 3 8 7 3 6 4 0 ,9 1 6 2 ,3 3 2 ,2 5 0 ,2 1 0 4 ,7 0 ,8 0 ,8 0 ,2 -3 5 ,7 0 ,5 5 C h u n 9 - 2 0 m 1 4 .2 9 .5 9 .1 5 1 ,5 0 7 7 4 7 7 4 3 8 8 3 6 3 0 ,9 1 6 2 ,5 3 2 ,2 5 0 ,2 1 0 3 ,8 0 ,8 0 ,7 0 ,2 -3 6 ,4 0 ,1 5 C h u n 1 0 - 0 m 1 0 1 8 .8 9 .0 6 1 ,0 6 8 5 2 3 6 3 2 3 9 2 4 6 0 ,8 1 1 3 ,7 2 5 ,9 3 9 ,9 6 6 ,3 1 0 ,8 1 ,0 0 ,2 -5 5 ,2 -3 ,1 2 S m a ll L a k e s C h u n 2 - 0 m 1 7 1 4 .9 1 0 .7 6 1 ,8 3 0 9 1 4 1 1 7 5 4 3 2 3 1 1 ,3 2 2 9 ,5 4 3 ,0 2 9 ,2 1 3 8 ,7 0 ,1 0 ,4 0 ,3 -1 2 ,9 5 6 ,2 4 C h u n 7 - 0 m 1 1 1 1 .8 8 .8 5 3 2 3 1 5 3 .9 1 4 5 5 1 0 5 0 ,4 3 0 ,5 6 ,6 1 9 ,8 1 3 ,9 2 5 ,9 < 0 .1 0 ,0 -1 0 8 -1 3 ,8 4 C h u n 8 - 0 m n a 1 1 .4 8 .6 2 7 6 1 3 7 1 4 6 1 1 8 1 8 2 0 ,8 6 9 ,7 9 ,9 4 3 ,7 4 5 ,5 1 0 ,6 0 ,3 0 ,1 -7 4 ,7 -6 ,2 6 R iv e r- S tr e a m s C h u n 3 - 0 m 8 .3 6 .3 1 0 .2 3 2 7 2 1 2 9 .3 3 6 0 8 1 0 ,1 1 8 ,8 3 ,8 1 9 ,9 1 4 ,6 3 1 ,7 0 ,4 0 ,0 -1 0 1 ,4 -1 3 ,9 5 C h u n 4 - 0 m n a 8 .4 8 .4 8 2 6 9 1 2 7 .9 3 6 8 7 3 0 ,1 1 7 ,5 3 ,5 1 9 ,2 1 3 ,7 3 3 ,1 0 ,4 0 ,0 -1 0 4 -1 6 ,3 7 C h u n 6 - 0 m n a 1 3 .7 8 .6 5 4 8 .9 2 2 .8 1 3 2 5 0 ,1 5 ,1 3 ,2 2 ,6 1 ,8 2 1 ,5 < 0 .1 0 ,0 -1 1 5 ,4 -1 5 ,0 7 T a b le 2 .1 . Is ot op e an d o th er c h em ic al a n d p h ys ic al d at a an al ys ed f ro m w at er s am p le s co ll ec te d i n L ag o C h u n ga rá , n ei gh b ou ri n g sm al l la ke s an d m ai n s u rf ac e w at er in fl ow s. S am p le s w er e co ll ec te d in D ec em b er 2 0 0 8 . P os it io n o f th e co ll ec te d s am p le s is in d ic at ed in F ig u re 3 .1 . Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 2.4 Earlier work undertaken at Lago Chungará Given that this PhD thesis was prompted by earlier multiproxy studies, some data employed here have been published by other authors from the same research group (Herrera et al. 2006; Moreno et al. 2007; Sáez et al. 2007; Giralt et al. 2008). Earlier studies have characterised the surface and underground waters from the Chungará and Cotacotani lake district (Section 2.3) as well as the sedimentary infill of Lago Chungará. 2.4.1 Sedimentary record In November 2002 fifteen sediment cores (6.6 cm inner diameter and up to 8 m long) were recovered from Lago Chungará using a raft equipped with a Kullenberg coring system. All cores were cut in 1.5 m sections and physical properties (GRAPE-density, p-wave velocity and magnetic susceptibility) were measured in the laboratory using a GEOTEKTM Multi-Sensor Core Logger (MSCL) at 1 cm intervals. Thereafter, the cores were split into two halves, scanned using a DMT colour scanner, and the textures, colours and sedimentary structures were described. Smear slides were used to characterise the sediment composition in order to estimate the biogenic, clastic and endogenic mineral content (Moreno et al. 2007). The uppermost sedimentary infill of Lago Chungará was characterised by the lithological description of the fifteen cores (Sáez et al. 2007) and by seismic imagery obtained in 1993 (Valero- Garcés et al. 2000). A 3D sedimentary model was obtained after the construction of correlation sedimentary panels (Sáez et al. 2007) (Fig. 2.8). These stratigraphic correlation panels allowed us to select cores 10 and 11, located offshore (Fig. 2.7 and 2.8) to carry out the palaeoenvironmental reconstruction. A composite core recording the whole sedimentary infill (minimum thickness of 10 m) of the offshore zone was constructed from the detailed description and correlation of these two cores. From the bottom to the top of the composite core, two sedimentary units (units 1 and 2) were identified and correlated over the offshore zone of the lake mainly using 14 tephra keybeds (Fig. 2.8). Unit 1 is made up of diatomaceous ooze with variable types and quantities of carbonates (calcite, aragonite) and amorphous organic matter. This unit, which extends across the lake, is thicker in the NW sector of the central plain and thinner towards the south and west, probably overlapping the Miocene substrate. This unit lies in the central plain and on the steep flank of the lake (Fig. 2.8). It is divided into two subunits: subunit 1a and subunit 1b. Subunit 1a ranges in thickness between 0.58 m and 2.56 m and is composed of light-white and dark-green diatomaceous ooze couplets (Fig. 2.9). Subunit 1b (from 0.62 m to 1.87 m thick) is made up of centimetre-to decimetre-thick laminated brown diatomaceous ooze and endogenic carbonates that occur at low concentrations (Sáez et al. 2007) (Fig. 2.9). 40 41 Geological, geographical, limnological and climate framework M 1 1 M 8 M 7 M 1 1 M 1 1 M 8 M 9 M 9 M 7 M 6 M 6 M 6 M 5 M 5 M 5 M 2 A A A D D D M 2 M 2 M 2M 3 M 4 M 4 M 4 M 3 M 3 M 3 M 8 M 9 M 7 M 1 1 M 1 1 M 8 M 9 M 5 M -8 M -4 M -9 M -2 M -3 M -7 M -5 M -4 W -E A D A T U M f U n it 3 M -6 M 5 4 9 0 0 6 7 3 0 1 9 9 3 5 m d e p th 1 5 2 4 m d e p th 0 9 2 4 m d e p th 0 4 3 9 .9 m d e p th 1 2 3 5 m d e p th 1 3 2 3 ,5 m d e p th 0 3 2 7 ,5 m d e p th E P L A T F O R M 0 .4 5 -0 .8 5 k m E S L O P E 0 .2 k m W S L O P E 0 .1 1 k m R IS E 0 .3 5 k m C E N T R A L P L A IN 5 .2 k m 3 1 3 8 1 5 1 2 9 9 3 4 2 k m U n it 6 . T A L U S -S L O P E M A S S F L O W D E P O S IT S M IO C E N E V O L C A N IC R O C K S (S e is m ic U n it A ) P R E -C O L L A P S E A L L U V IA L -F L U V IA L D E P O S IT S (S e is m ic U n it B ), C o a rs e s a n d s C O L L A P S E B R E C C IA V O L C A N IC L A S T IC D E P O S IT S L a p ill i la y e r (F a c ie s M ) W h it e te p h ra la y e r (s ilt g ra in -s iz e d ) (F a c ie s N ) G re y to b la c k (s ilt g ra in -s iz e d ) (F a c ie s N ) U n it 1 . S H A L L O W T O D E E P O F F S H O R E F . A . U n it 2 . D E E P (W IT H A S H A L L O W E P IS O D E ) O F F S H O R E D E P . B ro w n -w h it e la m in a te d a n d c a rb o n a te -b e a ri n g (F a c ie s B ) d ia to m it e G re e n -w h it e la m in a te d (F a c ie s A ) d ia to m it e D a rk g re y -b la c k m a s s iv e d ia to m it e s (F a c ie s E ). B ro w n re d d is h m a s s iv e a n d b a n d e d d ia to m it e (F a c ie s D ), in te rb e d d in g c a rb o n a te la y e rs (F a c ie s F ) M a g n e ti c s u s c e p ti b ili ty p e a k k e y le v e l F in e s a n d a n d s ilt (F a c ie s K ) U n it 3 . S H A L L O W T O D E E P P L A T F O R M D E P . D a rk g re e n m a s s iv e d ia to m it e , ri c h in m a c ro p h y te a n d m o llu s c s re m a in s (F a c ie s I) D a rk g re e n , m a s s iv e o rg a n ic -r ic h d ia to m it e (F a c ie s C ) C a rb o n a te la y e r L o w c o n te n t in c a rb o n a te la y e r 1 m (l o g s s c a le ) S u b u n it 2 a U n it 4 U n it 5 U n it 6 S u b u n it 2 b S u b u n it 1 b S u b u n it 1 a F a u lt re la te d b re c c ia . C a rb o n a te c la s t, d ia to m it e m a tr ix (F a c ie s G ) nolateralcontinuity b iv a lv e re m a in s g a s tr o p o d e re m a in s o s tr a c o d e re m a in s a q u a ti c v e g e ta b le re m a in s o n la p U n it 4 . S H A L L O W T O V E R Y S H A L L O W L IT T O R A L D E P . D a rk g re y to lig h t g re y c a rb o n a te s it ly s a n d ri c h , in s p re m a in s (F a c ie s H ) C h a ra a n d in te rb e d d e d fi n e -s iz e d te p h ra la y e rs (F a c ie s N ) B la c k p e a ty le v e l, ri c h in n o n -c h a ro p h y te -d o m in a te d m a c ro p h y te a n d m o llu s c re m a in s (F a c ie s J ) B la c k p e a ty le v e l, ri c h in n o n -c h a ro p h y te -d o m in a te d m a c ro p h y te a n d m o llu s c re m a in s (F a c ie s J ) a n d in te rb e d d e d fi n e -s iz e d te p h ra la y e rs (F a c ie s N ) W A F W A F W A F W A F W A F ? ? 2 3 8 2 3 0 U / T h a g e d a ta (c a . y rs B P ) A M S C a g e d a ta (c a . c a l y rs B P ) 1 4 U n it 5 . D IS T A L A L L U V IA L -F A N D E P O S IT S G ra v e l to fi n e s a n d (F a c ie s L ) o n la p A F ig u r e 2 .8 . A ) W es t– E as t st ra ti gr ap h ic c ro ss -s ec ti on ; B ) N or th – S ou th s tr at ig ra p h ic c ro ss -s ec ti on . S tr at ig ra p h ic c or re la ti on s ar e b as ed o n li th os tr at ig ra p h ic a n d s ed im en to lo gi ca l c ri te ri a (l im it s b et w ee n u n it s an d s om e ke y le ve ls a n d fa ci es ) an d m ag n et ic s u sc ep ti b il it y p ro fi le s. T o im p ro ve c la ri ty , t h e h or iz on ta l s ca le s ar e n ot t h e sa m e in t h e ce n tr al t ro u gh a s in t h e p la tf or m s (S áe z et a l. , 2 0 0 7) . Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles M 2 M 3 M 4 M 4 D A T U M M -8 A A A A A B B B B B C C C C C D D D D D M 5 M 6 M -1 M 1 1 M 1 0 M 1 2 M -8 M 4 M 5 M 4 M 1 2 M 1 3 M 1 3 M 8 M 9 M 9 M 9 M 9 M 9 M 1 1 M 1 1 M 1 2 M 1 3 M 1 4 M 1 4 M 8 M 7 M 7 M 5 M 6 M 6 M 6 M 6 N -S M 11 M 11 M 1 0 M 1 0 M 1 0 M 1 0 M 8 M 7 M 7 M 5 M 5 M 2 M 3 M 2 M 3 M 3 1 m (l o g s s c a le ) M -1 M 8 M 1 1 M 1 0 M 9 1 1 2 8 m d e p th 0 2 3 4 m d e p th 1 2 3 5 m d e p th 0 1 3 7 m d e p th 1 4 2 7 m d e p th 1 0 3 6 m d e p th 0 5 2 6 m d e p th 0 7 1 8 m d e p th 2 6 2 0 2 3 2 0 7 2 0 0 -9 5 5 0 8 3 1 0 -1 1 2 9 0 1 2 7 9 0 -1 5 5 1 0 5 7 0 0 S u b u n it 2 a S u b u n it 2 b S u b u n it 1 b S u b u n it 1 a 9 9 4 0 -1 3 0 7 0 1 1 2 4 0 -1 3 9 7 0 9 6 5 0 -1 2 9 4 0 C E N T R A L P L A IN 5 .5 k m 5 1 11 2 7 1 4 1 1 0 2 2 k m o n la p o n la p M 3 W A F W A F R IS E 0 .3 5 k m U n it 6 B F ig u r e 2 .8 . (c on ti n u ed ). 42 43 Geological, geographical, limnological and climate framework removed BASE TOP S U B U N IT 1 b S U B U N IT 1 b S U B U N IT 1 a S U B U N IT 1 a S U B U N IT 2 a S U B U N IT 2 a S U B U N IT 2 b 10 cm onlap sec. 2sec. 3sec. 4 sec. 5 sec. 6 3,465 2,747 1,950 6,391 9,612 11,186 11,502 11,942 12,415 10,379 10,941 7,289 8,151 8,802 2,075 L a te G la c ia l H o lo c e n e 1 2 3 Figure 2.9. DMT scanner (LRC, Minnesota) image of core 11. Lithological units, 14C AMS radiocarbon dates, intervals sampled for chapters 4, 5 and 6 (yellow squares), and the position of the samples analysed for chapter 7 (red arrows) are indicated. Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 44 Unit 2 is about 6 metres-thick and grades laterally to the west and south into alluvial and deltaic deposits, and towards the east into macrophyte, organic-rich facies (Fig. 2.8). It is mainly made up of massive to slightly banded diatomaceous ooze interbedded with 13 tephra layers. Unit 2 is also divided into two subunits: subunit 2a and subunit 2b. Subunit 2a (between 1.56 m and 3.44 m thick) is composed of massive brownish-red to slightly banded sapropelic diatomaceous ooze with abundant calcite crystals (silt grain-sized) and carbonate-rich layers (Fig. 2.9). The sediments of the uppermost subunit 2b range from 0.86 m to 3 m in thickness and consist of dark-grey diatomaceous ooze with frequent macrophyte remains alternating with massive black tephra layers, mainly constituted by plagioclase, glass and mafic minerals (Fig. 2.9) (for further details see Moreno et al. 2007 and Sáez et al. 2007). The Lago Chungará sediments were analysed by X-Ray Fluorescence (XRF), X-Ray Diffraction (XRD), Total Carbon and Total Organic Carbon (TC and TOC), Biogenic Silica (BSi), pollen and diatoms (Fig. 2.10). The XRD analyses indicated that offshore sediments of Lago Chungará are constituted by an amorphous and a crystalline fraction. The amorphous fraction, ranging from 40 wt% (towards the top of the composite core) to almost 100 wt% (in the lower two thirds of the composite core), is made up of organic matter and diatoms. The crystalline fraction, ranging between 1 wt% and 60 wt%, consists of Ca-plagioclase, carbonate (calcite and dolomite), muscovite, pyrite, quartz and an amphibole (probably riebeckite) (Fig. 2.10). In addition, XRF analysis of the distribution of chemical elements enabled us to detect three main components in the Lago Chungará sediments: a) lacustrine biological remains (mainly diatoms), b) volcanic minerals and c) endogenic offshore carbonates. The first component corresponds to the amorphous fraction revealed by the XRD and the other two correspond to the crystalline fraction (Fig. 2.10) (Moreno et al. 2007). The geology of the Lago Chungará catchment (see Sáez et al. 2007 for further details) indicates that volcanic minerals have two main provenances: a) the erosion of former volcanic rocks and that of previous deposited fallout material from the catchment, and b) the ash fallout from the Parinacota volcano. Conversely, the origin of the carbonates is not so straightforward. The presence of large amounts of Ca of volcanic origin dissolved in the lake water together with a marked lake level decrease could have promoted the precipitation of these carbonates (see Giralt et al. 2008 for further details). 2.4.2 Chronological Framework The chronological model for the sedimentary sequence of Lago Chungará is based on 17 AMS 14C dates of bulk organic matter and aquatic plant macrofossils, and one 238U/230Th date from carbonates (Table 2.2 and Fig. 2.11). The radiocarbon dates were performed in the Poznan Radiocarbon Laboratory (Poland) whereas the 238U/230Th sample was analysed by high-resolution ICP-IRMS multicollector at the University of Minnesota (Edwards et al. 1987; Cheng et al. 2000; Shen et al. 2002). The main problems encountered in the construction 45 Geological, geographical, limnological and climate framework Mineralogy (%) Amorphous material Plagioclase Calcite Quartz Muscovite Chlorite Dolomite Pyrite Amphibole Magnetic Susceptibility (10 S.I) -5 Biogenic Opal (%) Total Carbon (%) Total Organic Carbon (%) Grey curve C o re D e p th (m m ) L it h o lo g ic a l u n it s S u b u n it 1 a S u b u n it 1 b S u b u n it 2 a S u b u n it 2 b M in e ra lo g ic a l z o n e s 0 40 1000 600 30 0 3 0 2 0 0.2 0 0.6 0 1 0 1.5 1 2 3 0 80 0 750 120 12 100200 1,000 2,000 3,000 4,000 5,000 6,000 7,000 8,000 C o re D e p th (m m ) 0 1,000 2,000 3,000 4,000 5,000 6,000 7,000 8,000 Light elements (cps) Heavy elements (cps) Al 5001000 4000800010002000 5000100003500070000 30006000 200040002000040000 50 150 1000 2000 300600 600 900 2000 4000 Si S K Ca Ti Mn Fe Rb Sr Zr Sn BaL it h o lo g ic a l u n it s X R F z o n e s U p p e r M id d le L o w e r S u b u n it 1 a S u b u n it 1 b S u b u n it 2 a S u b u n it 2 b A B Figure 2.10. A) Mineralogy (expressed as percentages on total dry weight), magnetic susceptibility (expressed as standard units), Total Biogenic Silica (BSi), Total Carbon (TC), Total Organic Carbon (TOC) (expressed as percentages) and grey-colour curve of the Lago Chungará sediments (Giralt et al. 2008). B) Analysed light and heavy elements (expressed as counts per second) in the sediments of Lago Chungará. The zones correspond to those established in Moreno et al. (2007). of reliable chronological frameworks for the lacustrine sedimentary infill of most lakes in the Andean Altiplano were: a) the determination of the large and variable radiocarbon reservoir effect (Geyh et al. 1999; Geyh and Grosjean, 2000; Grosjean et al. 2001), and b) the temporal evolution of this radiocarbon reservoir effect. Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles Time (calendar years BP) 0 0 1,000 2,000 3,000 4,000 5,000 6,000 7,000 8,000 9,000 5,000 10,000 15,000 Subunit 2b Subunit 2a Subunit 1b Subunit 1a C o re d e p th (m m ) Figure 2.11. Chronological framework of Lago Chungará was constructed using 15 14C AMS dates and after applying Heegaard’s method (Heegaard et al. 2005) in calendar years BP (modified from Moreno et al. 2007). The dotted line represents the chronological model constructed considering a constant reservoir-effect of 3,260 years. The dashed line depicts the chronological model considering a reservoir effect of 3,260 years for the upper sedimentary unit and 0 years for the lower unit. The intermediate continuous line represents the average model of the two earlier ones and is applied to Lago Chungará. Horizontal bars are the confidence error of the radiometric dates after applying the method of Heegaard. Reproduced from Giralt et al. (2008). 46 The modern reservoir effect in lakes must be calculated as the difference between the 14C age of the water and that of the atmospheric 14CO 2 measured at the same time (Giralt et al. 2008). Radiocarbon dating of the modern DIC resulted in 2,320 ± 40 14C years BP (Table 2.2), but this value must be corrected owing to the atmospheric thermonuclear bomb tests carried out during the late 1950s–early 1960s. These nuclear tests doubled the amount of 14C in the atmosphere (Reimer et al. 2004b). The effects of these thermonuclear tests on the modern carbon cycle of a lake depend on a) the year in which the sample was taken and b) the residence time of the lake (Hua and Barbetti, 2004; Goslar et al. 2005). The Lago Chungara water residence time is about 15 years (Herrera et al. 2006) and the DIC sample was obtained and dated in 2004. Therefore, the real present-day reservoir effect of this lake is 3,260 years BP: 2,320 years BP (radiocarbon date of the DIC) minus –920 years (apparent radiocarbon age of the atmospheric 14CO 2 for the period 1988–2002) (Hua and Barbetti 2004; Goslar et al. 2005) (see Giralt et al. 2008 for further details). The reservoir effect in the Altiplano lakes proved to be highly variable over time. One of the factors that could influence the reservoir effect is the change in the volume/surface ratio of the lake, which is a function of the water depth (Geyh et al. 1998; Grosjean et al. 2001). According to these authors, the 47 Geological, geographical, limnological and climate framework reservoir effect decreases when the lake level diminishes, and viceversa. Therefore, the approach followed in Moreno et al. (2007) and Giralt et al. (2008) to correct the dates for the variable reservoir effect was based on two assumptions: a) the Lago Chungará ecosystem had an environmental status during the deposition of Unit 2b akin to the one existing at present and b) the present-day lake level is at its highest position. Accordingly, a different correction of the reservoir effect was applied to Units 1 and 2. A constant reservoir effect of 3,260 years was substracted from the radiocarbon dates present in Unit 2. However, it is only possible to speculate about the variations over time of the reservoir effect in Unit 1. Given that it was not possible to estimate the reservoir effect by applying the methodology of Geyh et al. (1998), the ages of the two extreme reservoir values (a minimum of 0 and a maximum of 3,260 years) were calculated (Table 2.2 and Fig. 2.11). A theoretical mid point between the two extreme reservoir effect points was calculated for Unit 1 and this theoretical value was employed to construct the final chronological model (Fig. 2.11). All the corrected radiocarbon dates were calibrated using the CALIB 5.02 software package (Reimer et al. 2004a) selecting the mid-point of 95.4% of the distribution. The upper and lower limits of the chronological model were calculated employing the Cagedepth software (Heegaard et al. 2005). Lithological Units Composite depth (mm) Lab ID Sample reference Sample material Uncalibrated 14 C (years BP) Calibrated age a (calendar years BP) Calibrated age b (calendar years BP)  13 C (%PDB) Unit 2 - Beta-188745 - Lake water 2,320 ± 40 - - - 375 Poz-8726 14 A-1, 5 Bulk organic remains 4,620 ± 40 1,850 ± 454 2,050 ± 500 -13.6 ± 0.2 420 Poz-8720 11 A-2, 39 Bulk organic remains 4,850 ± 40 1,964 ± 460 2,185 ± 520 -12.9 ± 0.4 670 AA56904 15 A-2, 48 Aquatic plants 6,635 ± 39 2,590 ± 630 2,900 ± 710 -25.46 951 Poz-8721 11 A-2, 84 Bulk organic remains 7,290 ± 50 3,290 ± 860 3,645 ± 910 -14.8 ± 0.2 2,573 Poo-8723 11 A-3, 2 Bulk organic remains 8,920 ± 50 6,227 ± 1,200 6,555 ± 1,010 -16.1 ± 0.1 3,440 AA56903 15 A-4, 27 Aquatic plants 9,999 ± 50 7,070 ± 1,200 7,505 ± 890 - 4,361 Poz-8724 11 A-3,86 Bulk organic remains 10,860 ± 60 7,530 ± 1,250 8,775 ± 940 -16.9 ± 0.1 Unit 1 4,909.1 Poz-7170 11 A-3, 123 Bulk organic remains 8,570 ± 50 7,705 ± 1,230 9,900 ± 950 -16.8 ± 0.1 5,504.8 Poz-8647 11 A-4, 10 Bulk organic remains 9,860 ± 60 7,940 ± 1,260 11,290 ± 1,080 -14.1 ± 0.3 6,152 Poz-7171 11 A-4, 63 Bulk organic remains 11,070 ± 70 8,270 ± 1,400 12,490 ± 910 -13.6 ± 0.2 6,650 AA56905 15 A-5, 77 Aquatic plants 4,385 ± 101 - - - 6,750 Poz-8725 13 A-4, 66 Bulk organic remains 8,810 ± 50 - - -22.9 ± 0.1 6,962 Poo-11891 11 A-4, 145.5 Bulk organic remains 11,460 ± 60 8,765 ± 1,420 13,120 ± 930 -16.2 ± 0.4 7,442 Poz-13032 11 A-5, 41 Bulk organic remains 10,950 ± 80 9,080 ± 1,540 13,290 ± 910 -22.7 ± 2.3 7,852 Paz-11982 11 A-5, 84 Bulk organic remains 11,180 ± 70 9,400 ± 1,740 13,605 ± 880 -28.7 ± 3.7 8,272 Poz-13033 11 A-6, 41 Bulk organic remains 12,120 ± 80 9,730 ± 2,090 14,155 ± 1,390 -19.6 ± 1.7 8,652 Poz-7169 11 A-6, 79 Bulk organic remains 13,100 ± 80 10,040 ± 2,640 14,795 ± 1,760 -23.1 ± 0.2 Composite depth (mm) Sample reference Carbonate Type 238 U (ppb) 232 Th (ppm)  234 U measured 230 Th/ 238 U 230 Th/ 232 Th Calendar age (years BP) 3,440 13 A-2, 105 Crystalline 467,4 35.2 413.5 0.1036 23 -6,728 ± 974 Table 2.2. A) AMS 14C datings carried out to construct the chronological framework of the Lago Chungará sedimentary sequence. B) Results of 238U/230Th datings. Reproduced from Giralt et al. (2008). Bolded radiocarbon dates were not taken into account in the construccition of the chronological framework 49 Chapter 3 Methods 3.1 Hydrochemical and isotopic water analyses Water samples were collected from Lago Chungará in December 2009 in order to complement samplings carried out by the research team in 2002 and 2004 (Herrera et al. 2006). Water samples were taken each 2 metres in vertical profiles up to water depths of 8 and 20 metres in the lake. These water samples were complemented by additional samplings from the nearby smaller lakes, rivers and streams (Table 2.1 and Fig. 3.1). The variables measured in situ were conductivity, oxygen concentration, pH, temperature and salinity. A total of 24 samples were collected for isotope analyses. The samples for the δ18O and δD analyses were stored in 50 ml polythene tubes. For the δ18O analysis, the waters were equilibrated with CO 2 prior to mass spectrometry measurements, whereas for δD analysis the samples were reduced to H 2 using Pt before mass spectrometry. The isotope analyses were conducted using a Finnigan MAT Delta S IRMS. Isotopic standards employed as reference were VSMOW, SLAP and GISP for δD and δ18O in water samples. Replicate analysis of all the samples indicated a precision of ±<0.1‰ (1 σ) (for δ18O) and 1.5‰ (1σ) (for δD). All these samples were also analysed by means of Ion-exchange chromatography, volumetric analysis and Inductively Coupled Plasma (ICP-OES) in order to measure the concentrations of the main chemical components (Cl-, SO 4 2-, NO 3 -, HCO 3 - and cations). The water samples were previously filtered (0.45 μm) during the field sampling. All the water samples were analysed at the Serveis Científico-Tècnics de la Universitat de Barcelona. 3.2 Sediment sampling The Lago Chungará sediments were sampled following different strategies in accordance with three main objectives: a) To characterise the climate and environmental changes, at different resolutions, taking place in the Late Glacial-Early Holocene transition (12,000-9,400 cal years BP). Three short intervals consisting Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles Figure 3.1. A) Aerial-view of the Lago Chungará region with the position of water sampling sites. B) and C) pictures of water sampling in Lago and Río Chungará, respectively. of light and dark finely-laminated sediments were selected. Light laminae were formed by the accumulation of massive short-term diatom blooms, probably lasting only days or weeks, whereas dark laminae would represent a normal annual cycle of the lake with alternating phases of stratification and mixing over several years. Interval 1 (11,990–11,450 cal years BP) was located at subunit 1a. Interval 2 (10,430–10,260 cal years BP) was located at the transition between subunit 1a and subunit 1b. Interval 3 (9,890–9,430 cal years BP) was located at subunit 1b. Individual laminae within the three intervals were sampled with a blade for isotope analyses. A total of 190 samples (111 samples from interval 1, 37 samples from interval 2, and 42 samples from interval 3) were obtained. A selection of 37 samples from dark laminae of these 3 intervals was selected to investigate the baseline hydrological evolution of Lago Chungará by means of δ18O diatom analysis. Additionally, 40 samples of dark laminae from a stretch of 46.5 cm corresponding to the interval 1 (11,990-11,450 cal years BP) were analysed by an ultra-high resolution δ18O diatom reconstruction to track the influence of the ENSO and solar activity at this time interval. b) To characterise the biogeochemical changes taking place at ultra-high frequency. To this end, 102 samples of both light and dark laminae from the interval 1 were selected for δ18O diatom . Additionally, 11 samples from interval 1 were also analysed for δ13C diatom and %C diatom . c) To characterise the high-frequency environmental changes recorded in laminated unit 1 (12,400- 8,400 cal years BP). A total of 51 bulk samples were selected for δ18O diatom and δ13C diatom analyses (Fig. 2.9). The sampling was carried out every 10 cm except where the sediments were either carbonate or tephra-rich. 50 Lago Chungará Río Chungará 5 Km Bofedal de Parinacota Lagunas de Cotacotani 5 A B C 4 3 10 5 6 91 2 7 8 51 3.3 Light and Scanning Electron Microscope sample preparation Each interval from unit 1 (Intervals 1, 2, 3) was continuously covered by thin sections. Thin sections of 120 mm x 35 mm (30 μm in thickness), with an overlap of 1 cm at each end, were obtained after freeze-drying and balsam-hardening (Fig. 3.2). Detailed petrographical descriptions and lamina thickness measurements were performed with a Zeiss Axioplan 2 Imaging petrographic microscope. A number of representative samples were also selected for observation with the Jeol JSM-840 and the Hitachi H-4100FE field emission Scanning Electron Microscopes (SEM) to complement the petrographical study using the thin sections and to check the purity of the samples for isotopic analyses. The samples were dried in two steps because of the high absorbed water content: a) most of the water was eliminated by capillary action using filter paper, and b) samples were freeze-dried and vacuum stored prior to being carbon coated. Pushing Knife Nylon wire Core 1.5 cm Sediment surface Met alic sam pler 0 cm 10 cm 20 cm Core 2 A D G E F B C Figure 3.2. Sampling procedure to obtain the material of the thin sections in soft sediment cores. A) Insertion of a metallic sampler into one side of the core. B) If more than one metallic sampler is used for the same core, a minimum overlapping of 1.5 cm is performed. C) Annotation of the bottom depth, the top depth, and some middle depths, indicating the polarity of the sample as well as the name of the core. D) Nylon wire cutting the sediment with the aid of a knife at the top end. E) To avoid horizontal displacement, the bottom of the metallic sampler is secured with the knife. Nylon wire is then pulled along to the end of the metallic sampler. F) Inclination of the core 45 degrees and extraction of the metallic sampler. G) Samples prepared for freezing for subsequent study of thin sections. All thin sections were performed at GFZ in Potsdam. 3.4 Isotope analyses in sediments 3.4.1 Cleaning of diatom frustules Analysis of δ18O diatom requires the material to be almost pure diatomite since fluorination techniques following Clayton and Mayeda (1963) will release oxygen from all the components in the sediment; i.e. silt, clay, tephra, carbonates and organic matter (Juillet-Leclerc, 1986). Hence, samples were treated in Methods Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles accordance with the four clean-up stage method proposed by Morley et al. (2004) with some variations (Fig. 3.3). Sample purity was checked after each stage using light microscopy: Stage 1: Organic and carbonate removal. Standard methods were used to remove organic and carbonate material (Battarbee et al. 2001). A few grams of sediment were heated at 90ºC in 30% H 2 O 2 until all reaction ceased. Carbonate was then removed using 10% HCl for 12h. A stronger oxidising agent (concentrated HNO 3 heated at approximately 80°C for 1h) was used to eliminate any remaining organic matter. The samples were washed in a centrifuge three times in distilled water between each step and before the next stage. Stage 2: Sieving. The samples were first sieved using a 125 μm sieve. This eliminated resistant charcoal and terrigenous particles >125 μm. The samples were then sieved again to obtain a fraction between 38 and 63 μm to remove clay and the remaining detrital grains. A diatom concentrate made up almost exclusively of valves of the large centric diatom Cyclostephanos andinus was obtained, thereby eliminating any species-specific effect variability in the samples. At this stage all diatoms of <38 μm were removed, an insignificant number of diatoms remained in the >63 μm fraction. Stage 3: Differential settling. Gravity settling in a water column during the sieving process also helped to remove any remaining tephra and clay particles. The 38–63 μm sieved fraction was placed in glass beakers. Differential settling occurred since faster settling coarser silt grains were deposited under the slower settling diatoms. The diatom layer was carefully removed using a pipette. Excess water was decanted away but samples were kept wet. Stage 4: Gravitational split-flow lateral-transport thin (SPLITT). The fourth stage was an alternative approach to heavy liquid separation. SPLITT was developed by J.C. Giddings at the University of Utah (Giddings, 1985) and first applied to the separation of diatoms at the University of Jülich (Schleser et al. 2001; Rings et al. 2004) (Fig. 3.4). SPLITT utilises the different densities and hydrodynamic properties of diatoms and contaminants to produce two distinct end fractions. Samples are passed along a narrow channel in a laminar flow at a constant velocity. Sediments of different density and hydrodynamic properties settle differently to create two distinct end fractions (Leng and Barker, 2006). A sample (A’) is introduced into a thin channel where it meets a carrier fluid (usually water for diatom samples) in a laminar flow (Fig. 3.4). The velocity of the flow is controlled to separate the sample into two streams: the upper stream and the lower stream. The former stream, which contains finer, less dense and more hydrodynamic particles, passes through an outlet (A). The latter stream enables the collection of the remaining particles to pass through another outlet (B) (Leng and Barker, 2006) (Fig. 3.4). SPLITT was employed at Lancaster University (UK). The SPLITT technique was only applied to the 38-63 μm fraction which contained clays and fine tephra particles after the previous settling stage. 52 Finally, the purified diatom samples were dried at 40ºC between 24h and 48h. After the cleaning process all the samples were checked under a light microscope and a number of random samples were verified with XRD, TC analysis and SEM observations. This verification process confirmed that the samples did not contain significant amounts of terrigenous matter. TC values were below 0.5 wt% and the terrigenous content (clays or tephra) was less than 1 wt% (Fig. 3.3). The final isotope data were not affected despite the fact that a large number of diatoms were broken during the cleaning process. Non-purified sample Purified sample STAGE 1 30% H O2 2 wash 3x distilled water wash 3x distilled water wash 3x distilled water 5% HCl HNO concent.3 Organic and carbonate removal STAGE 2 STAGE 3 63 m sieve 125 m sieve >125 m charcoal  >63 m - <125 m clasts + large diatoms   >38 m - <63 m mostly diatoms + some clasts   <38 m small diatoms + clay  38 m sieve clay fraction diatoms residue Sieving Differential settling STAGE 4 Gravitational split-flow lateral-transport thin (SPLITT) Peristaltic pump A Sediment input (A’) Water input (B’) Bubble trap OSP ISP Fine/less dense output (A) Coarse/dense paricle output (B)Peristaltic pump B Flow direction 10 o Less dense fraction (A) dry at 40ºc Purified sample A B Figure 3.3. A) Diagram showing the four-stage laboratory pretreatment before δ18O diatom analysis. Stage 1 chemically removes organic matter and carbonates. Stages 2, 3 and 4 eliminate mineral contaminants that may contribute to the δ18O diatom signal. The duration of the treatment varies from sample to sample (modified from Morley et al. 2004). B) SEM images of two samples before cleaning (left) and after cleaning (right) 53 Methods Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 3.4.2 Oxygen isotope extraction Diatom frustules consist of an inner tetrahedrally bonded silica skeleton (Si–O–Si) with an outer hydrous layer (Labeyrie and Juillet,1982) (Fig. 1.10). The distribution of these two components is more complex than simple layering and is related to the mode of formation of the frustule and its differential porosity. Through the dissolution of the hydrous parts of the frustule, Juillet (1980) demonstrated that the internal dense silica is isotopically homogenous whereas the outer hydrous layer freely exchanges with any water that the diatom silica comes into contact with. Knauth (1973) considered that biogenic opal has 7–12 wt% water, compared to 1 wt% for quartz, although Leng et al. (2001) suggested that 20– 30% of diatom oxygen needs to be removed before stable values for δ18O are reached. Thus, prior to analysis it is essential to remove the –Si–OH layer of the diatom frustule when attempting to obtain environmental records from δ18Odiatom (Swann and Leng, 2009). The extraction of the –Si–OH layer, however, which may require the removal of 7–40% of all oxygen within diatoms (Leng and Sloane, 2008), is technically challenging and requires specialised equipment, hazardous reagents and highly trained operators (Swann and Leng, 2009). The internal Si–O bond needs considerable energy to break and requires the use of an extremely powerful oxidising reagent (i.e., a fluorine based compound such as ClF 3 or BrF 5 ) or high temperatures (Leng and Barker, 2006; Leng and Sloane, 2008). The classic fluorination approach using the step wise approach method described here (Fig.3.5) is the standard methodology employed for extracting oxygen from diatom silica prior to mass spectrometry at the NERC Isotope Geosciences Laboratory (NIGL), UK. All isotope analyses described in this PhD thesis (with the exception of the water O and H) were analysed at NIGL. Peristaltic pump A Sediment input (A’) Water input (B’) Bubble trap OSP ISP Fine/less dense output (A) Coarse/dense paricle output (B)Peristaltic pump B Flow direction 10 o A B Figure 3.4. A) The SPLITT system utilises the different properties of sediment for separation into two fractions. Sediments with a higher density will settle faster than sediments with a lower density. Similarly, oval-shaped sediments will have greater drag than needle-shaped sediments. These properties create two distinct end fractions. The process can be repeated several times to further purify samples. B) Image of the SPLITT from the Lancaster Environment Centre at Lancaster University. 54 Stepwise Fluorination (SWF) involves a three-stage process (Haimson and Knauth, 1983; Matheney and Knauth, 1989). First, samples between 5 and 10 mg were outgassed in nickel reaction tubes at room temperature to remove the hydrous layer and excess water. Given that the first fluorination stage of the SWF methodology removed some non-diatom contamination (e.g. highly reactive clays) in the sample (Matheney and Knauth, 1989), the SWF method provides δ18O diatom data that is less distorted by contamination when sample purity is less than 100% (Swann and Leng, 2009). The second fluorination stage was used to release oxygen from the outer –Si–O–Si layer. This involved prefluorination using a stoichiometric deficiency of the reagent BrF 5 at a low temperature. The third stage involved a full reaction at 450ºC for 12 hours with an excess of BrF 5 to release the oxygen (gas) from the structurally bound component (Leng and Barker, 2006; Leng and Sloane, 2008). The oxygen liberated was then converted Purified BrF5 Gauge O yield2 Yield Collection vessel Graphite rod KBr trapLN trap2 Nickel reaction vessel Sample P P P LN trap2 A B Figure 3.5. The fluorination line used at the NERC Isotope Geosciences Laboratory for the extraction of oxygen (as CO 2 ) from biogenic silica for IRMS. A) Fish-eye lens image; B) schematic. Modified from Leng and Sloane (2008). 55 Methods Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles into CO 2 by exposure to graphite using the method of Clayton and Mayeda (1963). The CO 2 was measured with a Finnigan MAT 253 dual inlet mass spectrometer and normalised against a NBS-28 quartz international standard. During oxygen extraction, oxygen yields were monitored for every sample and compared with their calculated theoretical yield for SiO 2 . Most samples had mean yields of 69% - 70% of their theoretical yield based on silica. This suggests that around 30% of the material including hydroxyl and loosely bonded water (both OH– and H 2 O) was removed during prefluorination. A random selection of more than 50 samples were analysed in duplicate or in triplicate, giving a reproducibility between 0.01‰ and 0.6‰ (1σ) with a mean value of 0.15‰. The standard laboratory quartz and a diatomite control sample (BFC) had a mean reproducibility over the period of analysis of 0.2‰. The oxygen isotope composition of diatom silica is expressed on the delta-scale in terms of per mil (Equation 1). For diatom oxygen the reference is VSMOW (Vienna Standard Mean Ocean Water, which is specially-prepared distilled seawater). 3.4.3 Analyses of δ13C diatom and %C diatom 13C/12C ratios and %C of organic matter within the diatom frustules were analysed by combustion in an elemental analyser (Costech ECS4010) interfaced with a VG dual inlet isotope ratio mass spectrometer (Fig. 3.6). Samples containing 1 to 2 mg of pure diatomite were loaded into tin capsules and placed on the carousel of the elemental analyser. The samples were sequentially dropped into a continuous flow of helium carrier gas into a 1020ºC furnace. A pulse of oxygen gas promoted an exothermal flash oxidation of the tin, ensuring full combustion of the sample, and the product gases were further oxidised by chromium and cobaltous oxides in the lower part of the furnace. The excess oxygen and water were removed by passing them through hot copper and magnesium perchlorate. The remaining N 2 and CO 2 were then passed through a Gas Chromatrography column and a thermal conductivity detector. This generated an electrical signal proportional to the concentrations of N 2 and CO 2 present in the helium stream. The Costech software station acquired and evaluated this information, producing %C data for the sample. As a result, it was possible to calibrate the instrument and to quantify the content of carbon of the unknown sample by analysing a standard sample of a given composition under the same operating conditions. Meanwhile the helium stream had carried the CO 2 through a trap at –90ºC (for complete removal of water), before reaching the Triple Trap held at –196ºC. Here the CO 2 was frozen, allowing the N 2 and helium to vent to the atmosphere. The TripleTrap was then evacuated before warming the CO 2 trap and discharging the sample CO 2 into the inlet of the Optima (Fig. 3.6). 56 The Optima mass spectrometer had triple collectors allowing simultaneous monitoring of CO 2 ion beams at the mass-to-charge ratio (m/e) = 44, 45 and 46 and a dual-inlet allowing rapid comparison of sample CO 2 compared with a reference CO 2 . 45/44 molecular mass ratios were converted to 13C/12C ratios after correction for common ion effects (Craig correction). Samples were measured against a within-run laboratory standard (BROC1). Based on the knowledge of the δ13C values obtained from the lab standard (derived from regular comparison with international calibration and reference materials NBS-19 and NBS-22), the 13C/12C ratios of the unknown samples were converted into δ13C values versus VPDB (Leng, pers comm.). Replicate analyses of well-mixed samples indicated a precision of ±<0.1‰ (1 σ) (for δ13C) and 0.1 (1 σ) (for %C). 3.5 Statistical analyses and Grey-colour curve The Multi-Taper Method (MTM) and the Time-Frequency (TF) analysis were employed to examine the periodic components in the δ18O values from interval 1. These spectral analyses enabled us to examine statistically significant modes in the time series in both the frequency and time domains. MTM provided a means of spectral estimation (Thompson, 1982) and a signal reconstruction (e.g. Park, 1992) for time series with spectra that contain both singular and continuous components (Theissen et al. 2008). MTM has been widely employed in the analysis of geophysical data including data from palaeoclimate studies (e.g. Mann et al. 1995; Mann and Park, 1996). The TF analysis is a hybrid tool constituted by the Fourier Transform and wavelets to examine non-stationary phenomena that decompose a time series into time- frequency space. As a result, the dominant modes of variability and the manner in which these modes vary in time can be determined (Lau and Weng, 1995; Torrence and Compo, 1998). For this reason, this analysis does not use a fixed-size Gaussian window but a Gaussian window that adapts to the spectrum Figure 3.6. Images of the carbon isotope analyser used at the NERC Isotope Geosciences Laboratory for the extraction of carbon (as CO 2 ) from biogenic silica and for analysis by means of an Optima mass spectrometer. 57 Methods Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles (Stockwell et al. 1996). All the statistical treatments of the datasets were performed using the R software package (R Development Core Team, 2008). Grey-colour curve was calculated using the ImageJ software package (Rasband, 1997–2009). A surface from the interval studied was selected in order to obtain a digital image which was a two- dimensional array of pixels. Every pixel presented a value related to the light reflection from a grey- scale variation. The image was an 8-bit grey-scale, and each pixel represented a value ranging from 0 (black) to 255 (white). Thus, it was possible to calculate a curve showing the variations along every pixel in a row. The results are presented in a 21 running mean curve. 58 Chapter 4 The palaeohydrological evolution of Lago Chungará (Andean Altiplano, northern Chile) during the Late Glacial and Early Holocene using oxygen isotopes in diatom silica* 4.1 Introduction Oxygen isotopes of diatom silica have been widely used in palaeoenvironmental reconstructions from lake sediments in the last decade (see Leng and Barker, 2006 for a comprehensive review). Using δ18O in palaeoenvironmental reconstruction is however not easy, because the sedimentary record can be influenced by a wide range of interlinked environmental processes ranging from regional climate change to local hydrology. The oxygen isotopic composition of diatom silica depends on the isotope composition of the water when the skeleton of the siliceous micro-organisms is secreted, and also on the ambient water temperature (Shemesh et al. 1992). Therefore, knowledge of all the environmental factors that may have influenced the isotope composition of the lake water is vital for the interpretation of the δ18O diatom signal (Leng et al. 2005b). One of these environmental factors is evaporation, which has a major influence on the isotope composition of any standing water body (Leng and Marshall, 2004). The δ18O record can therefore be used, at least in closed lakes, as an indicator of changes in the P/E related to climate changes (Leng and Marshall, 2004). Yet, before any palaeoclimatic interpretation of the isotope records from a lake is considered, other local palaeohydrological intervening factors from the basin need to be taken into account (Sáez and Cabrera, 2002; Leng et al. 2005b). The sedimentary records of high-altitude, Andean Altiplano lakes, are good candidates for carrying out oxygen isotope studies to reconstruct the Late Quaternary palaeoclimatology of the region, because they preserve an excellent centennial- to millennial-scale record of effective moisture fluctuations and source changes during the Late Glacial and Holocene although the interpretation not always strength forward (Abbot et al. 1997; Argollo and Mourguiart, 2000; Valero-Garcés et al. 2000, 2003; Baker et al. *Chapter based on the paper published in: Journal of Quaternary Science (2008) vol. 23(4) 351-363. Armand Hernández, Roberto Bao, Santiago Giralt, Melanie J. Leng, Philip A. Barker, Alberto Sáez, Juan J. Pueyo, Ana Moreno, Blas L. Valero-Garcés and Hilary J. Sloane. 59 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 2001a,b;Grosjean et al. 2001; Tapia et al. 2003; Fritz et al. 2004, 2006; Placzek et al. 2006). The δ18O analyses of carbonates, cellulose and biogenic silica have successfully been used to reconstruct the hydrological responses to climate change in different Andean lacustrine systems (Schwalb et al. 1999; Abbott et al. 2000, 2003; Seltzer et al. 2000; Wolfe et al. 2001; Polissar et al. 2006). Up to now, only stable isotopes in carbonates have been examined in Lago Chungará (Valero- Garcés et al. 2003), although its sedimentary record is made up of rich diatomaceous ooze ideal for diatom silica oxygen isotope studies. Lago Chungará currently behaves as a closed lake, without any surface outlet and evaporation as the dominant water loss process (Herrera et al. 2006); however it has shown a complex depositional history since the Late Glacial (Sáez et al. 2007) and the relative role of other factors (groundwater versus evaporation) should be evaluated. Here we examine a high resolution δ18O diatom record of three selected sections belonging from the Late Glacial to Early Holocene (ca 12,000 – 9,400 cal years BP) from Lago Chungará. We emphasise the role that some local factors such as sedimentary infill and palaeohydrology can play on the interpretation of the δ18O diatom record and therefore the need to discriminate between the climate and local environmental signals. 4.2 Results: Petrography and isotope composition of diatoms Smear slide, SEM, and several analyses (XRD, TC, biogenic silica) of the lake sediments before they were prepared for isotope analysis showed that the samples were composed of both amorphous and crystalline material. The amorphous fraction comprises biogenic silica (between 47-58 wt%), organic matter and volcanic glass. The crystalline fraction represented <10% of the sediments. 4.2.1 Interval 1 (11,990 – 11,530 cal years BP) Diatom concentration range from 108.3 to 633.8 million valves g-1. The interval is dominated by euplanktonic diatoms ranging from 79.1% to 93.9% of the diatom assemblage. The thicknesses of the laminae are between 0.9 and 10.3 mm (Fig. 4.1.A). Smear slide, thin section and SEM observations showed that light laminae were quasi-monospecific layers of large Cyclostephanos andinus (diameter > 50 μm). The upper contact of the light laminae with the dark laminae is transitional, showing an increase in diatom diversity with subdominant tychoplanktonic (Fragilaria spp.) and benthic diatoms (mainly Cocconeis spp., Achnanthes spp., Navicula spp. and Nitzschia spp.) (Fig. 4.2.C) whereas the lower contact is abrupt (Fig. 4.2.A). Diatom valves show good preservation with no preferred orientation in the lower part, but increasingly orientated upwards. The content of the organic matter also increases upwards. Dark laminae comprise a more diverse mixture of diatoms, including the euplanktonic (those having a strict planktonic character) smaller Cyclostephanos andinus (diameter < 50 μm) than those found in light laminae, and diatoms of the Cyclotella 60 stelligera complex, as well as tychoplanktonic (those usually having a benthic life form but which can occasionally be facultatively planktonic) and benthic diatoms (bottom dwelling forms). These dark laminae are also enriched in organic matter probably from diatoms and other algal groups. Up to 41 light and dark laminae couplets were defined. The thickness of these couplets ranges between 4.2 mm and 22.5 mm and, according to the chronological model they are pluriannual (mean about 10 years). The rhythmite starts with the dominance of light laminae progressively changing to a dominance of dark laminae. The δ18O diatom values of the purified diatoms in interval 1 range from +35.5‰ to +39.2‰ (Fig. 4.1.A). Higher δ18O diatom occur in the lower part of the interval (around 822 cm of core depth). There is an upwards decreasing trend (~1.9‰/100 years) attaining a minimum of +35.5‰ around 803 cm depth. This stretch is followed by an increasing shift of ~2.9‰/100 years towards the upper part of the interval where a relative maximum of +38.8‰ is reached at 793 cm depth. The uppermost two samples show a light depletion. The mean δ18O diatom value of this interval is +37.8±0.85‰. 4.2.2 Interval 2 (10,430 – 10,260 cal years BP) Diatom concentration ranged from 95.2 to 218 million valves g-1 in interval 2. Almost 94% of the diatom assemblages of this interval were made up of euplanktonic diatoms. Benthic taxa show the 18 38 39 40 41 42 17 16 15 14 13 12 11 10 9 8 6 3 5 4 2 1 7 574 570 Organic matter in diatomaceous ooze 100% 0% 560 550 540 537 9431 9500 9600 9700 9800 9891 d e p th (c m ) a g e (c a l. ye a rs B P ) C yc le s O (SMOW) 18 10 3635 37 38 39 40 9 8 7 6 5 4 3 2 1 610 620 622 605 10258 10300 10400 10426 d e p th (c m ) a g e (c a l. ye a rs B P ) cy cl e s O (SMOW) 18 11989 11531 11600 11700 11800 11900 790 788 800 810 820 830 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 3635 37 38 39 40d e p th (c m ) a g e (c a l. ye a rs B P ) cy cl e s O (SMOW) 18 A B C Figure 4.1. Digital images of the three intervals selected according to its depth and timescale. The identified couplets and the δ18O values from diatom silica have been plotted for interval 1 (A), interval 2 (B) and interval 3 (C). Stippled line shows mean values. 61 The palaeohydrological evolution of Lago Chungará during the Late Glacial and Early Holocene Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles minimum values for the three analysed intervals. The thickness of diatomaceous ooze laminae ranged from 1.8 mm to 16 mm (Fig. 4.1.B). Light laminae were dominated by large Cyclostephanos andinus (diameter > 50 μm) with some tychoplanktonic (Fragilaria spp.) and benthic diatoms, as well as minor amounts of siliciclasts and organic matter. Dark laminae are composed of a mixture of small and large Cyclostephanos andinus valves, with more abundant tychoplanktonic and benthic diatoms (as well as organic matter) compared to light laminae. Diatom valves are not so well preserved as in interval 1 sometimes showing a high degree of fragmentation and a preferred orientation. The contact between the laminae is similar to those found in interval 1. Clear couplets were only observed in the upper two thirds of the interval and only 10 couplets could be identified (Fig. 4.1.B). They are pluriannual (mean couplet represents about 10 years of sedimentation) and their thicknesses range between 5.5 and 19 mm. Light laminae were more abundant in the upper part of the interval 2, whereas dark laminae are more abundant in the lower part. The δ18O diatom curve shows a clear increasing trend during this interval (Fig. 4.1.B). The lowest δ18O diatom value (+36‰) was recorded at the bottom of the interval (617 cm depth) and the maximum at the two uppermost samples (+39.7‰ and +39.6‰; 606-605 cm of core depth). The magnitude of the increasing trend is much higher between the two lowermost samples (~18.5‰/100 years) than for the rest of the interval (~0.6‰/100 years). The mean δ18O diatom value of this interval is +38.7±1.4‰. 4.2.3 Interval 3 (9,890 – 9,430 cal years BP) Diatom concentration ranges between 163.8 and 255.8 million valves g-1 for interval 3. Euplanktonic diatoms (68.6% - 98.1%) also dominate this interval, and have the minimum values for the three intervals. On the contrary, benthic diatoms show moderate values (up to 31.4%), being the highest for the three intervals. Light diatomaceous ooze laminae ranged between 0.9 and 12.3 mm in thickness (Fig. 4.1.C), and they comprise Cyclostephanos andinus (diameter > 50 μm) increasing upwards in both taxonomic diversity and organic matter content. The lower contact with dark laminae shows an abrupt change in diatom size whereas the upper one is gradual. Diatom valves show good preservation with no orientation in the lower part but are preferentially oriented upwards. Dark laminae comprise a mixture of smaller Cyclostephanos andinus (diameter < 50 μm), with subdominant tychoplanktonic and benthic diatoms, as well as a high organic matter content. The lower contact is gradual whereas the upper one abrupt. Up to 18 light and dark pluriannual couplets were defined (mean couplet represent around 12 years). These couplets are 3 to 18 mm thick. The rhythmite starts with light laminae progressively changing to dark laminae. The δ18O diatom curve for interval 3 (Fig. 4.1.C) shows an overall continuous increasing trend of ~0.9‰/ 100 years from +39.1‰ (570 cm of core depth) to +41.3‰ (548 cm of core depth). Superimposed over the general trend are short-term fluctuations. The mean δ18O diatom value of this interval is +40.1±0.77‰. 62 500 m 500 m Light Light Dark Dark (d a y s ) d ia to m b lo o m p h a s e b a s e lin e c o n d it io n s p h a s e (s e v e ra l y e a rs ) Cyclostephanos andinus Cyclotella stelligera complex benthic or tychoplanktonic diatom organic-rich mud 3 -2 3 m m L ig h t D a rk A B C D 30 diatom size nutrients 50 90 m - + transition A D B C Figure 4.2. Rhythmite type showing thickness, colour, ecological succession and temporal scale. (A) SEM image showing the contact between dark (bottom) and light lamina (top). (B) Petrographical microscope image of the dark lamina. (C) SEM image showing the transitional contact between light (bottom) and dark lamina (top). (D) Petrographical microscope image of the light lamina. See text for details. The three intervals have different δ18O diatom averages displaying a progressive low-frequency enrichment from the interval 1 (+37.8±0.85‰) to interval 3 (+40.1±0.77‰). The overall isotopic enrichment is 2.1‰ throughout these intervals. 63 The palaeohydrological evolution of Lago Chungará during the Late Glacial and Early Holocene Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 4.3 Discussion 4.3.1 The sedimentary model of diatom rhythmites Laminated diatomaceous oozes in the sedimentary record of Lago Chungará comprise variable thickness couplets of alternating light and dark laminae. These couplets display different features (colour and mean thickness) in the three intervals described here although they exhibit similar diatom assemblages and textural characteristics and therefore it is assumed that their formation is by similar environmental processes. Rhythmite types have been established (Fig. 4.2); light laminae are formed almost exclusively by diatom skeletons of a quasi-monospecific assemblage of Cyclostephanos andinus, while dark laminae, with a high organic matter content, comprise a mixture of a more diverse diatom assemblage including the euplanktonic Cyclostephanos andinus although diatoms of the Cyclotella stelligera complex are the dominant taxa. Subdominant groups are some tychoplanktonic (Fragilaria spp.) and benthic taxa (Cocconeis spp., Achnanthes spp., Navicula spp., Nitzschia spp.). Each couplet was deposited during time intervals ranging from 4 to 24 years according to our chronological model. Couplets are therefore not a product of annual variations in sediment supply but due to some kind of pluriannual processes. The good preservation and size of diatom valves in the light laminae suggest accumulation during short-term extraordinary diatom blooms, perhaps of only days to weeks in duration. These diatom blooms could have been triggered by climatically driven strong nutrient inputs to the lake and/or to nutrient recycling under extreme turbulent conditions and mixing affecting the whole water column. On the contrary, the baseline conditions are represented by the dark laminae. Each of these laminae is made up of the remains (organic matter and diatom skeletons) of a diverse planktonic community deposited throughout several years under different water column mixing regimes. The preserved remains are therefore a reflection of different stages in the phytoplankton succession throughout several years (Reynolds, 2006). 4.3.2 Lake level and δ18O diatom changes A preliminary lake level reconstruction of Lago Chungará was undertaken employing the variations of euplanktonic diatoms, Botryoccocus and macrophyte remains (see Sáez et al. 2007). This reconstruction shows a general deepening trend during the Late Glacial and Early Holocene. This overall increase in lake level is punctuated by one deepening (D1; Fig. 4.3) and by two shallowing episodes (S1 and S2; Fig. 4.3). According to Sáez et al. (2007) the three selected intervals described here represent two different lacustrine conditions. Intervals 1 and 3 are likely shallower episodes that ocurred in different 64 climate periods, whereas interval 2 occurred during a period between two shallow intervals, and likely with higher lake level conditions. However, the resolution of the lake level reconstruction provided by Sáez et al. (2007) does not preclude the occurrence of other shallowing episodes than those previously detected. The isotope analyses presented here of these three intervals have allowed us to characterise the hydrological evolution of the lake for these different lacustrine conditions during the Late Glacial and Early Holocene. Dark laminae were selected for δ18O diatom analyses to investigate the baseline hydrological evolution of Lago Chungará. These dark laminae would represent a normal annual cycle of the lake with alternating phases of stratification and mixing. These conditions would lead to the development of a complex diatom community, among other algal groups (Hernández et al. 2007). The δ18O diatom variation can result from a variety of processes (Jones et al. 2004; Leng et al. 2005a) but for closed lakes, particularly in arid regions where water loss is mainly through evaporation, measured δ18O lakewater values are always more enriched than those of ambient precipitation since the oxygen lighter isotope (16O) is preferentially lost via evaporation. Under these circumstances, the δ18O diatom record can be used as an indicator of changes in the P/E related to climate changes (Leng and Marshall, 2004). Lago Chungará is a hydrologically closed lake and its main water loss is currently via evaporation, thus meaning that changes in δ18O values should be directly related to shifts in the P/E. The lake level change from the deeper water conditions recorded during the sedimentation of interval 2 to the shallower conditions occurred during the deposition of interval 3 according to the Sáez et al. (2007) reconstruction, is S1 D1 S2 S3? 9,000 9,500 10,000 10,500 11,000 11,500 12,000 LAKE LEVEL +- 37 38  18 O (SMOW) 39 40 41 Lg e a rl y H o lo c e n e AGE (cal yr BP) CLIMATIC PERIODS Figure 4.3. Lake-level evolution curve based on biological indicators (modified from Sáez et al. 2007). Deepening–shallowing episode (D1) and shallowing–deepening episodes (S1 and S2) are indicated. The lake followed an overall deepening trend (see Sáez et al. 2007 for further details). Shaded bands mark the three studied intervals. On the right corresponding mean values of δ18O from diatom silica of the studied intervals are shown. 65 The palaeohydrological evolution of Lago Chungará during the Late Glacial and Early Holocene Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles compatible with the observed increase in δ18O values. However, the isotope values and the lake level reconstruction do not agree over the transition from interval 1 to interval 2. The isotope values suggest a reduced P/E (shallower) stage, whereas several proxy indicators suggest deeper conditions (Fig. 4.3). A possible explanation for this could involve shifts in δ18O diatom related to other environmental circumstances, such as variations in the morphometrical parameters and changes in the groundwater outflow. Changes in the surface to volume ratio and in groundwater outflow of Lago Chungará from the Late Glacial to Early Holocene are the factors likely to account for most of the shifts found in the δ18O diatom values. Besides fluctuations in the P/E, another factor to take into account is basin morphology. During the lake’s evolution the lake’s surface to volume ratio would have changed. A tentative palaeobathymetric reconstruction of Lago Chungará based on the lake level curve from Sáez et al. (2007) (Fig. 4.4) shows that during the Late Glacial the lake only occupied the present central plain area. The rise in the lake level during the Early Holocene, although punctuated by some oscillations, flooded the extensive eastern and southern margins of the basin. Under these circumstances, the lake underwent a significant increase in its surface area (Fig. 4.4). Because the eastern margin is much shallower than the central plain (Fig. 2.7), the whole lake’s surface area to volume ratio would have significantly increased, and also concurrently the relative importance of evaporation. So the observed δ18O diatom high values of the interval 3 could be explained not only by the shallowing trend from interval 2 to interval 3, but also by the increasing of the lake’s surface to volume ratio between both intervals. There are no signs of subaerial exposure in the recovered sediments of the eastern platform, which indicates that lake water level did not drop significantly afterwards. Although the lake was deeper during interval 3 than during the interval 1, the mean isotope value is higher during interval 3. This fact could be explained by the increase of the surface to volume ratio and by the reduction of groundwater losses. Hence, the morphology of the lake, and not only water depth, must be considered as a key factor in any interpretation of the δ18O diatom in terms of changes in P/E. Furthermore, changes in the groundwater fluxes in Lago Chungará could have been a significant factor in the shifts found in the δ18O diatom values from the Late Glacial to Early Holocene. The groundwater outflow from the lake during the Late Glacial was probably higher than during the Holocene. This condition would progressively change with the sedimentary infill of the basin. Drainage, through the breccia barrier would progressively become less efficient as the groundwater outflows silted-up (Leng et al. 2005b). Thus, evaporative losses would have predominated over groundwater during the Early Holocene. This highlights the fact that stable isotopes would not have, in this case, a direct correspondence with changes in the lake water level. In summary, the relative increase in evaporation due to the increase in the lake’s surface to volume ratio between the studied intervals could have played a significant role. Superimposed onto this situation, the increase in the δ18O diatom values from the Late Glacial (when the lake was at its shallowest) to the Early 66 2 km 11 Lateglacial-early Holocene 11,990-11,531 cal years BP early Holocene I 10,430-10,260 cal years BP early Holocene II 9,890-9,430 cal years BP Diatomites (Subunit 1a) Diatomites (Subunit 1b) Massflows deposits Groundwater outflow towards Cotacotani Pre-collapse fluvial deposits Miocene volcanic rocks Collapse deposits A-A’ A A’ 1 Km Deltaic deposits Lake Boundary of present lake evaporation 11 11 A B C Figure 4.4. Hydrological evolution of the Lago Chungará in the Lateglacial–early Holocene. North–South cross-section of the lake (left) and water lake surface area (right) for the sedimentation of interval 1 (11,990–11,530 cal. years BP (A)), interval 2 (10,430–10,260 cal. years BP (B)) and interval 3 (9,890–9,431 cal. years BP (C)). Holocene (when the overall deepening trend started) is also likely to have been related to a change to a predominantly evaporative lake as the lake’s bottom became more impermeable due to sediment basin sealing. 4.4 Conclusions The thin section study of the diatomaceous laminated sediments shows that the rhythmites are made up of light quasi monospecific lamina of the euplanktonic diatom Cyclostephanos andinus and a pluriannual dark lamina rich in organic matter and a mixture of a more diverse diatom assemblage. The formation of light laminae is apparently related to short term (days to weeks) diatom blooms whereas dark laminae represent baseline conditions lasting several years. 67 The palaeohydrological evolution of Lago Chungará during the Late Glacial and Early Holocene Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles The oxygen isotope record of the dark laminae diatoms of Lago Chungará indicates a progressive δ18O enrichment from the Late Glacial to Early Holocene. Besides changes in the P/E, two other factors could have governed shifts in the Lago Chungará δ18O diatom record. The basin’s stepped morphology forced the expansion of the lake towards the eastern and southern shallow margins during the rising trend. These changes could have caused an increase in the lake’s surface to volume ratio thus enhancing the evaporation which caused isotope enrichment during the Early Holocene. In addition, changes in the groundwater/evaporation loss ratio and changes in the lake’s extent. The hydrology of the lake was probably modified during the Late Glacial to Early Holocene transition as the lake’s groundwater outflow became progressively sealed by sediments, thereby increasing lake water residence time and potential evaporation. Previous work has focused on issues of diagenesis, contamination and host-water interactions that can all influence δ18O diatom whereas local hydrological factors have been largely neglected. These results point to the complex interplay among the different factors which intervene in the diatom oxygen isotope record of closed lakes and how interpretation needs to be adapted to the different evolutionary stages of the lake’s ontogeny. This study highlights the importance of reconstructing local palaeohydrology as this may be only indirectly related to palaeoclimate. 68 Chapter 5 ENSO and solar activity signals from oxygen isotopes in diatom silica during the Late Glacial- Holocene transition in Central Andes (18ºS)* 5.1 Introduction The study of Andean Altiplano lacustrine records plays a prominent role for interpreting the Quaternary palaeoclimatic history of the South American tropics and therefore for understanding the function of the tropics in the Earth’s climate system (Grosjean et al. 2001; Valero-Garcés et al. 2003; Placzek et al. 2006) (Fig. 2.3). For this reason, studies on the sedimentary records from this area have increased in the last few decades. Most of these studies have focussed on the reconstruction of climate events at millennial time scales, especially since the Last Glacial Maximum (Baker et al. 2001a). There is a general consensus that orbital forces are the main factor triggering the climate conditions at a millennial- scale (Rowe et al. 2002; Placzek et al. 2006), and are therefore responsible for those climate events. Superimposed onto this long term variability, changes in the hydrologic balance at a sub-millennial scale in the Andean Altiplano, have been attributed to the variability of the Pacific SSTs and the strength of the zonal winds (Rowe et al. 2002; Garreaud et al. 2003). Both factors are controlled by ENSO and changes in the solar activity (Theissen et al. 2008). A number of studies have detected multidecadal- to centennial-scale hydrological balance shifts, suggesting that these relationships have been active since, at least, the Mid-Holocene (Valero-Garcés et al. 2003; Theissen et al. 2008). δ18O diatom are increasingly being used for palaeoenvironmental reconstructions in lacustrine sedimentary records (Rietti-Shati et al. 1998; Barker et al. 2001). However, application of this proxy to high-resolution centennial to millennial lacustrine records is still in its infancy (Barker et al. 2007). δ18O diatom in decadal-to-centennial resolution palaeoclimatic reconstructions has not been utilised, mainly due to the difficulty in obtaining high resolution samples from sites with sufficient variation in δ18O diatom (outside of analytical error) that can be characterised at this fine temporal scale. Additional difficulties *Chapter based on the paper published in: Journal of Paleolimnology (In press). Armand Hernández, Santiago Giralt, Roberto Bao, Alberto Sáez,Melanie J. Leng, Philip A. Barker. 69 in using δ18O diatom are related to the difficulty in obtaining monospecific diatom samples in order to eliminate any species-specific effect variability, to acquire the necessary amount of sample from these short periods of time, and to have pure diatom samples, since significant contaminants can produce excursions in δ18O diatom that are similar to those produced by climate variations (Brewer et al. 2008). The diatomaceous ooze from Lago Chungará has previously been the subject of a preliminary diatom oxygen isotope study at low resolution. This earlier study was aimed at three non consecutive stretches of the sedimentary record, and did not include all the dark-green laminae (Hernández et al. 2008). For the present study we have analysed 40 consecutive dark-green laminae, corresponding to the Late Glacial and Early Holocene, which represent a continuous record of the background limnological conditions (Hernández et al. 2008) (see the sedimentary model in the sedimentary sequence and rhythmite type section below). The excellent preservation and high diatom content of the record of Lago Chungará allow a detailed study of the regional moisture balance at decadal and centennial timescales. Here, we present the decadal to centennial time scale moisture balance reconstruction for the Andean Altiplano during the Late Glacial-Holocene transition (11,990-11,450 cal years BP) based on high-resolution analysis of δ18O diatom . This analysis was performed on successive and continuous 40 dark-green laminae of lacustrine sediments present in a core located in the offshore zone of Lago Chungará. In order to support the interpretation, isotope data are compared with the reconstructions of the terrigenous inputs and inferred regional effective moisture in the Lago Chungará performed in the same core by Giralt et al. (2008). 5.2 Results 5.2.1 Oxygen isotopes The δ18O diatom record (Fig. 5.1.E) shows both short-term (decadal) and long-term oscillations (centennial time scales) ranging from +35‰ to +39.2‰ (mean= +37.4 ± 0.8‰). From the bottom to the top, the studied record can be subdivided into three phases. These intervals correspond to three enrichment/depletion phases (Fig. 5.1.D). Each phase starts with a continuous centennial isotope enrichment which abruptly ends with a sharp depletion: Phase 1. Lower interval (11,990 to 11,800 cal years BP). It shows the maximum and minimum δ18O diatom values (+39.2‰ and +35.1‰ respectively, with a mean value of +37.7 ± 1‰) throughout the whole record. It starts with an increasing trend of ~3.3‰/100 years which finishes at 11,860 cal years BP. This trend is followed by a shift to lighter values of ~8.1‰/100 years with a sharp final decrease in the δ18O diatom values of 3.5‰ in less than 10 years, acquiring the minimum value for the whole record at ca 11,800 cal years BP. Both trends are interrupted by ca 5 to 20 years depletion/enrichment excursions ranging between ±0.9 and ±1.7‰. 70 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles Phase 2. Middle interval (11,800 to 11,550 cal years BP). This section (mean values +37.3 ± 0.7‰) starts with an enrichment trend showing an upwards gradient of ~1.3‰/100 years which finishes at 11,570 cal years BP with a +38.3‰ δ18O diatom value. This trend is however punctuated by one sudden rise (+2.3‰) and up to four minor depletions (ranging from -0.6 to -1.3‰.) of the δ18O diatom values on a 40 to 55 years basis. The enrichment trend is followed by a shift of ~9.1‰/100 years to lighter values reaching a minimum value of +36.2‰. Phase 3. Upper interval (11,550 to 11,450 cal years BP). This interval (mean values +37.2 ± 0.7‰) also starts with an enrichment trend but, because the section only comprises three samples, this enrichment has not been estimated. This trend is also followed by depletion of 1.3‰ in 10 years. 5.2.2 Spectral analyses of the diatom oxygen isotope record MTM performed on the δ18O diatom values shows a number of clear periodicities (Fig. 5.2). Almost all identified periodicities (7.2, 8.9, 11.1, 13, 18.6, 22.3 and 39.4 years) exceed the 99% confidence interval  18 O SMOW diatom ( )‰ 35 36 37 38 39 40 11,990 11,475 11,600 11,500 11,700 11,800 11,900 Age (cal yrs BP) Isotope phases P h a s e 1 L a te g la c ia l e a rl y H o lo c e n e P h a s e 2 P h a s e 3 Terrigenous input Giralt et al. (2008)Present work Effective moisture availability NO DATANO DATA Wet DryWet DryWet Dry Sáez et al. (2007) Hernández et al. (2008) 0 100 200 300 400 500 600 Depth (cm) - -+ + Planktonic diatoms (%) 20 60 100 38  18 O (SMOW) 40 42 H o lo c e n e LG Increasing trend Depletion trend a b c d e e ef fA D E F GB C Figure 5.1. δ18O diatom data for the period 11,990–10,475 cal year BP from Lago Chungará, compared with other paleoenvironmental records of the lake. A) Planktonic diatoms percent abundance curve for the whole Lago Chungará sequence (Sáez et al. 2007). B) δ18O diatom data of non-consecutive dark-green laminae from three intervals of the record (Hernández et al. 2008). C) Photography of laminated sediments corresponding to the sampled interval of subunit 1a in core 11. D) Oxygen isotope enrichment/depletion phases, in the studied interval, interpreted from the data. E) δ18O diatom data from the present study and interpretation in terms of wet and dry conditions. The values correspond to 40 consecutive dark-green laminae throughout the whole selected interval. F) Terrigenous input variations derived from the first eigenvector of PCA on magnetic susceptibility, XRF, XRD, TC and TOC, BSi (Giralt et al. 2008). G) Effective moisture availability variations from the second eigenvector of the mentioned PCA (Giralt et al. 2008). Correlation lines correspond to the main oxygen isotope depletion peaks. Note that the main trends of the three curves are similar but there is a systematic temporal disagreement between them. 71 ENSO and solar activity signals during the Late Glacial-Holocene transition in Central Andes (18ºS) whereas only two (3.7 and 8 years) lie between 95-99% confidence interval (Fig. 5.2). Most of the sub- decadal identified frequencies are close to the minimum temporal resolution of the sampling (4.1 years), which explains in great part the weaker intensity of the short periodicities between 3 and 8 years. Therefore, only the most significant frequencies and above the minimum temporal sampling resolution have been taken into account in the discussion. TF analysis reveals the strongest energy for the lower values of frequency, mainly focussed on the 35-years cycles, whereas it decreases towards higher frequency values, i. e., the higher periodicities (Fig. 5.3). This fact can mostly be explained by the decadal sampling resolution, making periodicities lower than ten years less significant. Additionally, TF analysis indicates that the highest energies of the significant frequencies are located in the Late Glacial period between ca 11,950 and 11,700 cal years BP, decreasing just from the onset of the Holocene until, at least, approximately 11,550 cal years BP (Fig. 5.3). TF diagram also highlights that the identified frequencies did not have the same intensity (energy) during all the studied period. For instance, the shortest significant periodicity observed in the MTM (7.2 years) was mainly active during the first 150 years of the record, whereas it was only active during three short time windows in the following 500 years. A similar pattern is also observed for the rest of the significant periodicities (8.9, 11.1, 13, 18.6, 22.3 and 39.4 years). The maximum energy areas correspond to depletions in the δ18O diatom values, i.e. 11,800 and 11,550 cal years BP (Fig. 5.3). 3 .7 3 y e a rs 7 .1 6 y e a rs 8 y e a rs 8 .9 0 y e a rs 1 1 .1 3 y e a rs 1 2 .9 7 y e a rs 1 8 .6 2 y e a rs 2 2 .2 7 y e a rs 3 9 .3 7 y rs harmonic MTM Spectrum: Data vector O, npi=2, ntpr=3 18 Frequency (units) P o w e r reshaped median 90% 95% 99% 0 0.0001 0.01 1 100 0.1 0.2 0.3 0.4 0.5 Figure 5.2. Multi-taper analysis of the δ18O diatom values. The 90, 95 and 99% confidence levels are indicated and significant periodicities are shown. Note that periodicities with more than 99% of significance are shown in black and those with more than 95% significance in blue. 72 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 7.16 - 8.9 years (Strong ENSO: 7 - 9 years) 11.13 - 12.97 years (Schwabe cycles: 11 years) 18.62 years (Strong ENSO: 15 - 17 years) 22.27 years (Hale cycles: 23 years) 39.37 years (Brückner cycles: 35 years) 0.01 11,500 40 34 11,500  18 Odiatom 11,600 11,600 Time [cal. years BP] Time [cal. years BP] Energy F re q u e n c y [H z ] 11,700 11,700 11,800 11,800 11,900 11,900 11,950 11,950 0.5 0.02 0.05 0.1 0.2 EARLY HOLOCENE LATEGLACIAL Figure 5.3. Time–Frequency analysis of the δ18O diatom values. Pink indicates high energy whereas blue displays low energy areas. Energies below 0.03 were clipped in order to facilitate understanding of the graph. Red and blue horizontal bands mark different frequency bands of the ENSO and solar activity forcings. Yellow vertical bands show zones with δ18O diatom shifts and their corresponding power values for each frequency. A weakening pattern in ENSO and solar activity energies can be observed through the Late Glacial-Early Holocene transition. 5.3 Discussion 5.3.1 Controlling factors of δ18O diatom in Lago Chungará δ18O diatom in lake sediments is controlled by the δ18O lakewater , temperature, and the possible disequilibrium by vital effects or diagenesis (Leng and Barker, 2006). We discount vital effects and diagenesis as analyses were made on near-monospecific diatom samples and preservation of the diatom frustules is excellent (Fig. 5.4). δ18O lakewater depends on the balance between the isotope composition of water inputs (including the source and amount of precipitation, surface runoff and groundwater inflow) and outputs (evaporation and groundwater loss) in the lake. The measured δ18O of the inputs (springs, Río Chungará and rainfall) in the Lago Chungará is homogeneous, giving values close to the δ18O precipitation (Fig. 2.5.A). δ18O precipitation is 73 ENSO and solar activity signals during the Late Glacial-Holocene transition in Central Andes (18ºS) a function of the isotope composition of the moisture source and air-mass trajectory, but in the Lago Chungará there are no changes in the moisture source composition since the air masses always come from the Atlantic Ocean throughout the Amazon basin (Grosjean et al. 1997). During moisture transport from the Atlantic to the lake area, three processes are directly responsible for the low and variable values of the present δ18O precipitation throughout the Andean Altiplano (Aravena et al. 1999). These processes include interaction of the air masses within the Amazon basin, an altitude effect due to the ascent of the air masses along the eastern slope of the Andes, and the convective nature of the storms in the Altiplano region. Nevertheless, in the Lago Chungará region the values obtained for the measured δ18O precipitation are relatively stable with almost all values around –14 and –20‰ (Herrera et al. 2006) (Fig. 2.5.A) , whereas δ18O lakewater is much higher (Fig. 2.5.A). This result is in accordance with a δ18O lakewater enrichment via evaporation. Thus, any isotopic variation of δ18O lakewater will be more related to changes in the amount of precipitation («amount effect») and evaporation rather than to the variability of δ18O precipitation . Evaporation enriches δ18O lakewater by 14‰ relative to the inlets (precipitation, springs and river) in the present day (Fig. 2.5.A). During the Late Glacial and Early Holocene the water residence time of the lake was shorter than present because of the different palaeohydrological context, but even so it can be considered closed for that period (Hernández et al. 2008). Accordingly, the variations in the δ18O diatom must be mainly derived from changes in the δ18O lakewater resulted from shifts in the P/E, rather than dominated by temperature. However, two factors should be considered in the interpretation of the δ18O diatom values in terms of temperature oscillations. The first factor is related to δ18O precipitation that correlates directly with changes in the air temperature. The global relationship between changes in δ18O precipitation with air temperature is commonly referred to as the ‘Dansgaard relationship’, and it implies changes between +0.2 and +0.7‰/ºC (Dansgaard, 1964). The second is the water of the lake temperature dependence of oxygen isotope fractionation between diatom 50 m Figure 5.4. Diatom-rich sediment from Lago Chungará after the cleaning process. Large Cyclostephanos andinus valves are the unique component. 74 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles silica and the lake water (Brandriss et al. 1998). Nevertheless, the fractionation factor value of this temperature dependence is still controversial. Published fractionation factors range from –0.2‰ and – 0.5‰/ºC (Brandriss et al. 1998; Moschen et al. 2005). The two temperature factors have opposing effects on δ18O diatom but, owing to its larger variability, the effects of the first factor (air temperature) usually dominate over the second. However, even in the case of the largest change due to the Dansgaard relationship, its magnitude will be greatly damped by the effect of the isotope fractionation between diatom silica and lake water. Moreover, it is known that most of the tropical rainfall isotope datasets exhibit a far stronger correlation with total precipitation than with air temperature (Leng et al. 2005b), indicating in the Lago Chungará case a magnification of the P/E in wetter periods. Hence, we can assume that in the Lago Chungará the effects of precipitation variability and temperature oscillations in the δ18O diatom values will be small in comparison to evaporative concentration, as pointed by other authors for closed lakes in general (Gat 1980; Gasse and Fontes 1992). 5.3.2 Variations of the P/E in the lake Oxygen isotopes have widely been used to carry out lake level reconstructions and to establish consequent palaeoclimatic interpretations (Barker et al. 2001; Valero-Garcés et al. 2003). There is a relationship between lake level change and the P/E for Lago Chungará during the Late Glacial and Early Holocene, but this dependency is hampered by local palaeohydrological factors such as changes in the groundwater outflow and shifts in the lake surface/volume ratio which produce a background long term enrichment trend (Hernández et al. 2008). This effect is however negligible when considering isotopic changes at a decadal to centennial time scale. Both present (Fig. 2.5.A) and past (Thompson et al. 1998) rainfall isotope values in the Lago Chungará region are much lighter than those measured for the water of the lake, and the magnitude of the long-term enrichment trend is very small compared to them. Therefore, depletions of δ18O diatom would directly be related to wet episodes in the Andean Altiplano, whereas exceptionally high values, which stand out over the general enrichment trend, would indicate arid episodes. The observed δ18O diatom enrichment trends agree with periods where light-white laminae are more common, whereas depletion episodes coincide with poorly developed and less abundant light-white laminae (Fig. 5.1.C and D). These light-white laminae are most likely the result of exceptional periods of mixing of the shallow water column during lowstands, which recycle nutrients from the hypolimnion and therefore trigger extraordinary diatom blooms (Hernández et al. 2007). This interpretation is also supported by terrigenous input and regional effective moisture reconstructions previously performed 75 ENSO and solar activity signals during the Late Glacial-Holocene transition in Central Andes (18ºS) on the Lago Chungará sedimentary record (Giralt et al. 2008) (Fig. 5.1 .F and G). These reconstructions were carried out by applying multivariate statistical analyses (Cluster, Redundancy Analysis (RDA) and PCA) to magnetic susceptibility, XRF, XRD, TC, TOC, TBSi and grey-colour curve data. The terrigenous inputs curve was derived from the first eigenvector of the PCA, whereas the regional effective moisture reconstruction was obtained from the second eigenvector. For the lower part of Chungara sequence (Unit 1), the more positive values of the terrigenous inputs were interpreted, as increasing erosion rate of catchment volcanic sediments, suggesting humid conditions. Similarly, the effective moisture availability proxy depends on the P/E, with positive values corresponding to drier conditions (Giralt et al. 2008). The comparison of the three proxies (Fig. 5.1.E, F and G) shows that the hydrological response of the diatom silica oxygen isotopes (a biological proxy) and of the other two reconstructions to the environmental variations is not the same. The main trends in the three curves (Fig. 5.1.E, F and G) are similar but there is a systematic temporal disagreement (ranging between ca 5 and 50 cal years BP) between the terrigenous inputs and the regional effective moisture availability (which both react first) and the δ18O diatom (reacting afterwards). This time lag between the two proxies highlights the complex and non-linear response of the lacustrine ecosystem to environmental forcings (Fritz, 2008). After rainfall the increased runoff and input of terrigenous material is almost immediate. On the contrary, the oxygen isotope homogenisation of the water of the lake which later will be incorporated on the diatom frustule, has a delayed time of response. This depends on the epilimnion water residence time and, furthermore, whether the lake is hydrologically closed or not. Hence, the observed time lag can be showing these different responses of the system to the same forcing. However, we cannot discount the poorly understood concept of silica maturation, where pores in the silica matrix close through early diagenesis creating differences in the δ18O between living diatoms and sediment assemblages (Schmidt, 2001) and therefore a lag in the δ18O diatom record. At centennial-scale, the Lago Chungará isotopic values show a general pattern of increasing δ18O diatom (Fig. 5.1.B), with an enhanced enrichment period at the bottom, but interrupted by three major depletion events. The depletion events, accentuated by the «amount effect», correspond to heavy rainfall conditions, whereas enriched values would indicate exceptionally dry conditions favouring the evaporation. This interpretation is reinforced by the terrigenous input and effective regional moisture availability independent reconstructions. Three wet/dry phases have been identified in the δ18O diatom record (Fig. 5.1.D). Phase 1 (11,990 – 11,800 cal years BP) shows a significantly increased gradient in δ18O diatom suggesting that dominantly dry climate conditions played a key role triggering this isotope enrichment. Because of this drier situation the lake level would be lower, as also indicated by the important development and major presence of light-white laminae in this part of the interval. Three low-intensity and short-term wet episodes punctuate 76 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles the established Late Glacial arid period (Fig. 5.1.E). These episodes can also be recognised and correlated with events of increased terrigenous inputs and effective moisture availability (Fig. 5.1.E, F and G). The much weaker isotope enrichment for phase 2 (11,800 and 11,550 cal years BP) can be mainly ascribed to the general low magnitude palaeohydological background trend towards heavier isotope conditions of the Late Glacial-Early Holocene transition (Hernández et al. 2008). This fact, together with the poorer development and minor presence of the light-white laminae with respect to the previous interval, suggests that the enrichment via evaporation was much less important than during the sedimentation of phase 1, corresponding to a more humid period. Furthermore, the terrigenous inputs and effective regional moisture availability curves show relatively wetter conditions for this period (Fig. 5.1.F and G). This trend is also punctuated by a sudden rise in the lowest part of the interval indicating a short dry event and slight depletions in δ18O diatom indicating wet decadal-scale events (Fig. 5.1.E). In phase 3 any clear trend is difficult to identify (Fig. 5.1.E). Although the δ18O diatom record seems to show a new trend towards drier conditions after the sudden wet event dated at 11,550 cal years BP, the lack of suitable samples has hampered any firm conclusions. 5.3.3 Long-term, centennial- to millennial-scale palaeoclimatic implications There are many Late Quaternary palaeoclimatic reconstructions from the Andean Altiplano region (Sylvestre et al. 1999; Rigsby et al. 2005) but the climate context for the Late Glacial-Holocene transition still remains unclear. Some authors have defined a cold period (12,600-11,500 cal years BP) coincident with the Northern hemisphere’s Younger Dryas event (Baker et al. 2001b). The wet («Coipasa phase», Thompson et al. 1998; Placzek et al. 2006) or dry (Maslin and Burns 2000; Weng et al. 2006) character of this event remains controversial. On the contrary, other authors consider this period just the final part of the deglaciation towards the present Interglacial («Ticaña phase», Sylvestre et al. 1999), as part of a long-term dry pattern (Rowe et al. 2002; Abbott et al. 2003). The previous lake level reconstruction, mainly based on the abundance of planktonic diatoms, shows a shallowing followed by a long term rising trend for the interval presented here (Sáez et al. 2007). Additionally, recent data on the Lago Chungará record, mainly based on XRF core scanner analysis, has established the Late Glacial to Holocene transition as a relatively wet period (Giralt et al. 2008). The centennial scale δ18O diatom record is congruent with the lake level reconstruction performed by Saéz et al. (2007) which represents the palaeoclimatic evolution related to the major lake level variations (Fig. 5.1.B). The non continuous isotopic data (Fig. 5.1.B) also displays a persistent, but minor, background isotope enrichment trend. This enrichment is related to changes in the lake morphology due to shifts in its surface/volume ratio, as well as changes in the groundwater outflow during the lake ontogeny 77 ENSO and solar activity signals during the Late Glacial-Holocene transition in Central Andes (18ºS) (Hernández et al. 2008). In any case, the new δ18O diatom data presented here highlights that the Glacial- Interglacial transition in the central Andean Altiplano was punctuated by abrupt and high-frequency centennial climate variability. 5.3.4 Short-term, decadal- to centennial-scale palaeoclimatic implications Millennial-scale shifts in the Atlantic-Amazon-Altiplano hydrologic system have been attributed to orbitally induced changes in solar insolation, coupled with long-term changes in the ENSO variability (Rowe et al. 2002; Abbott et al. 2003; Servant and Servant-Vildary 2003). However, higher-resolution changes are not directly related to orbitally induced insolation forcing (Abbott et al. 2003). The interannual climate variability in the Andean Altiplano is most likely related to changes in the Pacific Tropical SSTs, and the sign and strength of the zonal winds above the Altiplano (Garreaud et al. 2003). Both factors would affect the strength and position of the Bolivian high and, hence, the moisture distribution over the region. The main force controlling the SSTs is the ENSO variability, involving dry or wet situations in the Altiplano during El Niño- or La Niña-like conditions respectively (Garreaud et al. 2003; Vuille and Werner, 2005). This is consistent with instrumental data from the Chungará area where precipitation is reduced during moderate to intense El Niño years (1965, 1972, 1983, and 1992) (Fig. 2.5.C). Additionally, the sign and strength of the zonal winds above the Altiplano would be modulated by decadal and multidecadal variations in solar activity, possibly related to the mode of the ENSO system (Theissen et al. 2008). Although ENSO modulation by solar activity has been suggested (Velasco and Mendoza, 2008), no clear relationship has been demonstrated between both forcings. Nevertheless, there is broad agreement that ENSO events are the main control governing the moisture distribution in the Altiplano (Servant and Servant-Vildary, 2003), and that decadal-scale changes in the effective moisture could be related to the solar activity during the Mid-Holocene (Theissen et al. 2008). The results presented here would suggest a similar pattern during the Late Glacial-Holocene transition over the Andean Altiplano (Fig. 5.3). The identified frequencies can be attributed to different periodicities of the solar activity cycles such as Schwabe 11 years (identified as 11.1 and 13 years), Hale 23 years (22.3 years) and Brückner 35 years (39.4 years), and of the ENSO frequency (main frequency at 7-9 years (7.2 and 8.9 years) and its decadal frequency 15-17 years (18.6 years)). The influence of solar activity and ENSO variability on the isotope record is supported by the fact that several periodicities concordant with both forces were identified. The time-frequency analysis suggests that the driest period (11,950 - 11,800 cal years BP) was ruled by high solar activity, mainly represented by a Brückner cycle, and strong ENSO-like conditions. 78 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles The ENSO and solar activity signals remain present for the Early Holocene period (between 11,750 until 11,500 cal years BP), although they show a weakening pattern through this period (Fig. 5.03). This fact is congruent with the progressive weakening of the ENSO suggested by other authors for the Late Glacial-Holocene transition (Rodbell et al. 1999; Moy et al. 2002; Rodó and Rodriguez-Arias, 2004). In Lago Chungará, the onset of the Holocene was characterised by minor δ18O diatom enrichment by evaporation and by the occurrence of multi-decadal weak depletions that would be governed by the more humid La Niña-like conditions. This would agree with previous observations that suggest a reduction in the El Niño intensity within the region during the Early-Holocene in favour of long-term La Niña-like conditions in the tropical Pacific (Betancourt et al. 2000; Koutavas et al. 2002). 5.4 Conclusions The Late Glacial to Holocene transition from the Lago Chungará record is made up of laminated diatom-rich sediments which provide excellent material for the application of oxygen isotope analysis in biogenic silica. δ18O diatom data have for the first time provided palaeoclimatic reconstruction at decadal- to-centennial resolution. The well-laminated nature of these sediments allowed a lamina by lamina continuous sampling, giving one of the highest resolution records available for δ18O diatom . It has also revealed important insights into the usefulness of this method, as well as provided decisive palaeoenvironmental information for this critical period. δ18O diatom from dark-green diatom laminae represent the baseline in the environmental evolution of Lago Chungará, and show decadal to centennial variability in the moisture conditions of the Andean Altiplano. The isotopic record displays a persistent background isotope enrichment trend related to changes in the lake morphology and groundwater outflow during the Late Glacial and Early Holocene. Overprinted onto this long-term (centennial to millennial) trend there are cyclically short-term (decadal to centennial) shifts which are not related to changes in temperature or isotopic composition of the source of precipitation, but to the P/E variability in the Altiplano. The record shows two major isotope depletions, occurring at a centennial time scale (11,800 and 11,550 cal years BP) indicating a long-term increase in moisture conditions, and one major isotope enrichment above the background levels that occurred between 11,990 and 11,800 cal years BP indicating a short dry phase during the Late Glacial. Minor depletions at a decadal time scale are associated with weaker rainfall short-term events. The comparison with terrigenous input and effective moisture availability reconstructions previously performed for Lago Chungará shows agreement, but includes a systematic lag time (up to 50 years) among these proxies and δ18O diatom . This is mainly due to the time necessary to change the δ18O lakewater values and its subsequent incorporation into the diatom frustules, 79 ENSO and solar activity signals during the Late Glacial-Holocene transition in Central Andes (18ºS) but other factors should not be completely disregarded. The time lag highlights the fact that not all the proxies react at the same time to environmental forcing and this needs to be more often recognised in high resolution palaeolimnological reconstructions. Sub-millennial shifts in the hydrological balance of Lago Chungará are hypothesised to be the result of changes in the strength and position of the Bolivian High. Spectral analyses of δ18O diatom suggest that these changes in the atmospheric conditions over the Altiplano during the wet events were triggered by both ENSO and solar activity. The change from the Late Glacial dry period to a wetter Early Holocene period confirms a weakening of El Niño intensity in the Andean Altiplano region in favour of La Niña-like conditions found elsewhere. Nested upon the underlying climate dynamics are the different cyclicities of solar activity (Schwabe, Hale and Brückner) that were active during different time windows. There is undoubtedly an interaction between these and ENSO at the decadal and greater scales and it is likely that apparent solar forcing of the Lago Chungará record is transmitted via ENSO modulation of the South American monsoon. The complexity of Andean Altiplano palaeoenvironmental conditions, and the absence of other high resolution studies for this time interval, does not allow us to establish any clear conclusion on the existence of significant climate events synchronous to the Younger Dryas in the northern hemisphere. While many studies have demonstrated ENSO-like forcing during the Glacial-Interglacial transition, this highly resolved record is one of the few that preserves key ENSO frequencies, therefore further implicating this major climate process with events governing the transition to the Holocene. 80 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles Chapter 6 Biogeochemical processes controlling oxygen and carbon isotopes of diatom silica in lacustrine rhythmites* 6.1 Introduction Rhythmites are finely laminated sequences (millimetre- to submillimetre thick) made up of regular alternations of two or three contrasting sediment types called couplets or triplets (Talbot and Allen, 1996). Rhythmite formation is generally associated with seasonally heterogeneous sediment supply and a lack of physical or biological reworking processes (Grimm et al. 1996). Thus, laminated sediments indicate high-frequency environmental change through time. A number of studies have described laminated lacustrine sediments, but they have mainly dealt with annual-rhythmites (varves) with different clastic grain-size and/or biogenic content deposited over different seasons (e.g. Bird et al. 2009). At mid- to high latitudes the processes that lead to rhythmite formation are often well constrained (e.g. Chang et al. 2003), whereas the biogeochemical processes and climate events which prompt laminated sediments in tropical lacustrine sediments are often less understood. In these cases, tropical rainfall regimes associated with intense storms and wind may be responsible for extraordinary external nutrient loading or upwelling of nutrient rich-waters which trigger phytoplankton blooms (Talbot and Allen, 1996). These tropical climate regimes follow a seasonal behaviour (e.g. monsoons), but they can also be highly influenced by climate multiannual phenomena (e.g. ENSO). Changes in δ18O diatom in lacustrine sediments are used to infer hydrological variations. For closed lakes in the tropics, these variations are mostly related to the P/E, which is generally directly linked to lake level change (Leng and Barker, 2008). The isotope-inferred reconstructions can thus be used to unreveal the climate history of the region (e.g. Barker et al. 2007) although this may be mitigated by biological and sedimentary processes. Besides δ18O diatom , the δ13C diatom , can give other relevant palaeoenvironmental information, including insights on the lakes’ carbon cycle. There are few studies of *Chapter based on the paper submitted in: Palaeogeography, Palaeoclimatology, Palaeoecology (submitted). Armand Hernández, Roberto Bao, Santiago Giralt, Philip A. Barker, Melanie J. Leng, Hilary J. Sloane, Alberto Sáez. 81 carbon isotopes from organic inclusions within diatom frustules, and of those published, most have dealt with marine sedimentary records (e.g. Crosta and Shemesh, 2002). Studies on δ13C diatom in lake sediments are now emerging and providing valuable insights into the complex carbon cycle of lakes (Hurrell et al. submitted). The aim of this paper is to understand high frequency biological, chemical and sedimentary processes which cause the laminae formation in the sedimentary record of Lago Chungará, a high altitude tropical lake located in the Central Andes. δ18O diatom and δ13C diatom data from individual lamina are presented for a period between 11,990 and 11,530 cal years BP. High frequency environmental perturbations brought about by interannual-decadal climate events are rarely recorded in lake sediments, and therefore, the laminated sediments here are a good record of their intensity and their effect on lacustrine hydrological and carbon cycles. 6.2 Results 6.2.1 Laminae biogenic composition The present study extends the petrographical examination of the diatomaceous laminated sediments of Lago Chungará for the Late Glacial to Early Holocene transition (11,990 - 11,530 cal years BP) described in Hernández et al. (2008). A hundred laminae have been differentiated and grouped under the white, light-green and dark-green laminae categories according to their diatom composition, organic matter content and colour. Additionally, nine laminae were undifferentiated due to their mixed features between the three groups (Figure 6.1). White laminae are formed almost exclusively by diatom frustules of the large (diameter > 50 μm) euplanktonic diatom Cyclostephanos andinus (Fig. 6.2.G). Dark-green laminae, which contain a higher organic matter content, probably derived from diatoms and other algal groups, are made up of a mixture of different diatom species. This mixture is mainly composed of smaller (diameter < 50 μm) Cyclostephanos andinus valves, with diatoms of the Discostella stelligera species complex as co-dominant taxa. Subdominant diatom taxa comprise a number of tychoplanktonic (mainly Staurosira construens aff. venter and Fragilaria spp.) and benthic life forms (including Cocconeis placentula, Gomphonema minutum, Nitzschia tropica and Opephora spp. aff. mutabilis) (Fig. 6.2.C). The light-green laminae are made up of components from the white laminae progressively grading upwards to the typical constituents of the dark-green laminae. Diatoms of the light-green laminae are usually embedded in an organic matrix creating a preferential orientation of the valves (Fig. 6.2.B and E). Thus, a lower white lamina, an intermediate light-green lamina and an upper dark-green lamina form a typical sedimentary triplet. These light-green laminae may be variable in thickness or even absent. The transition between well-defined laminae within the triplets (from here on called intra- cycle relationships) is gradual, whereas the transition between different triplets is abrupt (from here on called inter-cycle relationships) (Fig. 6.2.B, D, F and H). 82 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 6.2.2 Laminae isotope composition In spite of the very high sampling resolution (mean=4 years, sd=1.5, n=101) δ18O diatom values display a large variability, ranging between +40.1‰ and +31.1‰ with a mean value of +37.5‰ for the whole record (sd=1.1, n=97) (Fig. 6.1). The studied interval shows three δ18O diatom major enrichment trends which 31.1 35340 40 80 120 160 36 37 38 4039 41 O (SMOW) 18 diatomGrey scale 11,990 11,530 11,600 11,700 11,800 11,900 790 788 800 810 820 830 D ep th (c m ) A ge (c al . y rs B P ) C yc le s C or e 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 1 2 ? ? ? ? ? ? ? 3 4* 5 6* 7 8 9 10 11 12 13 14 15 16 17 18 19 20* 21 22 23 24 25 26* 27 28* 29 30 31 32 33* 34 35* 36 37 38 39 40 41 42 43 44* 45 46* 47 48 49 A DB C Figure 6.1. A. Digital XRF ITRAX core scanner image from the selected and sampled interval indicating the age and its correspondent core depth. B. The 49 defined cycles composed by couplet/triplets from 102 sampled laminae. C. The smoothed grey-colour curve D. δ18O diatom values associated to each lamina. Note the diatom super-blooms are indicated by thicker white laminae and the higher values of the curve. 83 Biogeochemical processes controlling oxygen and carbon isotopes of diatom silica in lacustrine rhythmites 1 mm2 mm A D D G G H H E E F F B B C C Figure 6.2. A. Digital XRF ITRAX core scanner image of laminated sediments of core 11 corresponding to the sampled interval of Subunit 1a. Note that the lamination is composed by millimetre thick white lamina and green lamina forming rhythmites. B. Photomosaic from a thin-section showing an ideal triplet rhythmite sequence made up of (from base to top): (H) Abrupt contact between dark-green and white laminae; (G) A white lamina formed by skeletons of the large diatom Cyclostephanos andinus (> 50 μm); (F) Gradual contact between white and light-green laminae; (E) A light-green lenticular and discontinous lamina which is made up of a mixture of white and dark-green lamina; (D) Gradual contact between light- and dark- green laminae; (C) A dark-green lamina made up of diatoms embedded in an organic matter matrix. C. SEM image of dark-green lamina mainly made up by Cyclostephanos andinus (black arrows) and diatoms of the Discostella stelligera species complex (white arrows). Note the smaller Cyclostephanos andinus size (diameter < 50 μm). D. SEM image showing the decreasing upwards size of the diatoms throughout an intra-cycle contact between a light- green lamina and a dark-green lamina. Arrows indicate the different size of the diatoms. E. SEM image of a light-green lamina. The lamina is made up of complete valves and fragments of Cyclostephanos andinus valves, both showing a preferential orientation. F. SEM image showing an intracycle contact between the white and light-green laminae. Note the preferential orientation of de diatoms placed at the top of the image (light-green lamina). G. SEM image of a white lamina. The lamina is exclusively composed by large Cyclostephanos andinus (diameter > 50μm). The excellent preservation of the diatom frustules can be observed in the image (white arrow). There are no signs of dissolution. H. SEM image showing an intercycle contact between a dark-green and a white lamina. The arrows indicate the exact position of the contact which can be perfectly followed. Note the different size of the diatoms. 84 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles coincide with similar trends in the grey-colour curve (Fig. 6.3). The %C diatom values range from 0.63% in the uppermost sample (rhythmite 48) to 0.32% in the lowermost sample (rhythmite 8) (mean=0.42%, sd=0.10, n=11) whereas δ13C diatom values oscillate between –26.1‰ and –29.5‰ (mean=–28.1‰, sd=0.95, n=11). The white laminae generally display lower %C diatom and δ13C diatom values than the dark laminae from the same rhythmite, in addition there is an increase in C/Si ratios and δ13C diatom values throughout the 5 studied intra-cycle relationships (Table 6.1). δ18O diatom inter-cycle relationships have been studied in 49 cases. From these, 12 cases could not be taken into account due to the absence of δ18O diatom data or because the difference between the two consecutive isotopic values was below the mean analytical error. Valid δ18O diatom inter-cycle relationships (n=37) are characterised by higher oxygen isotope values. The most common inter-cycle relationship is the dark-green to white laminae (n=25), and it shows isotope enrichment (i.e. values increase) in 60% of the cases. Likewise, the difference between dark-green laminae to undifferentiated laminae shows similar levels of increasing δ18O diatom , whereas relationships between undifferentiated and white laminae show both increases and decreases in δ18O diatom (Table 6.2.A). 34 Gr ey sc ale Ag e ( ca l y r B P) 25 5 03 1.1 11,53011,600 11,700 11,800 11,90011,990 C o re se ctio n 41  18Odia tom B C A Figure 6.3. A. Digital XRF ITRAX core scanner image from the selected interval. B. Grey-colour surface plot elaborated from the digital image. Decadal-scale main grey-colour trends to whiter values are indicated by means of red arrows. C. δ18O diatom record. Decadal-scale main δ18O diatom trends to higher values are indicated by means of blue arrows. Note the good agreement between both proxies. 85 Biogeochemical processes controlling oxygen and carbon isotopes of diatom silica in lacustrine rhythmites Sample Cycle Colour Depth (cm) Age (cal yr BP)  18 Odiatom (SMOW)  13 Cdiatom (PDB) %Cdiatom 1 White 787.8 11,529 38.97 2 49 Dark-green 788.1 11,532 36.87 3 49 Light-green 788.5 11,537 35.63 4 49 White 788.8 11,540 38.35 5 48 Dark-green 789.1 11,543 36.27 -28,99 0,63 6 48 White 789.5 11,547 38.01 -28,51 0,47 7 47 Dark-green 789.7 11,549 37.85 8 47 White 790.2 11,554 37.93 9 46 Undifferentiated 790.6 11,559 39.67 10 45 Dark-green 791.3 11,566 38.35 11 45 White 792 11,573 37.37 12 44 Undifferentiated 792.6 11,580 37.50 13 43 Dark-green 793.4 11,588 38.07 -26,05 0,57 14 43 Light-green 794.1 11,595 37.29 -28,30 0,38 15 43 White 794.5 11,599 38.65 -28,46 0,39 16 42 Dark-green 794.9 11,604 37.36 17 42 Light-green 795.4 11,609 36.91 18 42 White 795.8 11,613 37.79 19 41 Dark-green 796.1 11,616 36.86 20 41 Light-green 796.3 11,618 38.35 21 41 Light-green 796.6 11,621 38.73 22 41 White 797 11,626 38.49 23 40 Dark-green 797.3 11,629 38.10 24 40 White 797.5 11,631 36.93 25 39 Dark-green 797.7 11,633 37.58 26 39 White 798 11,636 36.52 27 38 Dark-green 798.5 11,641 37.62 28 38 White 799 11,647 35.96 29 37 Dark-green 799.3 11,650 37.67 -27,87 0,40 30 37 White 799.6 11,653 38.16 -29,01 0,33 31 36 Dark-green 800 11,657 36.51 32 36 White 800.4 11,661 37.92 33 35 Undifferentiated 801 11,668 35.70 34 34 Dark-green 801.5 11,673 37.49 35 34 White 802 11,678 36.18 36 33 Undifferentiated 802.5 11,683 na 37 32 Dark-green 803.4 11,693 35.46 38 32 White 804 11,699 37.05 39 31 Dark-green 804.3 11,702 na 40 31 White 804.6 11,705 37.57 41 30 Dark-green 804.8 11,707 37.35 42 30 White 805.1 11,711 38.20 43 29 Dark-green 805.5 11,715 36.62 44 29 White 805.9 11,719 37.91 45 28 Undifferentiated 806.3 11,723 37.70 46 27 Dark-green 806.7 11,727 37.30 47 27 Light-green 807 11,730 37.84 48 27 White 807.3 11,734 37.72 49 26 Undifferentiated 807.7 11,738 37.71 50 25 Dark-green 808.4 11,745 na 51 25 White 809 11,751 37.30 52 24 Dark-green 809.5 11,757 37.16 53 24 Light-green 809.9 11,761 36.53 54 24 Light-green 810.3 11,765 36.50 55 24 White 810.5 11,767 37.18 Table 6.1. Analysed samples and their features (number, cycle, colour, depth, age, δ18O diatom , δ13C diatom and %C diatom . Dark grey stripes indicate not available (na) samples. 86 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles There are 51 valid (out of 62) relationships between laminae that take place within a rhythmite (intra-cycle relationships). These intra-cycle relationships are dominated by isotope depletion (values decrease). The most usual case shows changes from white to dark-green laminae (n=23) where isotope decreases occur in 67% of the cases (Table 6.2.B). Sample Cycle Colour Depth (cm) Age (cal yr BP)  18 Odiatom  13 Cdiatom %Cdiatom 56 23 Dark-green 811 11,772 37.39 57 23 White 811.7 11,780 na 58 22 Dark-green 812.2 11,785 37.16 59 22 White 812.6 11,789 37.99 60 21 Dark-green 813.4 11,798 35.86 61 21 White 814 11,804 36.98 62 20 Undifferentiated 814.5 11,809 38.70 63 19 Dark-green 814.8 11,812 35.06 64 19 White 815.2 11,816 37.19 65 18 Dark-green 815.6 11,821 38.49 66 18 Light-green 816.2 11,827 37.99 67 18 White 816.5 11,830 38.28 68 17 Dark-green 816.9 11,834 na 69 17 White 817.2 11,837 37.30 70 16 Dark-green 817.6 11,842 37.70 71 16 White 818.2 11,848 38.66 72 15 Dark-green 818.7 11,853 37.59 73 15 White 819 11,856 37.81 74 14 Dark-green 819.3 11,859 38.92 75 14 Light-green 819.6 11,862 40.13 76 14 White 820 11,867 38.34 77 13 Dark-green 820.5 11,872 38.44 -27,21 0,44 78 13 White 820.9 11,876 38.91 -29,53 0,32 79 12 Dark-green 821.2 11,879 38.02 80 12 White 821.4 11,881 37.90 81 11 Dark-green 821.8 11,886 39.24 82 11 White 822 11,888 38.66 83 10 Dark-green 822.3 11,891 37.67 84 10 White 822.8 11,896 39.24 85 9 Dark-green 823.1 11,899 37.98 86 9 White 823.4 11,902 38.00 87 8 Dark-green 824 11,909 38.62 -27,69 0,38 88 8 White 824.3 11,912 37.03 -28,89 0,32 89 7 Dark-green 824.7 11,916 37.95 90 7 White 825.3 11,922 37.57 91 6 Undifferentiated 825.6 11,925 38.31 92 5 Dark-green 826.1 11,931 37.43 93 5 White 826.5 11,935 38.54 94 4 Undifferentiated 827 11,940 37.50 95 3 Dark-green 827.4 11,945 37.57 96 3 White 827.9 11,951 38.50 97 2 Dark-green 828.3 11,956 37.38 98 2 Light-green 828.8 11,962 37.60 99 2 White 829.2 11,967 31.14 100 1 Dark-green 829.5 11,971 35.34 101 1 Light-green 830 11,977 35.45 102 1 White 830.3 11,981 37.05 87 Biogeochemical processes controlling oxygen and carbon isotopes of diatom silica in lacustrine rhythmites Table 6.1. Continued. 6.3 Discussion 6.3.1 Biological and sedimentary processes forming rhythmites White laminae features (relatively thick, good diatom preservation and monospecific diatom composition) suggest accumulation during short-term massive diatom blooms, perhaps of only days to weeks in duration. According to the chronological model rhythmites are not a product of annual variations in sediment supply, but due to some kind of multiannual processes (Hernández et al. 2008). Causes of super-blooms can be different to regular seasonal blooms which occur as part of the normal phytoplankton succession (Reynolds, 2006). We suggest that our diatom super-blooms may have been triggered by abnormally high nutrient concentrations coupled with hydrological conditions prompting diatom population growth. Low lake level stages and/or strong wind episodes would favour upwelling of nutrient- rich hypolimnion waters (Talbot and Allen, 1996). Strong mixing would also select diatoms over other types of phytoplankton due to their relative buoyancy. Alternatively, the increase in nutrient external loading due to exceptional catchment erosion during wet events could also have had the same effect (Bradbury et al. 2002). ENSO cyclicity signals recorded at this time in the Lago Chungará record (Hernández et al. in press) provide support to the existence of those two contrasting dry (El Niño) and wet (La Niña) conditions (Vuille et al. 2000; Valero-Garcés et al. 2003). Dark-green laminae (made up of a mixture of diatom valves belonging to several planktonic and benthic taxa, all embedded in an organic matter matrix) represent the baseline lake conditions where the complete phytoplankton successions over several years are preserved. These laminae therefore record the ‘normal’ intra- and inter-annual changes in the water column mixing regime characterised by the shifting species composition throughout general annual phythoplankton cycles. Preservation occurs as skeletons belonging to several diatom taxa, or simply as the organic matter mainly belonging to other algal groups (likely Chlorophycean, Cyanobacteria, etc.) that embed the valves in the dark-green laminae. Regular seasonal diatom blooms, are likely manifested in the dark-green laminae by the abundance of the small Cyclostephanos andinus (<50 μm), a large centric diatom whose buoyancy depends on the existence of a turbulent regime. Seasonal Cyclostephanos andinus (<50 μm) blooms reflected in the dark-green laminae would therefore be triggered by the same process during the super-blooms of the larger Cyclostephanos andinus (>50 μm) that make up the white laminae (i.e. water stratification breakdown). The dark-green laminae are sometimes preceded by light-green laminae. This observation indicates that recovery of the baseline conditions from the super-blooms can be more or less gradual (forming couplets or triplets, respectively). Flocculation of diatoms by extracellular polymeric substances is a common feature in the marine realm (Thornton, 2002). This phenomenon occurs towards the end of a diatom bloom, due to the onset of 88 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles Intercycle relationship types Enrichments (%) Depletions (%) n Dark-green to white laminae 60 40 25 Dark-green to undifferentiated laminae 67 33 6 Undifferentiated to white laminae 50 50 6 Total 37 Intracycle relationship types Enrichments (%) Depletions (%) n White laminae to light-green laminae 33 67 9 Light-green to light-green laminae 0 100 1 Light-green to dark-green laminae 56 44 9 White laminae to dark-green laminae 35 65 23 White laminae to dark-green laminae (non- consecutive laminae, base to top of the rhythmite) 33 67 9 Total 51 nutrient limitation. Diatom aggregation and subsequent rapid sedimentation of species having any kind of resting cell stages would favour future recruitment once nutrient resources were again available (Smetacek, 1985). Biosiliceous laminae in marine sediments have been interpreted as the product of changes in the mass sedimentation of diatoms by means of the formation of aggregates (Grimm et al. 1996, 1997). At Lago Chungará a similar phenomenon could have taken place in the formation of the light-green laminae once the super-blooms of the large (>50 μm) Cyclostephanos andinus come to an end. Aggregation of cells enclosed in a gelatinous matrix could therefore have taken place, being rapidly deposited in the form of the transitional light-green laminae. Although the life cycle details of Cyclostephanos are far from fully known, the closely related genera Stephanodiscus, to which Cyclostephanos once belonged (Round et al. 1990), is known to produce resting cells (Sicko-Goad et al. 1989), whose aggregation and rapid sedimentation represents a transition to a resting phase (Smetacek, 1985; Alldredge et al. 1995). It is therefore likely that the mechanism of formation of triplets is mediated by processes of self-sedimentation triggered by Cyclostephanos andinus (Grimm et al. 1997). 6.3.2 δ18O diatom and δ13C diatom interpretation Variation in δ18O diatom can result from a variety of processes, such as δ18O lakewater , temperature, vital effects and post depositional diagenesis (Leng and Barker, 2006). In hydrologically closed lakes under arid climate conditions evaporative concentration processes have a much larger effect on δ18O lakewater Table 6.2. A. Intercycle isotope relationships between the defined rhythmites. B. Intracycle isotope relationships between the defined rhythmites. Relationship types are established according to the colour of the laminae that are in contact. 89 Biogeochemical processes controlling oxygen and carbon isotopes of diatom silica in lacustrine rhythmites than any other process (Gasse and Fontes, 1992; Leng and Marshall, 2004; Hernández et al, in press). In these circumstances, the δ18O diatom record can be used as an indicator of changes in the P/E related to climate change (Leng and Barker, 2006). At present, Lago Chungará can be considered a closed lake due to its water residence time (ca 15 years), and the fact that δ18O lakewater is enriched by 14‰ relative to δ18O of the inputs (precipitation, springs and river) (Herrera et al. 2006). This was probably also the case in the Late Glacial-Early Holocene transition described here because δ18O diatom values are similar (around +37.5‰) to other diatom-isotope sequences in tropical sites (e.g. Lakes from Mount Kenya (Kenya), Barker et al. 2001; Lake Malawi (Malawi, Mozambique, Tanzania), Barker et al. 2007; Lake Tilo (Ethiopia); Lamb et al. 2005). Thereby the variations in the δ18O diatom from Lago Chungará sediments must be mainly derived from changes in the δ18O lakewater resulting from shifts in the balance between P/E, rather than other factors. The organic matter enclosed within diatom frustules contains polysaccharides, proteins and long- chain polyamines (Kröger and Poulsen, 2008). These substances host carbon which is protected from post-depositional diagenetic alteration (Des Combes et al. 2008). As these carbon compounds will be synthesised from the surrounding waters, isotope analysis of the carbon contained in the diatom frustules can be used as a proxy for reconstructing the lake’s carbon cycle. Previously published studies suggest primary productivity and CO 2(aq) concentration as the main factors which determine δ13C diatom in marine environments (Schneider-Mor et al. 2005), although lake δ13C diatom is likely controlled by more complex environmental conditions (Hurrell et al. submitted). δ13C diatom variations due to the species effect, cell size, growth rate or/and metabolic pathway are neglected here since in our case the δ13C diatom analysis was always carried out on similar sized-cells (38-62 μm) and on the same diatom species (Cyclostephanos andinus). In lakes, it is usually assumed that the carbon pool in the water becomes enriched in 13C during the periods of enhanced productivity (Leng et al. 2005b; Singer and Shemesh, 1995) since phytoplankton preferentially use the lighter isotope. However, the maximum productivity events found here, associated with the white laminae (short-term diatom super-blooms), show the lowest δ13C diatom values. Therefore, although the diatom blooms will have preferentially incorporated 12C, this cannot have been sufficient to positively shift the isotope value of the dissolved carbon. Instead, the supply of carbon available to the diatoms must have been sufficient not to lead to limiting conditions. The δ13C bulk in the Lago Chungará laminated unit range from –21‰ to –19‰ (J.J. Pueyo, unpublished data), yielding a difference of more than 5‰ when compared to the measured δ13C diatom values. Nevertheless, the C/N ratio from bulk sediments of the laminated unit have values ranging between 7 and 11 (J.J. Pueyo, unpublished data), indicating that the δ13C bulk signal would have a mainly algal origin (Meyers and Terranes, 2001). For this reason, it seems that the δ13C diatom , rather than being mainly affected by changes in the source of organic matter, is mostly conditioned by changes in dissolved carbon concentration. Mineralisation of terrestrial or previously deposited carbon through microbial 90 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles decomposition could create a pool of isotopically lighter carbon available to the diatoms. Release of this through respiration (CO 2(aq) and possible also CH 4 ) would be partly controlled by lake dynamics under the control of external forcing factors. Lake water dynamics are mainly governed by two contrasting situations (Hernández et al. 2008): (1) a water column subjected to episodes of very strong mixing, which is represented by the white laminae, and (2) a more stable condition, including periods of lake water stratification, and represented by the dark-green laminae. During the stratified periods, concentrations of oxygen and other electron acceptors typically decrease in the hypolimnion, while CO 2(aq) , CH 4 , and nutrients accumulate (Bedard and Knowles, 1991). These dissolved nutrients, as well as the accumulated CO 2(aq) and CH 4 , are released into the entire lake during mixis (Houser et al., 2003), when the diatom blooms occur, being the carbon incorporated into the diatom frustules. 6.3.3 δ18O diatom inter-cycle relationships (white laminae formation) The characterisation of δ18O diatom values through the inter-cycle relationships gives clues to understand the underlying processes involved in the formation of the white laminae. The massive diatom blooms that produce the white laminae have to be triggered by an exceptional injection of nutrients into the water column which may or may not be associated with a water mass change. The start of the rhythmite is usually accompanied by δ18O diatom enrichment (Table 6.2.A), indicating a decrease in the P/E ratio, a drop in the lake water level and a remobilization of nutrients from the hypolimnion as the more likely scenario during the white laminae formation (Fig. 6.4.A and B, transition 1). Episodes of diatom super-blooms occur throughout the whole studied section, but their formation is a time scale-dependent process. At decadal-centennial scales white laminae are more marked (higher values in the grey colour curve) and thicker (around 6 mm) with higher isotope oxygen values (up to 39.2‰) than during other laminae deposition periods (Hernandez et al. in press). Deposition of these white laminae stretches are related to low-stand conditions, as shown in the uppermost part of the three shallowing upwards trends observed in the δ18O diatom record (Fig. 6.3). However, at interannual scales the inter-cycle isotope relationships reveal that both changes to drier or wetter conditions may trigger the formation of the white laminae, but falls in lake level were more likely responsible for the development of the massive diatom blooms (Table 6.2). 6.3.4 δ18O diatom and δ13C diatom intra-cycle relationships (green laminae formation) The intra-cycle relationship between δ18O diatom and δ13C diatom provides a means of better understanding of the environmental processes involved in the origin of the green laminae. The δ18O diatom 91 Biogeochemical processes controlling oxygen and carbon isotopes of diatom silica in lacustrine rhythmites intra-cycle relationships show that transitions from white to dark-green laminae are mainly governed by δ18O diatom depletions by up to −2.7‰ (65%; n=23), although there are also a significant percentage of enrichments (Table 6.2). As in the case of the white laminae formation, green laminae can be formed under both lake-water level drops and rises, but their formation is clearly favoured by increasing P/E ratios with subsequent lake-water level rises (Fig. 6.4.A and C, transitions 2 and 3). Green laminae record the baseline conditions in the water column mixing regime including water table stratification periods. The intra-cycle relationships which show δ18O diatom depletions indicate that the lake tended to progressively recover the previous environmental conditions by means of a gradual increase in water availability (Fig. 6.4.A and C, transition 2 and 3). Conversely, the intra-cycle relationships which show δ18O diatom enrichments would indicate the recovery to a lower lake level after a super-bloom caused by a massive allochthonous nutrient input associated to enhanced rainfall. This is 30 diatom size nutrients 50 90 m - + Cyclostephanos andinus Cyclotella stelligera complex benthic or tychoplanktonic diatom Sedimentary facies (W=white laminae, L=light-green laminae, D=dark-green laminae) Rhythmite and sedimentary facies transitions organic-rich mud Anoxic water Baseline conditions Mixing D Extraordinary bloom conditions W Nutrient resuspension Transitional conditions Weak Mixing L Anoxic water Baseline conditions Mixing D Extraordinary bloom conditions Strong Mixing W Water level drop Increasing O 18 diatom Water level rise Decreasing Increasing   18 Odiatom diatom 13 C Water level rise Decreasing Increasing C   18 Odiatom diatom 13 Water level drop Increasing Increasing   18 13 O C diatom diatom D D D L W W 43 2 1 1 4 3 3 2 Nutrient resuspension Strong Mixing A B C Figure 6.4. A. Rhythmite log succession showing facies and transitions (indicated by letters and numbers, respectively). B. The most frequent intercycle relationship scenarios. Transition case 1: From dark-green to white laminae, the white laminae formation (diatom super-blooms) is more often favoured by drops of the lake water level (increases in δ18O diatom values) and therefore related to recycled nutrients from the hypolimnion. C. The most common intracycle relationship scenarios. Transition case 2: From white to dark-green laminae, the dark-green lamina formation is usually favoured by rises of the lake level water (decreases in δ18O diatom values). Transition case 3: From white to light-green laminae, the light-green laminae formation is usually favoured by rises of the lake water level (lower δ18O diatom values). Transition case 4: From light-green to dark green laminae, the dark-green laminae formation is almost indistinctly favoured by drops or rises of the lake water level, with a slight predominance of the former as the δ18O diatom show. 92 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles suggested by the prevalence of δ18O diatom depletions that precede those super-blooms (90%; n=10) (Table 6.1). This model suggests that the green laminae occurred most of the time as a result of the recovery phase favoured by lake water rises. Finally, when the lake is already in the recovery phase (transitional and baseline conditions) it may evolve, indistinctly, towards rise or fall lake water level stands, as indicated by the light- to dark-green isotope transitions (enrichments= 56%; n=9) (Fig. 6.4.A and C, transition 4). Comparison between δ13C diatom and δ18O diatom in the intra-cycle relationships show that the former can be associated to either δ18O diatom enrichments or depletions. However, δ13C diatom values show that all intra-cycle relationships yield carbon isotope enrichment during the formation of the organic-rich green laminae (Table 6.1). This occurs because during the white laminae formation, the strong mixing necessary for the formation of the massive diatom blooms break the lake water stratification. Transport of CO 2(aq) and CH 4 from an hypolimnion enriched in these compounds is then allowed, depleting the δ13C of the total carbon pool. 6.3.5 Climate forcing of the laminae formation The biogeochemical reconstruction presented above suggests that laminae are formed by the occurrence of diatom super-blooms. These are directly affected by nutrient availability which, in turn, is mainly controlled by the lake level fluctuations and mixing as stable isotopes (δ13C diatom and δ18O diatom ) demonstrate. Hence, the laminae formation in Lago Chungará seems to be mainly induced by environmental forcing such as short-term climate variability. ENSO and solar activity, as well as interactions between both phenomena, have been key factors prompting changes in the atmospheric conditions over the Altiplano region during the Late Glacial to Early Holocene at decadal and longer term time scales (Hernandez et al. in press). ENSO and solar activity, as responsible for the more accentuated sub-millennial wet or dry conditions over the Andean Altiplano (Theissen et al. 2008), are very likely the main environmental factors in the frequency and production of the white laminae. In the Central Andes, there is a weak trend towards wet conditions during La Niña phase and to dry conditions during El Niño phase (Valero-Garcés et al. 2003). According to our depositional model, white laminae formation could be triggered by both phases. However, El Niño events seem most likely to be responsible, since the white laminae formation is usually favoured by changes from wet-to-dry conditions, which is seen in the isotope enrichments. It is important to point out that white laminae are more intense (higher intensity grey colour values) and better developed (thicker) during periods showing higher isotope values which point to their formation during drier conditions (Hernández et al. in press). Thus, the diatom super-blooms likely occurred during El Niño- like periods. The presence of exceptionally intense and thick white laminae could, on the other hand, be indicative of the overlapping of both ENSO and solar activity phenomena during such periods. Isotope data show that high intensity of ENSO and solar activities can be recorded beyond the white laminae deposition. 93 Biogeochemical processes controlling oxygen and carbon isotopes of diatom silica in lacustrine rhythmites The short duration (days) of extreme blooms in relation to lake water residence time gives an isotope signature that will remain for longer periods (years) being recorded in the dark laminae (Hernández et al. in press). 6.4 Conclusions Lago Chungará rhythmites record multiannual diatom super-blooms lasting from days to weeks (white laminae) and the lake hydrology recovery towards the baseline conditions throughout several years (dark-green laminae). Self-sedimentation phenomena taking place immediately after the diatom super-blooms cannot be discarded as a sign of the end of the super-bloom (light-green laminae). The diatom super-blooms are favoured episodes of extreme turbulent conditions affecting the whole water column, and/or by strong runoff during wet episodes. In the first case upwelling from nutrient-rich hypolimnion waters allowed an extraordinary nutrient availability, whereas in the second case allochthonous nutrient enrichment would be implicated. In Lago Chungará, the δ18O diatom record can be used as an indicator of changes in the P/E related to climate changes, whereas the δ13C diatom variability would be mainly influenced by changes in CO 2(aq) concentration. δ18O diatom values show that both white and green laminae formation may occur in either dry or wet conditions, but the diatom super-blooms were more intense (thicker white laminae) during decadal- centennial lowstands. δ18O diatom composition shows that the white laminae formation was mainly favoured by low lake levels, whereas the green laminae formation was especially prompted by lake level rises. ENSO and solar activity are the most likely main climate forcing mechanisms triggering the white laminae formation. Both El Niño and La Niña phases could be responsible for this, but geochemical data indicate that dry conditions associated to El Niño could have the primary role since the white laminae formation was usually favoured by changes from wet-to-dry conditions in the Altiplano region. Moreover, the periods where the white laminae present major thickness and whiter colours might be indicative of phases with overlapping El Niño and solar activity. On the contrary, green-laminae were deposited during the baseline climate phases, when the normal plankton succession throughout several years and associated regular diatom blooms occur. High resolution isotope analysis of the oxygen and carbon isotopes in diatom silica in this uniquely laminated sequence has displayed links between limnology, catchment runoff variations, hydrology and climate forcing at different time scales. Strong El Niño phases have triggered nutrient and carbon release from the hypolimnion and sediments that has led to diatom super-blooms. Such phenomena may be found in many lakes but few preserve evidence in their sedimentary architecture. Further work on other parts of this record and in similarly laminated sites may reveal the full impact of these multi-annual events on lake ecosystems and biogeochemical cycles. 94 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles Chapter 7 Oxygen and carbon diatom isotope records from the Lago Chungará laminated unit (12,400 to 8,400 cal years BP) 7.1 Introduction A sound grasp of the tropical circulation and its temporal evolution is essential for characterising moisture transport mechanisms towards extratropical areas. Long-term response and climate variability of tropical systems is therefore crucial for understanding future climate change (e.g. Barker et al. 2007; Giralt et al. 2008; Chiang, 2009). However, there are currently few data on the frequency and amplitude of abrupt climate changes in tropical latitudes of South America, and the forcings responsible for this climate variability at different temporal scales are poorly documented (e.g. Moreno et al. 2007; Hyllier et al. 2009). Although it is believed that orbital parameters are the most important drivers of tropical precipitation at millennial-time scales (Baker et al. 2001a; Clement et al. 2001; Plazeck et al. 2006), a growing body of evidence suggests that climate variability is also governed by other centennial- to millennial-scale forcings such as the ENSO (Rowe et al. 2002; Baker, 2002). Despite the fact that many studies have focused on the period from the LGM until the present, a number of questions such as whether or not the Younger Dryas-like conditions arose in the Central Andes (Clapperton et al. 1997; Rodbell and Seltzer, 2000; Lowell and Kelly, 2008) or whether or not the ENSO weakening in the Early Holocene had a major effect on the lake levels (Rodbell et al.1999; Moy et al. 2002; Hernández et al. in press) remain unresolved. One way to answer these questions could be to undertake multiproxy studies of high-resolution lacustrine sequences. Recent palaeoenvironmental studies conducted at Lago Chungará have shown a clear response of this lake to regional climate changes, evidenced by shifts in the lake water level (Valero-Garcés et al. 2003; Moreno et al. 2007; Sáez et al. 2007; Giralt et al. 2008). The laminated unit of the Lago Chungará sedimentary infill is made up almost exclusively of diatom frustules that are exceptionally well preserved. Such a sedimentary record enables us to apply the well-established technique of δ18O diatom (e.g. Shemesh et al. 2001; Leng and Barker, 2006) as well as the recent and much less developed technique of analysis 95 of δ13C diatom (Hernández et al. submitted; Hurrell et al. submitted). The combination of both δ18O diatom and δ13C diatom proxies provides valuable insights into the nutrient input and carbon cycle during the Late Glacial-Early Holocene transition. The consumption and production of CO 2 and CH 4 by microorganisms exert an influence both on the concentration of these gases and on the atmospheric heat budget (Tranvik et al. 2009). Lakes play an important role in the global carbon cycle, potentially affecting regional climate and global change (Cole et al. 2007). Changes in the lacustrine carbon cycle may be revealed by δ13C diatom from lacustrine sediments, as has been reported in marine environments (e.g. Rosenthal et al. 2000; Crosta and Shemesh, 2002; Schneider-Mor et al. 2005). Diatom frustules contain polysaccharides and proteins (pleuralins, silaffins and long chain polyamines) enclosed within the silica cell wall structure (Kroger and Polusen, 2008). This organic matter is protected by silica against diagenetic processes that might affect the sediments (Des Combes et al. 2008). Therefore, it may reflect the features of dissolved carbon in water during cell formation. However, the δ13C diatom has not been widely applied to lake sediments where the complexity of sedimentary conditions such as the existence of different carbon sources and highly dynamic aquatic food webs raise further questions for interpretation (Hurrell et al. submitted). In this study, we present records of δ18O diatom and δ13C diatom from the Lago Chungará diatom-rich laminated sedimentary unit which spans from 12,400 to 8,400 cal years BP. These records provide evidence of significant palaeoenvironmental changes at local and regional scales in the Central Andes over this time window. 7.2 Results Isotopic values of the δ18O diatom record range from 34.7 to 40.3‰. This proxy displays a long-term enrichment trend (~+3‰) between 12,400 and 10,100 cal years BP. This trend is however punctuated by short-term depletions centered at ca 11,500, 11,200 11,000, 10,700, 10,400 and 10,200 cal years BP. Subsequently, a period with maximum isotope values (ca +40‰) lasts for approximately one thousand years. Finally, a progressive depletion trend (–3‰) characterises the last 800 years of the record (Fig. 7.1). The δ13C diatom values range from -30.3‰ (12,400 cal years BP) to -22.6‰ (10,100 cal years BP). The record shows considerable fluctuation with excursions to lower values that match the δ18O diatom record. Changes in δ13C diatom reveal a pattern that is similar to that of δ18O diatom from 12,400 to 10,100 cal years BP (Fig. 7.1). The following millennium is characterised by a depletion of δ13C diatom whereas the δ18O diatom values remain high. The uppermost part of the record displays a δ13C diatom enrichment trend that is particularly pronounced in the last 200 years. This trend is however interrupted by a large depletion reversal (–5‰) at 9,200 cal years BP. 96 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 440 460 480 500 520 540 560 580 600 620 640 660 680 700 720 740 760 780 800 820 840 L a te G la c ia l e a rl y H o lo c e n e 860 8,400 9,000 9,600 10,200 10,800 11,400 12,000 1 1 2 2 3 3-4 5-6 4 5 6  18 O ( SMOW) diatom ‰ C om po si te co re de pt h (c m ) F ac ie s as so ci at io n A ge (c al . Y ea rs B P )  13 C ( VPDB) diatom ‰ I I ? 34 35 36 37 38 39 40 41 -32 -30 -28 -26 -24 -22 Dark green-beige laminated diatomite Green-White laminated diatomite Dark green-Light green laminated diatomite Dark green, massive and banded organic-rich diatomite Brown-White interbedding and carbonate-bearing laminated diatomite Brown-laminated diatomite Carbonate layers Grey-to-black volcanoclastic layers (tephras) Figure 7.1. Facies association, δ18O diatom and δ13C diatom records for the period 12,400–8,400 cal years BP from Lago Chungará. Arabic numerals indicate the main humid events as revealed by the δ18O diatom record and the related δ13C diatom depletions during the Late Glacial-Early Holocene transition. The Roman numeral indicates the maximum dry phase associated with the highest δ13C diatom values. Arrows show the main trends during the established climate stages: Late Glacial-Early Holocene humid phase (12,400-10,100 cal years BP), Early Holocene dry phase (10,100-9,100 cal years BP) and Early Holocene dry-to-wet transition phase (9,100-8,400 cal years BP). 97 Oxygen and carbon diatom isotope records from the Lago Chungará laminated unit (12,400 to 8,400 cal years BP) 7.3 Discussion 7.3.1 Controlling factors in δ18O diatom and δ13C diatom Diatoms are autotrophic organisms containing a shell or frustule made up of biogenic silica (SiO 2 ·nH 2 O) and various specific organic macromolecules such as proteins and polysaccharides (Round et al. 1990; Kröger and Poulsen, 2008). Diatoms, like plants, require CO 2 for photosynthesis. However, they also have the ability to utilise HCO 3 - through biophysical concentrating mechanisms (Johnston et al. 2001). Thus, the interpretation of changes in the stable isotope composition of diatom silica is not straightforward since a number of influential factors cannot be ruled out. A good grasp of the processes that govern the δ18O diatom signal is therefore required to interpret the past oscillations of this proxy (Leng et al. 2005b). δ18O diatom depends on both the water temperature and the isotopic composition of the lake water when the shell is secreted (Shemesh et al. 1992). The δ18O diatom record may thus result from one or more of the following factors: a) changes in the P/E (e.g. Barker et al. 2007), b) oscillations in water temperature (e.g. Hu and Shemesh, 2003), c) fluctuations in δ18O or in temperature at the source of the precipitation (e.g. Swann et al. 2010), and/or d) variations in the δ18O lakewater attributed to changes in the contribution of precipitation from different air masses (e.g. Rosqvist et al. 2004). The interpretation of δ18O diatom may be complicated by changes in the specific composition of diatom assemblages, diatom dissolution, and silica maturation (Swann et al. 2010, Jonsson et al. 2010). Despite the fact that little is known about these factors, our analyses, which were performed in an exceptionally well preserved single species (Cyclostephanos andinus) with no signs of diagenetic alteration, show that these factors exert little influence on the δ18O diatom record of Lago Chungará (Fig. 5.4). In closed lakes, oscillations in δ18O lakewater will usually be far greater than fluctuations due to temperature or isotope changes in rainfall (Leng and Barker, 2006). Oxygen isotope records in tropical regions have also proved to be sensitive to the amount of precipitation (Cole et al. 1999), and it is likely that this relationship is important in tropical lake records (Barker et al. 2001; Hernández et al. in press). Under these circumstances, the δ18O diatom record can be used as an indicator of change in the P/E related to climate change (Leng and Marshall, 2004). A modern day calibration was carried out to determine the environmental forcings governing changes in δ18O diatom from Lago Chungará (Herrera et al. 2006; Table 2.1). At present, evaporation enriches the lake by ~14‰ with respect to precipitation. It may be assumed therefore that significant changes in water isotope values in the past mainly resulted from shifts in the P/E. Although the P/E is usually governed by climate oscillations, non-climate effects such as the ontogeny of the lake can also exert an influence on the δ18O signal. Changes in the residence time of the lake caused by variations in basin hydrology or in groundwater fluxes will also influence the degree of enrichment (Leng et al. 2005b; Hernández et al. 2008). During transgressive periods, the stepped morphology of Lago Chungará forced the expansion of the lake towards the shallow margins (Fig. 4.4), resulting in an increase in the lake surface/volume ratio that enhanced evaporation. This caused a background isotope enrichment during the Late Glacial-Early Holocene transition. Nevertheless, oscillations in the P/E seem to be mainly responsible for changes in the δ18O diatom in Lago Chungará (Hernández et al. 2008). As photosynthesisers, diatoms offer a more reliable substrate for reconstructing past changes in δ13C than bulk sediment organic matter, which often includes a mixture of both autotrophs and heterotrophs (Rosenthal et al. 2000). In addition, organic carbon from diatom frustules is less likely to be affected by diagenesis (Des Combes et al. 2008). However, most of the information about how δ13C diatom 98 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles isotopes records are interpreted is mainly derived from the marine realm (e.g. Schneider-Mor et al. 2005; Singer and Shemesh, 1995). Productivity plays a major role in determining the isotopic signature of oceanic DIC because reduced phytoplankton productivity increases the availability of 12C with the result that diatoms yield lower δ18O diatom values (Singer and Shemesh, 1995; Schneider-Mor et al. 2005). Consequently, δ18O diatom from marine records has commonly been used as a proxy for primary production and CO 2(aq) concentration (Des Combes et al., 2008). Nevertheless, there are other factors that influence δ13C diatom , such as taxonomic composition, pH, cell growth rate, size, shape and CO 2(aq) permeability and biochemical metabolic pathways (Popp et al. 1998, Hurrell, 2009). These obstacles could be overcome if a previous physical selection of single-size diatom species is undertaken (e.g. Singer and Shemesh, 1995; Hernandez et al. submitted). In addition to the factors previously mentioned as major influences, the scant literature on lacustrine sediments highlight other factors given that δ13C diatom in lakes is controlled by more complex environmental conditions (Hurrell et al. submitted). The larger catchment carbon pool of lakes in relation to their water mass exerts a considerable influence, whereas the significance of inputs of terrestrial organic matter to the ocean basins can be considered low. δ13C diatom from lakes can thus indicate changes in the nature and concentration of lake carbon apart from productivity oscillations. Therefore, the δ13C diatom signal is not only a function of in-lake processes but also of the catchment characteristics that influence carbon (Hurrell et al. submitted). External loadings would increase the carbon inputs to the lake of terrestrial isotopically depleted carbon-rich materials which would become an important source of carbon and would therefore produce lower δ13C diatom signals (Hurrell et al. submitted). Isotope values from Lago Chungará terrestrial plants range between –26 and –23‰, whereas planktonic algae values are higher than –15‰ (Pueyo JJ, unpublished data). Thus, even small external loadings could produce significant excursions in δ13C diatom values. Climate drives soil and vegetation dynamics in the watershed as well as stratification and mixing patterns in lakes. Thus, the transport of organic matter from catchments to lakes and its subsequent CO2 (aq) concentration are ultimately controlled by climate (Catalan and Fee, 1994; Catalan et al. 2009). δ13C diatom depletions in Lago Chungará are consistent with humid events in Lago Chungará as indicated by the δ18O diatom record. External loadings favoured by these humid periods introduce isotopically depleted carbon from the terrestrial organic matter in the lake water. On the other hand, periods of enhanced mixing of the water column also prompt influx of CO 2 enriched waters from the hypolimnion to the photic zone, causing isotopic depletion in DIC (Houser et al. 2003). The δ13C diatom in the Lago Chungará sedimentary record seems to corroborate this explanation. Major trends in δ13C diatom are in agreement with a rough indicator of water column mixing derived from the multivariate statistical analysis of the diatom assemblages (Bao et al. in prep) (Fig. 7.2). 99 Oxygen and carbon diatom isotope records from the Lago Chungará laminated unit (12,400 to 8,400 cal years BP) Another concordance is found between the δ13C diatom record and the BSi flux to the sediments (Bao et al. in prep), an indicator of past biosiliceous productivity conditions (Conley, 2001) (Fig. 7.2). All these observations suggest that the δ13C diatom signature in the sediments are due to changes in the carbon input from the catchment, degree of mixing, biosiliceous productivity or to a combination of some of these factors. 7.3.2 Palaeoenvironmental reconstructions The palaeonvironmental reconstruction of Lago Chungará during the Late Glacial and Early Holocene is based on a) the δ18O diatom record as an indicator of changes in the P/E and hence of the moisture balance, and b) the δ13C diatom record, which reveals changes in productivity, carbon inputs from the catchment and in CO 2(aq) concentration. Late Glacial-Early Holocene transition humid phase (12,400-10,100 cal years BP) The Late Glacial-Early Holocene transition in the Andean Altiplano was a humid period (e.g. Wolfe et al. 2001; Giralt et al. 2009) coeval with the Coipasa humid phase found elsewhere (Sylvestre et al. 1999; Placzek et al. 2006). The δ18O diatom record in Lago Chungará however displays a slight enrichment trend during the first half of this period (12,400-11,000 cal years BP) and a higher enrichment trend during the second half (11,000-10,100 cal years BP), which would suggest a lake level decline (Fig. 7.1). Contrary to expectations, this increasing trend can be ascribed to the flooding of the shallow eastern and southern margins after ca 11,000 cal years BP (Sáez et al. 2007). The reason for this apparent contradiction is that the surface area to volume ratio of the lake significantly increased because these flooded margins were much shallower than the central plain (Fig. 2.7). As a result the relative importance of evaporation was enhanced. This induced an unexpected enrichment trend of δ18O diatom in a transgressive phase (Hernández et al. 2008). Therefore, the six major isotope depletions of the δ18O diatom curve at this time would be the real indicators of the overall humid character of this period. The magnitude of these six δ18O diatom depletions decreased from 10,500 to 10,100 cal years BP when the surface/volume ratio was probably at its maximum prior to a gradual shift towards arid conditions. There is a general consensus that millennial-scale shifts in moisture conditions in tropical South America are orbitally induced by changes in solar insolation (e.g. Baker et al. 2001b; Clement et al. 2001) although the ENSO long-term changes also play a major role (e.g. Rowe et al. 2002; Abbott et al. 2003). The Late Glacial-Early Holocene transition is a period where the summer insolation decreases towards an insolation minimum at 10,000 cal years BP (Berger and Loutre, 1991). This insolation decrease 100 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles induces a diminution in the convective activity with clearer skies and less rainfall (Rowe et al. 2002; Hillyer et al. 2009), triggering a reduction in the P/E and lake level falls. However, the evidence for a highstand in Lago Chungará during this period suggests a humid phase. The most plausible explanation for this apparent contradiction would be the establishment of long-term La Niña-like conditions in the tropical Pacific (Betancourt et al. 2000; Koutavas et al. 2002), giving rise to a humid phase over the Altiplano during the Early Holocene (Hernández et al. in press). In addition, a regional effective moisture reconstruction was previously performed using the Lago Chungará sedimentary record (Giralt et al. 2008). This reconstruction was carried out by applying multivariate statistical analyses (Cluster, RDA and PCA) to magnetic susceptibility, XRF, XRD, TC, TOC, BSi and grey-colour curve of the sediment data. For the lower part of Chungará sequence (Unit 1), the effective moisture availability proxy depends on the P/E, with positive values corresponding to drier conditions (Giralt et al. 2008) (Fig. 7.3). The data obtained from this reconstruction were filtered for the ENSO (0.19-0.27 Hz) and solar activity cycles (0.075-0.11 Hz) frequency bands (Fig. 7.3) (Giralt et al. in prep). The moisture balance reconstruction shows a good agreement with a humid period between 11,000 and 10,500 cal years BP, and with a transition towards drier conditions for the next ca 500 years. Moreover, the ENSO intensity displays the highest values for this period, indicating an enhanced ENSO activity. This activity was mainly 460 500 540 580 620 660 700 740 780 820 L G e a rl y H o lo c e n e 860 8,400 9,600 9,000 10,800 10,200 12,000 11,400 0 -32 34 41 -22 32 - + BSi MAR (mg /cm yr) 2 Bao et al., In prepPresent work Berger and Loutre, 2001 Mixing  18 O ( SMOW) diatom ‰ C om po si te co re de pt h (c m )A ge (c al . y ea rs B P )  13 C ( VPDB) diatom ‰ Insolation 18ºS Dec (Wm ) -2 485450 Figure 7.2. Comparison between δ18O diatom and δ13C diatom records obtained in this work with the data of Biogenic Silica (BSi) mass accumulation rates in the sediments and the mixing intensity in the lake water column as deduced from the multivariate analysis of diatom assemblages (Bao et al. in prep). The insolation curve in the austral summer at 18ºS for this period is also shown. 101 Oxygen and carbon diatom isotope records from the Lago Chungará laminated unit (12,400 to 8,400 cal years BP) governed by the La Niña conditions, which are characterised by a humid situation over the Andean Altiplano (Valero-Garcés et al. 2003) in accordance with the δ18O diatom values. δ13C diatom data are consitent with this interpretation. δ13C diatom values also show an increasing trend with depletions that generally match the δ18O diatom drops (Fig. 7.1). The long-term increasing upwards trend agrees with the increase in productivity as revealed by the BSi flux to the sediments, suggesting a depletion of the available 12C and hence higher values of δ13C diatom . On the other hand, the coincidence of negative excursions in δ13C diatom with δ18O diatom drops suggests that increased organic matter transported by runoff could have added light carbon to the carbon pool, resulting in lower isotope values. There is no clear evidence of a Younger Dryas-like event between 12,400 and 11,000 cal years BP in Lago Chungará, which is consistent with other tropical Andean records, e.g. Lakes Pacucha (Hillyer et al. 2009), Titicaca (Paduano et al. 2003) and Junin (Seltzer et al. 2000). Early Holocene dry phase (10,100-9,100 cal years BP) The driest period recorded during the Early Holocene corresponds to the highest values of δ18O diatom (Fig. 7.1). The reduction in the relative abundance of planktonic diatoms (Sáez et al. 2007; Bao et al. in prep) and XRF data (Moreno et al. 2007; Giralt et al. 2008) indicate a fall in the water availability in Lago Chungará (Fig. 7.3). The brown-white interbedding and carbonate-bearing laminated diatomite facies (Fig. 7.1) are also indicative of low lake levels and hence of a reduction in the moisture of the region (Sáez et al. 2007). Although this dry event seems to follow a heterogeneous spatial and temporal pattern, it appears in most of the records from the Andean Altiplano (e.g. Seltzer et al. 2000; Baker et al. 2001a; Abbott et al. 2003; Hyllier et al. 2009). This event coincides with the minimum values of the summer insolation precessional curve, favouring drier conditions and the lowest lake levels throughout the Holocene (Rigsby et al. 2005; Hyllier et al. 2009). During this insolation minimum, the ITCZ experienced a northward shift with a reduction in the moisture transport and in the precipitation over the South American tropics (McGregor and Nieuwolt, 1998). Moreover, the ENSO and solar activity intensities show the minimum values for this period (Fig. 7.3). When the ENSO forcing is weak, small northward shifts of the ITCZ or the establishment of a tropical North Atlantic warm pool can induce periods of drought in Amazonia (Baker, 2002; Marengo, 2004; Marengo, 2007; Vuille et al., 2000). Thus, the Andean Altiplano region would fall in the most arid period of the Holocene (Hyllier et al. 2009). A thick stretch of mostly white laminae was deposited during this lowstand (Fig. 7.1). These white laminae accumulated during massive short-term diatom blooms and were triggered by high nutrient 102 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles availability as a result of strong mixing periods (Hernández et al. submitted). Although lake productivity decreases in this period the δ13C diatom record displays high values followed by a sudden depletion (Fig. 7.1). This indicates either an increase in terrestrial organic matter inputs or in dissolved CO 2(aq) concentration from the hypolimnion. Reduced loadings to the lake, which are associated with dry conditions, make the latter hypothesis the most plausible explanation. Enhanced mixing conditions governing this period (Bao et al. in prep.) (Fig. 7.2) would have released the CO 2(aq) available in the hypolimnion, depleting the carbon pool and the δ13C diatom values (Hernandez et al. submitted). Bacterial induced methanogenesis providing CH 4 to the lake water must not be ruled out (J.J. Pueyo, pers comm.). However, the decreasing trend of δ13C diatom after this maximum could be ascribed to a depletion of the available CO 2(aq) and to the subsequent depletion of 12C in the carbon pool. 103 Oxygen and carbon diatom isotope records from the Lago Chungará laminated unit (12,400 to 8,400 cal years BP) Effective moisture balance Age (cal. Years BP) 8,000 8,500 9,000 9,500 10,000 10,500 11,000 11,500 12,000 12,500 Solar activity intensity ENSO intensity Wet Dry High HighLow Figure 7.3. A) Effective moisture balance derived from the second eigenvector of PCA from magnetic susceptibility, XRF, XRD, TC and TOC, BSi (Giralt et al. 2008) for the laminated unit of Lago Chungará. The raw data are shown in grey and the smoothed data in yellow. B) Intensity of ENSO and solar activity signals (red and green, respectively) from the effective moisture balance curve filtered at 0.19-0.27 Hz and 0.075-0.11 Hz frequency bands (Giralt et al. in prep). Early Holocene dry-to-wet transition phase (9,100-8,400 cal years BP) A transition towards moister conditions is recorded by δ18O diatom between 9,100 and 8,400 cal years BP. This proxy displays a slight decreasing upwards trend to the top of the unit. These moister conditions are consistent with an increase in the summer insolation, and with the effective moisture reconstruction (Giralt et al. 2008) that indicates a shift to more humid conditions between 9,000 and 8,500 cal years BP (Fig. 7.3). Moreover, this humid phase is well known elsewhere over the Andean Altiplano (Baker et al. 2001a; Grosjean et al. 2001). On the other hand, δ13C diatom values display a slight enrichment trend in this period. This indicates that the δ13C diatom values are more influenced by lake productivity (which also increases) than by the terrestrial organic matter inputs from an increased runoff. The transition to more humid conditions from a severe drought situation, when vegetation coverage of the catchment was scarce, reduced external organic matter inputs to the lake. This would have limited the carbon pool, resulting in the subsequent δ13C diatom enrichment during this enhanced productivity period. Furthermore, the reduced mixing (Fig. 7.2) also affected the availability of isotopically light carbon for diatoms living in the photic zone. Intensification of the stratification conditions prevented the release of CO 2(aq) from the hypolimnion to surface waters, thereby stimulating the δ13C diatom enrichment. 7.4 Conclusions Measurements of δ18O diatom and δ13C diatom are useful proxies for understanding regional climate patterns and lake-catchment processes in Lago Chungará. The interpretation of the δ18O diatom record is based on the hydrology of the lake, and reflects the centennial-to-millennial regional moisture balance, whereas the interpretation of δ13C diatom is probably conditioned by changes in productivity, CO 2(aq) concentration and the source of carbon in the lake. Three main climate phases were identified during the Late Glacial and Early Holocene: a) a humid phase during the Late Glacial-Early Holocene transition (12,400-10,100 cal years BP), b) a dry phase in the Early Holocene (10,100-9,100 cal years BP), and c) a dry-to-humid phase in the latter part of the Early Holocene period (9,100-8,400 cal years BP). Although the Late Glacial to Early Holocene wet phase coincides with a trend to an insolation minimum, the establishment of long-term La Niña-like conditions in the tropical Pacific, as indicated by the high values in the ENSO activity intensity, seems to have overcome the precessional forcing. Evidence of the Younger Dryas-like event was not obtained in Lago Chungará. The insolation minimum at ca 10,000 cal years BP played a key role during the Early Holocene (10,100-9,100 cal years BP), giving rise 104 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles to the probable northward migration of the ITCZ. This, in addition to the ENSO weakening, would have plunged the tropical Andes into a severe drought. The return to more humid conditions began around 9,000 cal years BP, following a slight increase in summer insolation. This period was probably brief since many Andean Altiplano records show the start of a major dry period at ca 8,000 cal years BP as a result of an ENSO weakening. Further analyses throughout the non-laminated unit 2 from the Lago Chungará record would be required to confirm this hypothesis. Isotope analysis of δ13C diatom reveals that several factors, such as productivity, carbon source and carbon concentration account for the δ13C diatom oscillations. Furthermore, these oscillations reveal interactions between the lake water carbon pool and the catchment. During wet periods δ13C diatom shows that the relative contribution of external loadings, in addition to lake productivity, significantly enriched the carbon pool, whereas dry periods favoured the in-lake lacustrine carbon accumulation and the subsequent enrichment of the δ13C diatom values. The release of CO 2 from the hypolimnion during mixing conditions can also deplete the δ13C diatom values. Both δ18O diatom and δ18O diatom analyses can help us to gain a better understanding of the role of lakes in the carbon cycle in the context of global change. Oxygen and carbon diatom isotope records from the Lago Chungará laminated unit (12,400 to 8,400 cal years BP) 105 Chapter 8 Conclusions 8.1 Concluding remarks The isotope analyses and the petrographical study of diatomaceous laminated sediments from the Lago Chungará record yield the following methodological, limnological and climate conclusions: 8.1.1 Methodological conclusions - Analysis of δ18O diatom shows that a number of factors in addition to climate forcings can exert an influence on the δ18O diatom signal. Given that δ18O records cannot only be interpreted in terms of changes in wet–dry conditions, it is essential to investigate the hydrology of each system. The δ18O diatom records can provide valuable insights into the sedimentological processes that trigger the input, transport and accumulation of sedimentary particles, and into the ontogeny of the lake apart from palaeoclimatic issues. - Analysis of δ13C diatom has proved to be a useful tool to reconstruct the carbon cycle in lakes and to improve our understanding of the global carbon cycle. The δ13C diatom record yields information on lake lacustrine productivity, carbon sources, lake-catchment relationships and changes in the concentration of CO 2(aq) . - The Lago Chungará laminated diatom-rich sediments demonstrated that δ18O diatom and δ13C diatom can be used to reconstruct abrupt and rapid climate and enviromental changes. The well laminated nature of the Lago Chungará sediments has enabled us to conduct a continuous sampling lamina by lamina in accordance with the sedimentary structures. This sampling has provided one of the highest resolution records for the study of isotopes in diatom silica. The ultra-high resolution of this record showed significant high-frequency climate and limnological changes which would not otherwise have been identified. As a result, this technique can be used to characterise key climate and environmental events at an ultra-high temporal resolution, complementing the usual lower resolution studies in which these techniques are commonly applied. 107 8.1.2 Limnological conclusions - The laminated sedimentary unit of Lago Chungará is made up of millimetric (3-23 mm) rhythmites composed of multiannual (4-24 years) triplets of white, light-green and dark-green laminae. White laminae consist almost exclusively of very large valves of the euplanktonic diatom Cyclostephanos andinus. Dark-green laminae are constituted by a mixture of diatoms (mainly smaller valves of smaller Cyclostephanos andinus and diatoms of the Discostella stelligera complex) and organic matter. Light- green laminae are made up of components from the white laminae progressively grading upwards into the typical constituents of the dark-green laminae. The typical rhythmite starts with the white laminae deposition changing to dark-green laminae, and occasionally presenting interbedded light-green laminae. The contacts within the rhythmites are transitional, whereas contacts between rhythmites are abrupt. - White laminae were formed during extraordinary short-term diatom blooms (days to weeks). These extraordinary diatom blooms were favoured by episodes of extreme turbulent conditions affecting the water column and/or by episodes of exceptional erosion of the top soil of the catchment. In the former case, upwelling from nutrient-rich hypolimnetic waters gave rise to an extraordinary nutrient availability whereas, in the latter case, allochthonous nutrient inputs would be implicated. Light-green laminae were formed during the end of the diatom super-blooms probably by self-sedimentation processes. Finally, dark-green laminae were deposited over several years under different water column regimes representing baseline lake conditions. - Values of δ18O diatom showed that although both white and green laminae formations occurred in either dry or humid conditions, the extraordinary diatom blooms were more intense (thicker white laminae deposition) during decadal-to centennial-lowstands (i.e. in enhanced δ18O diatom periods). On the other hand, δ13C diatom values indicated that the carbon availability was higher during diatom super-blooms owing to the release of accumulated CO 2(aq) from the hypolimnion in the mixing periods. In most cases, the δ18O diatom composition showed that the white laminae formation was mainly due to falls in the lake level, whereas the green laminae formation was mainly caused by rises in the lake level. - High resolution isotope analysis of the diatoms in this laminated sequence showed complex links between limnology, catchment processes, hydrology and climate forcings. The δ18O diatom and δ13C diatom results highlight the importance of using both proxies in combination. Changes in carbon and biogeochemical cycles indicated by δ13C diatom are linked to hydrological changes evidenced by δ18O diatom . During wet events indicated by δ18O diatom , low values of δ13C diatom show that the relative contribution of 108 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles external loadings, in addition to lake productivity, significantly increased the carbon pool. By contrast, dry periods favoured the in-lake lacustrine carbon accumulation and an increase in δ13C diatom values. However, the release of the CO 2(aq) from the hypolimnion during the mixing periods could also favour the depletion of the δ13C diatom values. Therefore, in this region, periods with enhanced precipitation led to a larger productivity, a greater exportation of CO 2 to the atmosphere, or both. 8.1.3 Climate conclusions - This record clearly shows, that depending on the temporal-scale, one forcing type could prevail over the others when interpreting δ18O diatom and δ13C diatom in Lago Chungará: - At centennial-to millennial-scales, in addition to climate, hydrological factors such as changes in the groundwater/evaporation loss ratio and in the lake size can also play a major role. This must be borne in mind when interpreting the δ18O diatom results. Nevertheless, the long-term evolution of the diatom oxygen isotope record from the Lago Chungará laminated unit reflects long-term changes in the regional moisture balance that are related to orbital forcing (insolation) and ENSO-like conditions. Three main climate stages were identified during Late Glacial and Early Holocene: a) A wet phase during Late Glacial-Early Holocene transition (12,400-10,100 cal years BP) b) A dry phase during the Early Holocene (10,100-9,100 cal years BP) c) A dry-to-wet phase throughout the last years of the Early Holocene period (9,100-8,400 cal years BP) - At decadal scales, this proxy was used as an indicator of high-frequency climate phenomena. The δ18O diatom analysis of the dark-green laminae present in the Late Glacial-Early Holocene transition (11,990-11,450 cal years BP) showed two major events marked by an increase in moisture conditions (11,800 and 11,550 cal years BP) and one major dry event (between 11,990 and 11,800 cal years BP). Apart from this, several minor depletions at a decadal time scale were associated with short-term events with more rainfall. These abrupt and rapid shifts in the hydrological balance of Lago Chungará may due to changes in the strength and position of the Bolivian High. Changes in the atmospheric conditions over the Altiplano at this time scale, and hence in the position of the Bolivian High were triggered by both the ENSO and solar activity. Therefore, the shift from Glacial to Interglacial conditions in the tropical latitudes of South America consisted in see-saw 109 Conclusions moisture oscillations rather than in a progressive moisture increase, which has been reported in other low-resolution environmental reconstructions. - The comparison of δ18O diatom values from the Late Glacial-Early Holocene transition with terrigenous input and water availability reconstructions (previously performed for Lago Chungará) shows a good agreement. Nevertheless, it is necessary to consider a systematic time lag of up to 50 years since the terrigenous inputs and the effective moisture availability react earlier than δ18O diatom . This is mainly due to the time needed to change the values of δ18O lakewater and its subsequent incorporation into the diatom frustules. The time lag highlights the fact that the proxies do not always react at the same time to environmental forcings. This should be recognised in high resolution palaeolimnological reconstructions. 8.2 Perspectives and future work The results obtained provide valuable insights into isotope research and into the use of lacustrine sediments in palaeonvironmental reconstructions. As the study of isotopes in biogenic silica has developed mainly in the last decade, further work is required to improve this technique and its interpretation. The Isotopes in Biogenic Silica (IBiS) group is currently developing a number of new approaches to the use of different isotopes and different organisms. These approaches include the new analyses of δ13C diatom , δ15N diatom and δ30Si diatom , which will assume greater importance in the near future. Some of these new isotopes were employed in this PhD Thesis with promising results. In addition, studies on diatom cultures would provide new opportunities to control diatom-environment relationships or even species specific comparisons. In the present work, isotope analyses were successfully applied to the well preserved diatoms of the laminated sedimentary unit from Lago Chungará. The samples contained little organic matter and few carbonates, and the diatoms showed no signs of dissolution or diagenesis. Nevertheless, the upper sedimentary unit is made up of diatoms, organic matter and tephra clasts. This non siliceous and/or non diatomitic material hampers the direct application of δ18O diatom as environmental and climate proxies. The study of this upper unit would offer fresh insights into the methodology, especially in sample purification and in the effect of early-diagenesis on δ18O diatom signal. Local and regional palaeoclimate patterns for the period spanning from 8,400 to 1,300 cal years BP would also be obtained. This would probably include the long- and short-temporal scale evolution of local and regional moisture, the main forcings that govern it and the impact of abrupt climate phenomena, such as the ENSO. 110 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles Bibliography Abbott MB, Seltzer GO, Kelts K, Southon J. 1997. Holocene paleohydrology of the tropical Andes from Lake Records. Quaternary Research 47: 70–80. Abbott MB, Wolfe PW, Aravena R, Wolfe AP, Seltzer GO. 2000. Holocene hydrological reconstructions from stable isotopes and paleolimnology, Cordillera Real, Bolivia. Quaternary Science Reviews 19: 1801–1820. Abbott MB, Wolfe BB, Wolfe AP, Seltzer GO, Aravena R, Mark BG, Polissar PJ, Rodwell DT, Rowe HD, Vuille M. 2003. Holocene paleohydrology and glacial history of the central Andes using multiproxy lake sediment studies. Palaeogeography, Palaeoclimatology, Palaeoecology 194: 123–138. Aceituno P. 1988. On the functioning of the Southern Oscillation in the South American sector. Part I: surface climate. Monthly Weather Review 116: 505–524. Aceituno P, Montecinos A. 1993. Circulation anomalies associated with dry and wet periods in the South American Altiplano. Proccedings Fourth International Conference on Southern Hemisphere Meteorology, American Meteorological Society: Hobart, Australia; 330–331. Alldredge AL, Gotschalk C, Passow U, Riebesell U. 1995. Mass aggregation of diatom blooms: Insights from a mesocosm study. Deep Sea Research Part II: Topical Studies in Oceanography 42: 9–27. Allison GB, Barnes CJ, Hughes MW, Leaney FWJ. 1984. Effect of climate and vegetation on oxygen-18 and deuterium profiles in soils. In Isotope Hydrology. Proccedings Symposium Vienna: IAEA, Vienna, Austria; 105–122. Allmendinger RW, Jordan TE. 1997. The Central Andes. In Earth structure, An introduction to Structural Geology and Tectonics, Van der Pluijm VA, Marshak S. (Eds.). WCB/McGraw-Hill: 430–434. Allmendinger RW, Jordan TE, Kay SM, Isacks BL. 1997. Evolution of the Puna- Altiplano Plateau of the central Andes. Annual Review of Earth and Planetary Sciences 25: 139–74. Amilibia A. 2002. Inversion tectónica en la Cordillera de Domeyko, Andes del Norte de Chile. PhD Thesis, Universitat de Barcelona: Barcelona. Amilibia A, Sàbat F, McClay KR, Muñoz JA, Roca E, Chong G. 2008. The role of inherited tectono-sedimentary architecture in the development of the central Andean mountain belt: Insights from the Cordillera de Domeyko. Journal of Structural Geology 30: 1520–1539. Anderson NJ. 2000. Diatoms, temperature and climatic change. European Journal of Phycology 35: 307–314. Anderson RY, Dean WE. 1988. Lacustrine varve formation through time. Palaeogeography, Palaeoclimatology, Palaeoecology 62: 215–135. Antico PL. 2009. Relationships between autumn precipitation anomalies in southeastern South America and El Nino event classification. International Journal of Climatology 29: 719–727. Aravena R, Suzuki O, Peña H, Pollastri A, Fuenzalida H, Grilli A. 1999. Isotopic composition and origin of the precipitation in northern Chile. Applied Geochemistry 14: 411–422. Argollo J, Mourguiart P. 2000. Late Quaternary climate history of the Bolivian Altiplano. Quaternary International 72: 37–51. Armbrust EV. 2009. The life of diatoms in the world’s oceans. Nature 459: 185–192. Ashley GM, Moritz LE. 1979. Determination of lacustrine sedimentation rates by radioactive fallout (137Cs), Pitt Lake, British Columbia. Canadian Journal of Earth Science 16: 965–970. Baker P. 2002. Trans-Atlantic climate connections. Science 296: 67–68. Baker PA, Seltzer GO, Fritz SC, Dunbar RB, Grove MJ, Tapia PM, Cross, SL, Rowe HD, Broda JP. 2001a. The history of South American tropical precipitation for the past 25,000 years. Science 291: 640–643. Baker PA, Rigsby CA, Seltzer GO, Fritz SC, Lowenstein TK, Bacher NP, Veliz C. 2001b. Tropical climate changes at millennial and orbital timescales on the Bolivian Altiplano. Nature 409: 698–701. Bao R, Saez A, Servant-Vildary S, Cabrera L. 1999. Lake-level and salinity reconstruction from diatom analysis in Quillagua Formation (Late Neogene, Central Andean Forearc, Nothern Chile). Paleogeography, Paleoclimatology, Paleoecology 153: 309–335. Barichivich J, Sauchyn DJ, Lara A. 2009. Climate signals in high elevation tree-rings from the semiarid Andes of north-central Chile: responses to regional and largescale variability. Palaeogeography, Palaeoclimatology, Palaeoecology 281: 320–333. Barker PA, Street-Perrott FA, Leng MJ, Greenwood PB, Swain DL, Perrott RA, Telford RJ, Ficken KJ. 2001. A 14 ka oxygen isotope record from diatom silica in two alpine tarns on Mt Kenya. Science 292: 2307–2310. Barker P, Telford R, Gasse F, Thévenon F. 2002. Late Pleistocene and Holocene palaeohydrology of Lake Rukwa, Tanzania, inferred from diatom analysis. Palaeogeography Palaeoclimatology Palaeoecology 187: 295–305. Barker PA, Leng MJ, Gasse F, Huang Y. 2007. Century-to-millennial scale climatic variability in Lake Malawi revealed by isotope records. Earth and Planetary Science Letters 261: 93–103. Barry RG, Chorley RJ. 1992. Atmosphere, weather and climate. 6th ed. Publisher Routledge: London. Battarbee RW, Flower RJ. 1984. The inwash of catchment diatoms as a source of error in the sediment-based reconstruction of pH in an acid lake. Limnology and Oceanography 29: 1325–1329. 111 Battarbee RW, Carvalho L, Jones VJ, Flower RJ, Cameron NG, Bennion H, Juggins S. 2001. Diatoms. In Tracking Environmental Change Using Lake Sediments,Terrestrial, Algal, and Siliceous Indicators, vol. 3. Last WM, Smol JP. (Eds.). Kluwer Academic Publishers: Dordrecht, The Netherlands; 155–202. Bedard C, Knowles R. 1991. Hypolimnetic O 2 consumption, denitrification, and methanogenesis in a thermally stratified lake. Canadian Journal of Fisheries and Aquatic Sciences 48: 1048–1054. Berger A, Loutre MF. 1991. Insolation values for the climate of the last 19 million years. Quaternary Science Reviews 10: 297–317. Berger WH, Killingley JS, Vincent E. 1978. Stable isotopes in deep-sea carbonates. Oceanologica Acta 1: 203–316. Berner EK, Berner RA. 1987. The Global Water Cycle: Geochemistry and Environment. Prentice-Hall, Inc.: Englewood Cliffs, NJ; 142–155. Betancourt JL, Latorre C, Rech JA, Quade J, Rylander KA. 2000. A 22,000-year record of monsoonal precipitation from northern Chile’s Atacama desert. Science 289: 1542–1546. Bigeleisen J, Mayer MG. 1947. Calculation of equilibrium constants for isotopic exchange reactions. Journal of Chemical Physics 15: 261–267. Bigeleisen J, Wolfsberg M. 1958. Theoretical and experimental aspects of isotope effects in chemical kinetics. Advances in Chemical Physics 1: 15–76. Bird BW, Abbott MB, Kutchko B, Finney BP. 2009. A 2000-year Varve-Based Climate Record from the Central Brooks Range, Alaska. Journal of Paleolimnology 41: 25–41. Birks HJB, Line JM, Juggins S, Stevenson AC, Ter Braak CJF. 1990. Diatoms and pH reconstruction. Philosophical Transactions of the Royal Society of London, Series B. Bootsma HA, Hecky RE, Johnson TC, Kling HJ, Mwita J. 2003. Inputs, outputs, and internal cycling of silica in a large, tropical lake. Journal of Great Lakes Research 29: 121– 138. Bos DG, Cumming BF, Smol JP. 1999. Cladocera and Anostraca from the interior plateau of British Columbia, Canada, as paleolimnological indicators of salinity and lake level. Hydrobiologia 392: 129–141. Boschker HTS, Middelburg JJ. 2002. Stable isotopes and biomarkers in microbial ecology. Fems Microbiology Ecology 40: 85–95. Bradbury P, Cumming B, Laird K. 2002. A 1500-year record of climatic and environmental change in Elk Lake, Minnesota III: measures of past primary productivity. Journal of Paleolimnology 27: 321–40. Bradley R, Vuille M, Hardy DR, Thompson LG. 2003. Low latitude ice cores record Pacific sea surface temperatures. Geophysical Research Letters 30: 1174. Brandriss ME, O’Neil JR, Edlund MB, Stoermer EF. 1998. Oxygen isotope fractionation between diatomaceous silica and water. Geochimica et Cosmochimica Acta 62: 1119–1125. Brauer A, Endres C, Gu¨nter C, Litt T, Stebich M, Negendank JFW. 1999. High resolution sediment and vegetation responses to Younger Dryas climate change in varved lake sediments from Meerfelder Maar, Germany. Quaternary Science Reviews 18: 321–329. Brenner M, Whitmore TJ, Lasi MA, Cable JE, Cable PH. 1999. Stable isotope δ13C, δ15N signatures of sedimented organic matter as indicators of historical lake trophic state. Journal of Paleolimnology 22: 205–221. Brewer TS, Leng MJ, Mackay AW, Lamb AL, Tyler JJ, Marsh NG. 2008. Unravelling contamination signals in biogenic silica oxygen isotope composition: the role of major and trace element geochemistry. Journal of Quaternary Science 23: 321–330. Brönmark C, Hansson LA. 2005. The Biology of Lakes and Ponds. Oxford University Press: Oxford. Cabrera L, Gierlowski-Kordesch EH, Alonso-Zarza AM, Arenas-Abad C. 2009. Limnogeology: Ancient and modern tales of an evolving Earth. Sedimentary geology 222: 1–2. Cane MA. 2005. The evolution of El Niño, past and future. Earth and Planetary Science Letters 230: 227–240. Carrera N. 2009. Inversió tectònica i evolució estructural de la Cordillera Oriental Meridional (Valls Calchaquís, NW d’Argentina). PhD Thesis, Universitat de Barcelona: Barcelona. Catalan J, Fee EJ. 1994. Interannual variability in limnic ecosystems: origin, patterns and predictability. In Limnology Now: A Paradigm of Planetary Problems. Margalef R. (Ed.). Elsevier: Armsterdam, The Netherlands; 81–97. Catalan J, Pla S, Garcia J, Camarero L. 2009. Climate and CO 2 saturation in an alpine lake throughout the Holocene. Limnology and Oceanography 54: 2542–2552. Chacko T, Cole DR, Horita J. 2001. Equilibrium oxygen, hydrogen, and carbon isotope fractionation factors applicable to geologic systems. In Stable Isotope Geochemistry, Valley JW, Cole DR. (Eds.). Reviews in Mineralogy and Geochemistry 43: Mineralogical Society of America, Washington, D.C.; 1–81. Chalié F, Gasse F. 2002. Late-Glacial–Holocene diatom record of water chemistry and lake-level change from the tropical East African Rift Lake Abiyata (Ethiopa). Palaeogeography, Palaeoclimatology, Palaeoecology 187: 259–283. Chang AS, Patterson RT, McNeely R. 2003. Seasonal sediment and diatom record from late Holocene laminated sediments, Effingham Inlet, British Columbia, Canada. Palaios 18: 477–494. Chepurnov VA, Mann DG, Sabbe K, Vyverman W. 2004. Experimental studies on sexual reproduction in diatoms. International Review of Cytology 237: 91–154. Chiang JCH. 2009. The tropics in Paleoclimate. Annual Review in Earth and Planetary Sciences 37: 263–297. Chikita K, Joshi SP, Jha J, Hasegawa H. 2000. Hydrological and thermal regimes in a supraglacial lake: Imja, Khumbu, Nepal Himalaya. Hydrological Sciences Journal 45: 507–521. Chu G, Liu J, Schettler G, Li J, Sun Q, Gu Z, Lu H, Liu Q, Liu T. 2005. Sediment fluxes and varve formation in Sihailongwan, a maar lake from northeastern China. Journal of Paleolimnology 34: 311–324. Clapperton CM. 1997. Late Quaternary glacier advances and paleolake high-stands in Bolivian Altiplano. Quaternary International 34: 39–49. Clark I, Fritz P. 1997. Environmental isotopes in Hydrogeology. Lewis Publishers: New York. Clavero JE, Sparks SJ, Huppert HE. 2002. Geological constraints on the emplacement mechanism of the Parinacota debris avalanche, northern Chile. Bulletin of Volcanology 64: 40–54. Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 112 Clavero JE, Sparks SJ, Polanco E, Pringle M. 2004. Evolution of Parinacota volcano, Central Andes, northern Chile. Revista Geológica Chile 31: 317–347. Clayton RN, Mayeda TK. 1963. The use of bromine pentafluoride in the extraction of oxygen from oxide and silicates for isotope analysis. Geochimica et Cosmochimica Acta 27: 43–52. Clement AC, Cane MA, Seager R. 2001. An orbitally driven tropical source for abrupt climate change. Journal of Climate 14: 2369–2375. Cohen AS. 1990. Tectono-stratigraphic model for sedimentation in Lake Tanganyika, Africa. In Lacustrine Basin Exploration- Case Studies and Modern Analogs, Katz BJ. (Ed.). American Association of Petroleum Geologists 50: 137–150. Cohen AS. 2003. Paleolimnology: the history and evolution of lake systems. Oxford University Press: New York. Coira B, Davidson J, Mpodozis C, Ramos V. 1982. Tectonic and magmatic evolution of the Andes of Northern Argentina and Chile. Earth-Science Reviews 18: 303-332. Cole JE, Rind D, Webb RS, Jouzel J, Healy R. 1999. Climatic controls on interannual variability of precipitation delta O-18: simulated influence of temperature, precipitation amount, and vapor source region. Journal of Geophysical Research— Atmospheres 104: 14223–14235. Cole JJ, Caraco NF, Kling GW, Kratz TK. 1994. Carbon-dioxide supersaturation in the surface waters of lakes. Science 265: 1568–1570. Cole JJ, Carpenter SR, Kitchell JF, Pace ML. 2002. Pathways of organic carbon utilization in small lakes: Results from a whole- lake 13C addition and coupled model. Limnology and Oceanography 47: 1664–1675. Cole JJ, Prairie Y T, Caraco NF, McDowell WH, Tranvik LJ, Striegl RG, Duarte CM, Kortelainen P, Downing JA, Middelburg JJ, Melack J. 2007. Plumbing the global carbon cycle: Integrating inland waters into the terrestrial carbon budget. Ecosystems 10: 171–184. Colman SM, Peck JA, Karabanov EB, Carter SJ, Bradbury JP, King JW, Williams DF. 1995. Continental climate response to orbital forcing from biogenic silica records in Lake Baikal. Nature 378: 769–771. Conley DJ, Kilham SS, Theriot E. 1989. Differences in silica content between marine and freshwater diatoms. Limnology and Oceanography 34: 205–213. Conley DJ, Schelske CL. 2001. Biogenic silica. In Tracking Environmental Change Using Lake Sediments: Biological Methods and Indicators, Smol J P, Birks HJB, Last WM. (Eds.). Kluwer Academic Press: 281-293. Coutand I, Cobbold PR, de Urreiztieta M, Gautier P, Chauvin A, Gapais D, Rossello E, Lopez- Gamundi O. 2001. Style and history of Andean deformation, Puna plateau, northwestern Argentina. Tectonics 20: 210–234. Crawford SA, Higgins MJ, Mulvaney P, Wetherbee R. 2001. Nanostructure of the diatom frustule as revealed by atomic force and scanning electron microscopy. Journal of Phycology 37: 543–554. Criss RE. 1999. Principles of Stable Isotope Distribution. Oxford University Press: Oxford. Cross SL, Baker PA, Seltzer GO, Fritz SC, Dunbar RB. 2001. Late Quaternary climate and hydrology of tropical South America inferred from an isotopic and chemical model of lake Titicaca, Bolivia and Peru. Quaternary Research 56: 1–9. Crosta X, Koç N. 2007. Diatoms:frommicropaleontologytoisotope geochemistry. In Proxies in Late Cenozoic Paleoceanography, Developments in Marine Geology Series, Vol.1. Hilaire-Marcel C, de Vernal A. (Eds.). Elsevier: Amsterdam, The Netherlands; 327–369. Crosta X, Shemesh A. 2002. Reconciling down core anticorrelation of diatom carbon and nitrogen isotopic ratios from the Southern Ocean. Paleoceanography 17: 1010. Damnati B, Taieb M, Williamson D. 1992. Laminated deposits from Lake Magadi (Kenya); climatic contrast effect during the maximum wet period between 12,000-10,000 yrs BP. Bulletin de la Societe Geologique de France 163: 407–414. Dansgaard W. 1964. Stable isotopes in precipitation. Tellus 16: 436–468. Darling WG, Bath A, Gibson JJ, Rozanski K. 2005. Isotopes in water. In Isotopes in Palaeoenvironmental Research, Leng MJ (ed.). Springer: Dordrecht, Netherlands;1–66. Davison W. 1993. Iron and manganese in lakes. Earth-Science Reviews 34: 119–163. Dean WE, Gorham E. 1998. Magnitude and significance of carbon burial in lakes, reservoirs, and peatlands. Geology 26: 535–538. De Batist M, Imbo Y, Vermeesch P, Klerkx J, Giralt S, Delvaux D, Lignier V, Beck C, Kalugin I, Abdrakhmatov KE. 2002. Bathymetry and sedimentary environments of Lake Issyk-Kul, Kyrgyz Republic (Central Asia): a large, high altitude, tectonic lake. In Lake Issyk-Kul: Its Natural Environment, Klerkx J, Imanackunov B. (eds.). NATO Science Series, IV. Earth and Environmental Sciences, Vol. 13. Kluwmer Academic Publisher: Dordrecht; 101–123. De la Rocha CL. 2002. Measurement of silicon stable isotope natural abundances via multicollector inductively coupled plasma mass spectrometry (MC-ICP-MS). Geochemistry Geophysics Geosystems 3: 1–8. Des Combes HJ, Esper O, De la Rocha CL, Abelmann A, Gersonde R, Yam R, Shemesh A. 2008. Diatom δ13C, δ15N, and C/N since the Last Glacial Maximum in the Southern Ocean: Potential impact of species composition. Paleoceanography 23: PA4209. De Silva S, Francis P. 1991. Volcanoes of the Central Andes. Springer-Verlag: Berlin. Devevey E, Bitton G, Rossel D, Ramos LD, Guerrero LM, Tarradellas J. 1993. Concentration and bioavailability of heavy-metals in Lake Yojoa (Honduras). Bulletin of Environmental Contamination and Toxicology 50: 253–259. Dewey JF, Bird JM. 1970. Mountain belts and the new global tectonics. Journal of Geophysical Research 75: 2625–2647. Diaz HF, Markgraf V. 1992. El Niño. Cambridge University Press, Cambridge. Dincer T, Al-Mugrin A, Zimmermann U. 1974. Study of the infiltration and recharge through sand dunes in arid zones with special reference to the stable isotopes and thermonuclear tritium. Journal of Hydrology 23: 79–87. Dinsmore WP, Scrimgeour GJ, Prepas EE. 1999. Empirical relationships between profundal macroinvertebrate biomass and environmental variables in boreal lakes of Alberta, Canada. Freshwater Biology 41: 91–100. Dorador C, Pardo R, Vila I. 2003. Variaciones temporales de parámetros físicos, químicos y biológicos de un lago de altura: el caso del Lago Chungará. Revista Chilena de Historia Natural 76: 15–22. Drucker DG, Bocherens H, Billiou D. 2003. Evidence of shifting environmental conditions in southwestern France from 33,000 to 15,000 years ago derived from carbon-13 and nitrogen-15 natural abundances in collagen of large herbivores. Earth and Planetary Science Letters 216: 163–173. Bibliography 113 Duarte CM, Prairie YT. 2005. Prevalence of heterotrophy and atmospheric CO 2 emissions from aquatic ecosystems. Ecosystems 8: 862–870. Durand A. 1982. Oscillations of Lake Chad over the past 50,000 years: new data and new hypothesis. Palaeogeography, Palaeoclimatology, Palaeoecology 39: 37–53. Dussart B. 1961. Proprietes utiles de certains sediments lacustres. Comptes rendus hebdomadaires des seances de l’academie des sciences 252: 2581. Earle LR, Warner BG, Aravena R. 2003. Rapid development of an unusual peat-accumulating ecosystem in the Chilean Altiplano. Quaternary Research 59: 2–11. Edgar LA, Pickett-Heaps JD. 1984. Diatom Locomotion. In Progress in Phycological Research. Round FE, Chapman DJ. (Eds.). Biopress Ltd: Bristol; 47–88. Emiliani C. 1955. Pleistocene Temperatures. Journal of Geology 63: 538–78. Emiliani C. 1991. Planktic/Planktonic, Nektic/Nektonic, Benthic/Benthonic. Journal of Paleontology 65: 329. Encyclopedia Britannica. 1962. William Benton: London, 23 vols. Eronen M, Zetterberg P, Briffa KR, Lindholm M, Meriläinen J, Timonen M. 2002. The supra-long Scots pine tree-ring record for Finnish Lapland: Part 1, chronology construction and initial inferences. The Holocene 12: 673–680. Eugster HP, Hardie LA. 1978. Saline lakes. In Lakes: Chemistry, Geology. Physics, Lerman A (Ed.). Springer-Verlag: New York; 237–293. Falkowski, P. G., Katz, M. E., Knoll, A. H., Quigg, A., Raven, J. A., Schofield, O. and Taylor, F. J. R. 2004. The Evolution of Modern Eukaryotic Phytoplankton. Science 305: 354-360. Falvey M, Garreaud R. 2005. Moisture variability over the South American Altiplano during the SALLJEX observing season. Journal Geophysical Research 110: D22105. Field CB, Behrenfeld MJ, Randerson JT, Falkowski P. 1998. Primary production of the biosphere: integrating terrestrial and oceanic components. Science 281: 237–240. Filippi ML, Lambert P, Hunziker J, Kübler B, Bernasconi S. 1999. Climatic and anthropogenic influence on the stable isotope record from bulk carbonates and ostracodes in Lake Neuchâtel, Switzerland, during the last two millennia. Journal of Paleolimnology 21: 19–34. Fogel ML, Cifuentes LA. 1993. Isotope fractionation during primary production. In Organic Geochemistry, Engel MH, Macko SA. (Eds.). Plenum Press: New York; 73–98. Francus P, Bradley RS, Lewis T, Abbott M, Retelle M, Stoner JS. 2008, Limnological and sedimentary processes at Sawtooth Lake, Canadian High Arctic, and their influence on varve formation. Journal of Paleolimnology 40: 963–985. Fritz P, Fontes JCh. 1986. Handbook of Environmental Isotope Geochemistry, Vol 2. Elsevier: Amsterdam. Fritz SC. 2008. Deciphering climatic history from lake sediments. Journal of Paleolimnology 39: 5–16. Fritz SC, Junggins F, Battarbee RW, Engstrom DR. 1991. Reconstruction of past changes in salinity and climate using a diatom- based transfer function. Nature 352: 706–708. Fritz SC, Cumming BF, Gasse F, Laird KR. 1999. Diatoms as indicators of hydrologic and climatic change in saline lakes. In The Diatoms: Applications for the Environmental and Earth Sciences, Stoermer EF, Smol JP. (Eds.). Cambridge University Press: Cambridge; 41–72. Fritz SC, Baker PA, Lowenstein T K, Seltzer GO, Rigsby CA, Dwyer GS, Tapia PM, Arnold KK, Ku TL, Luo S. 2004. Hydrologic variation during the last 170,000 years in the southern hemisphere tropics of South America. Quaternary Research 61: 95–104. Fritz SC, Baker PA, Tapia P, Garland J. 2006. Spatial and temporal variation in cores from Lake Titicaca, Bolivia/Peru during the last 13 000 years. Quaternary International 158: 23–29. Fry B. 1996. 13C/12C fractionation by marine diatoms. Marine Ecology-Progress Series 134: 283–294. Garreaud RD, Battisti DS. 1999. Interannual (ENSO) and interdecadal (ENSO-like) variability inthe Southern Hemisphere tropospheric circulation. Journal of Climate 2: 2113–2123. Garreaud RD, Vuille M, Clement AC. 2003. The climate of the Altiplano: observed current conditions and mechanisms of past changes. Palaeogeography, Palaeoclimatology, Palaeoecology 194: 5–22. Garreaud R, Vuille M, Compagnucci R, Marengo J. 2009. Present-day South American climate. Palaeogeography, Palaeoclimatology, Palaeoecology 281: 180–195. Gasse F. 2000. Hydrological changes in the African tropics since the Last Glacial Maximum. Quaternary Science Reviews 19: 189–211. Gasse F, Fontes JC. 1992. Climatic changes in northwest Africa during the last deglaciation (16–7 ka BP). NATO ASI Series 12: Kluwer, Dordrecht; 295–325. Gasse F, Barker P, Gell PA, Fritz SC, Chalié F. 1997. Diatom-inferred salinity in paleolakes: an indirect tracer of climate change. Quaternary Science Reviews 16: 547–563. Gat JR. 1980. Isotope hydrology of very saline lakes. In Hypersaline brines and evaporitic environments, Nissenbaum A. (Ed.). Elsevier: Amsterdam; 1–8. Gat JR. 1996. Oxygen and hydrogen isotopes in the hydrologic cycle. Annual Review of Earth and Planetary Sciences 24: 225–262. Gaupp R, Kött A, Wörner G. 1999. Palaeoclimatic implications of Mio-Pliocene sedimentation in the highaltitude intra-arc Lauca Basin of northern Chile. In Ancient and Recent Lacustrine Systems in Convergent Margins, Cabrera L, Sáez A. (Eds.) Paleogeography, Paleoclimatology, Paleoecology 151: 79–100. Gervais F, Riebesell U. (2001) Effect of phosphorus limitation on elemental composition and stable carbon isotope fractionation in a marine diatom growing under different CO 2 concentrations. Limnology and Oceanography 46: 497–504. Geyh MA, Grosjean M. 2000. Establishing a reliable chronology of lake level changes in the Chilean Altiplano: a result of close collaboration between geochronologists and geomorphologists. Zentralblatt für Geologie und Paläontologie: Teil I 7/8; 985–995. Geyh MA, Schotterer U, Grosjean M. 1998. Temporal changes of the 14C reservoir effect in lakes. Radiocarbon 40: 921–931. Geyh MA, Grosjean M, Núñez L, Schotterer U. 1999. Radiocarbon reservoir effect and the timing of the Late-Glacial/early Holocene Humid phase in the Atacama desert (Northern Chile). Quaternary Research 52: 143–153. Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 114 Bibliography 115 Giddings JC. 1985. A system based on split-flow lateral-transport thin (SPLITT) separation cells for rapid and continuous particle fractionation. Separation Science and Technology 20: 749–768. Gierlowski-Kordesch EH. 2010. Lacustrine carbonates. In Carbonates in Continental Settings: Processes, Facies, and Application, Developments in Sedimentology 61. Alonso-Zarza AM, Tanner LH. (Eds.). Elsevier, Amsterdam; 1–101. Giralt S, Burjachs F, Roca JR, Julià R. 1999. Late glacial to early Holocene environmental adjustment in the Mediterranean semi-arid zone of the Salines playa-lake (Alicante, Spain). Journal of Paleolimnology 21: 449–460. Giralt S, Moreno A, Bao R, Sáez A, Prego R, Valero BL, Pueyo JJ, González-Sampériz P, Taberner C. 2008. Statistical approach to disentangle environmental forcings in a lacustrine record: the Lago Chungará case (Chilean Altiplano). Journal of Palaeolimnology 40: 195–215. Goslar T, van der Knaap WO, Hicks S, Andric M, Czernik J, Goslar E, Räsänen S, Hyötylä H. 2005. Radiocarbon dating of modern peat profiles: pre- and post-bomb 14C variations in the construction of age–depth models. Radiocarbon 47: 115–134. Gosling WD, Bush MB, Hanselman JA, Chepstow-Lusty A. 2008. Glacial–Interglacial changes in moisture balance and the impact on vegetation in the southern hemisphere tropical Andes (Bolivia/Peru). Palaeogeography Palaeoclimatology Palaeoecology 259: 35–50. Gregory-Wodzicki KM. 2000. Uplift history of the Central and Northern Andes: a review. Geological Survey of America Bulletin 112: 1091–1105. Grimm KA, Lange CB, Gill AS. 1996. Biological forcing of hemipelagic sedimentary laminae: evidence from ODP site 893, Santa Barbara Basin, California. Journal of Sedimentary Research 66: 613–624. Grimm KA, Lange CB, Gill AS. 1997. Self-sedimentation of phytoplankton blooms in the geologic record. Sedimentary Geology 110: 151–161. Grosjean M, Geyh MA, Messerli B, Schotterer U. 1995. Late-glacial and early Holocene lake sediments, groundwater formation and climate in the Atacama Altiplano 22-24ºS. Journal of Paleolimnology 14: 241–252. Grosjean M, Núñez L, Cartajena I, Messerli B. 1997. Mid-Holocene climate and culture change in the Atacama Desert, northern Chile. Quaternary Research 48: 239–246. Grosjean M, van Leeuwen JFN, van der Knaap WO, Geyh MA, Ammann B, Tanner W, Messerli B, Núñez L, Valero-Garcés BL, Veit H. 2001. A 22,000 14C year BP sediment and pollen record of climate change from Laguna Miscanti (23ºS), northern Chile. Global and Planetary Change 28: 35–51. Grosjean M, Cartajena I, Geyh MA, Núñez L. 2003. From proxy data to paleoclimate interpretation: the mid-Holocene paradox of the Atacama Desert, northern Chile. Palaeogeography, Palaeoclimatology, Palaeoecology 194: 247–258. Grosjean M, Santoro CM, Thompson LG, Núñez L, Standen VG. 2007. Mid-Holocene climate and cultural change in the South- Central Andes. In Climate Change and Cultural Dynamics: A Global Perspective on Holocene Transitions, Anderson G, Sandweiss DF, Maasch KA. (eds.). Academic Press: San Diego, CA; 51–115. Grottoli A, Eakin CM. 2007. A review of modern coral δ18O and δ14C proxy records. Earth-Science Reviews 81: 67–91. Gu B, Schelske GL, Brenner M. 1996. Relationship between sediment and plankton isotope ratios (δ13C and δ15N) and primary productivity in Florida lakes. Canadian Journal of Fisheries and Aquatic Sciences 53: 875–883. Guillard RRL, Kilham P. 1977. The ecology of marine planktonic diatoms. In The Biology of Diatoms. Werner D. (ed.). Oxford: Blackwell Scientific Publications; 372–469. Gustavson TC. 1975. Bathymetry and sediment distribution in proglacial Malaspina Lake, Alaska. Journal of Sedimentary Petrology 45: 450–461. Haimson M, Knauth LP. 1983. Stepwise fluorination – a useful approach for the isotopic analysis of hydrous minerals. Geochimica et Cosmochimica Acta 47: 1589–1595. Hall RI, Smol JP. 1999. Diatoms as indicators of lake eutrophication. In The Diatoms: Applications for the Environmental and Earth Sciences, Stoermer EF, Smol JP. (Eds.). Cambridge University Press: Cambridge; 128–168. Hamilton-Taylor J, Davison W. 1995. Redox-driven cycling of trace elements in lakes. In Physics and Chemistry of Lakes, Lerman A, Imboden D, Gat J. (Eds.). Springer-Verlag: 217–263. Hamilton-Taylor J, Smith EJ, Davison W, Sugiyama M. 2005. Resolving and modeling the effects of Fe and Mn redox cycling on trace metal behaviour in a seasonally anoxic lake. Geochimica et Cosmochimica Acta 69: 1947–1960. Harris GP. 1986. Phytoplankton Ecology. Chapman and Hall: London. Hecky RE, Mopper K, Kilham P, Degens ET. 1973. Amino-acid and sugar composition of diatom cell-walls. Marine Biology 19: 323–331. Heegaard E, Birks HJB, Telford RJ. 2005. Relationships between calibrated ages and depth in stratigraphical sequences: an estimation procedure by mixed-effect regression. The Holocene 15: 612–618. Heine K. 2000. Tropical South America during the Last Glacial Maximum: evidence from glacial, periglacial and fluvial records. Quaternary International 72: 7–21. Hernández A, Bao R, Giralt S, Leng MJ, Barker PA, Pueyo JJ, Sáez A, Moreno A, Valero-Garcés B, Sloane HJ. 2007. A high- resolution study of diatom oxygen isotopes in a Late Pleistocene to early Holocene laminated record from Lake Chungará (Andean Altiplano, Northern Chile). Geochimica et Cosmochimica Acta 71: A398. Hernández A, Bao R, Giralt S, Leng MJ, Barker PA, Sáez A, Pueyo JJ, Moreno A, Valero-Garcés BL, Sloane HJ. 2008. The palaeohydrological evolution of Lago Chungará (Andean Altiplano, northern Chile) during the Lateglacial and early Holocene using oxygen isotopes in diatom silica. Journal of Quaternary Science 23: 351–363. Hernández A, Giralt S, Bao R, Leng MJ, Barker PA. In press. ENSO and solar activity signals from oxygen isotopes in diatom silica during late glacial-Holocene transition in Central Andes (18ºS). Journal of Paleolimnology. Hernández A, Bao R, Giralt S, Barker PA, Leng MJ, Sloane HJ, Sáez A. Submitted. Biogeochemical processes controlling oxygen and carbon isotopes of diatom silica in lacustrine rhythmites. Palaeogeography, Palaeoclimatology, Palaeoecology. Herrera C, Pueyo JJ, Sáez A, Valero-Garcés BL. 2006. Relación de aguas superficiales y subterráneas en el área del lago Chungará y lagunas de Cotacotani, norte de Chile: un estudio isotópico. Revista Geológica de Chile 33: 299–325. Hill WR. 1996. Effects of light. In Algal Ecology: Freshwater Benthic Ecosystems, Stevenson RJ, Bothwell ML, Lowe RL. (Eds.). Academic Press: New York; 122–149. Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 116 Hill KJ, Taschetto AS, England MH. 2009. South American rainfall impacts associated with inter-El Niño variations. Geophysical Research Letters 36: L19702. Hillyer R, Bush M, Valencia BG, Steinitz-Kannan M, Silman MR. 2009. A 24,000-year paleolimnological history from the Peruvian Andes. Quaternary Research 71: 71–82. Hodell DA, Schelske CL, Fahnenstill GL, Robbings LL. 1998. Biologically-induced calcite and its isotopic composition in Lake Ontario. Limnology and Oceanography 43: 187–199. Hoefs J. 2004. Stable isotope geochemistry. 5th edition. Springer: Berlin Heidelberg. Hollander DJ, MacKenzie JA. 1991. CO 2 control on carbon-isotope fractionation during aqueous photosynthesis: A paleo-pCO 2 barometer. Geology 19: 929–932. Hora J, Singer B, Wörner G. 2007. Volcan eruption and evaporative flux on the thick curst of the Andean Central Volcanic Zone: 40Ar/39Ar constrains from Volcán Parinacota, Chile. Geological Survey of America Bulletin 119: 343–362. Horne AJ, Goldman CR. 1994. Limnology, Second Edition. McGraw-Hill: New York, NY. Houser JN, Bade DL, Cole JJ, Pace ML. 2003. The dual influences of dissolved organic carbon on hypolimnetic metabolism: organic substrate and photosynthetic reduction. Biogeochemistry 64: 247–269. Hu FS, Shemesh A. 2003. A biogenic-silica δ18O record of climatic change during the last glacial–interglacial transition in southwestern Alaska. Quaternary Research 59: 379–385. Hua Q, Barbetti M. 2004. Review of tropospheric bomb C-14 data for carbon cycle modeling and age calibration purposes. Radiocarbon 46: 1273–1298. Huc AY, Le Fournier J, Vandenbroucke M, Bessereau G. 1990. Northern Lake Tanganyika: an example of organic sedimentation in an anoxic rift lake. In Lacustrine basin exploration. Case studies and modern analogs. American Association of Petroleum Geologists 50: 169–185. Hurrell ER. 2009. Climate change and biogeochemical cycles on East African mountains by stable isotopes of diatom frustules. PhD Thesis. Lancaster Environment Center, Lancaster University: Lancaster. Hurrell ER, Barker PA, Leng MJ, Vane CH, Wynn P, Kendrick CP, Verschuren D, Street-Perrott FA. Submitted. Developing a methodology for carbon isotope analysis of lacustrine diatoms. Journal of Palaeolimnology. Hutchinson GE. 1969. Eutrophication, past and present. In Eutrophication: Causes, Consequences, Correctives. National Academy of Sciences: Washington DC; 17–26. Hutchinson GE, Loffler J. 1956. The thermal classification of lakes. Proceedings of the National Academy of Sciences 42: 84–86. Imboden DM, Wüest A. 1995. Mixing mechanisms in lakes. In Physics and chemistry of lakes, Lerman A, Imboden D, Gat J (Eds.). Springer-Verlag: Berlin; 83–138. Isacks BL. 1988. Uplift of the central Andean plateau and bending of the Bolivian orocline. Journal of Geophysical Research 93: 3211–3231. Ishihara S, Kato M, Tanimura Y, Fukusawa H. 2003. Varved lacustrine sediments and diatom assemblages of Lake Fukami, central Japan. Quaternary International 105: 21–24. Ito E. 2001. Application of stable isotope techniques to inorganic and biogenic carbonates. In Tracking Environmental Change Using Lake Sediments, Physical and Geochemical Techniques, vol. 2. Last WM, Smol JP. (Eds.). Kluwer Academic Publishers: Dordrecht, The Netherlands; 351–371. Jacque JMS, Cumming BF, Smol JP. 2009. A 900-yr diatom and chrysophyte record of spring mixing and summer stratification from varved Lake Mina, west-central Minnesota, USA. The Holocene 19: 537–547. James DE. 1971. Plate-tectonic model for the evolution of the central Andes. Geological Society of America Bulletin 82: 3325–3346 Johnson RK, Weiderholm T. 1989. Classification and ordination of profundal macroinvertebrate communities in nutrient poor, oligo-mesohumic lakes in relation to environmental data. Freshwater Biology 21: 375–386. Johnston AM, Raven JA, Beardall J, Leegood RC. 2001. Carbon fixation-photosynthesis in a marine diatom. Nature 412: 40–41. Jones V. 2007. Diatom introduction. In Encyclopedia of Quaternary Science, Elias SA. (ed.). Amsterdam: Elsevier, 476-484. Jones VJ, Leng MJ, Solovieva N, Sloane HJ, Tarasov P. 2004. Holocene climate of the Kola Peninsula; evidence from the oxygen isotope record of diatom silica. Quaternary Science Reviews 23: 833–839. Jonsson A, Meili M, Bergstrom AK, Jansson M. 2001. Whole-lake mineralization of allochthonous and autochthonous organic carbon in a large humic lake (Ortrasket, N. Sweden). Limnology and Oceanography 46: 1691–1700. Jonsson C, Andersson S, Rosqvist GC, Leng MJ. 2010. Reconstructing past atmospheric circulation changes using oxygen isotopes in lake sediments from Sweden. Climate of the Past 6: 49–62. Juillet A. 1980a. Structure de la silice biogenique: nouvelles donnes apportees par l’analyse isotopique de l’oxygene. C.R. Academy of Science: Paris 290D; 1237–1239. Juillet A. 1980b. Analyse isotopique de la silice des diatomees lacustres et marines: fractionnement des isotopes de l’oxygene en fonction de la temperature. Diss. Paris XI These de 3e cycle. Juillet-Leclerc A. 1986. Cleaning process for diatomaceous samples. In 8th Diatom Symposium, Ricard M (Ed.). Koeltz Scientific: Koenigstein, Germany; 733–736. Karlsson J, Byström P, Ask J, Persson L, Jansson M. 2009. Light limitation of nutrient-poor lake ecosystems. Nature 460: 506–9. Katsui Y, González-Ferrán O. 1968. Geología del área neovolcánica de los Nevados de Payachata. Publicación 29: Universidad de Chile, Facultad de Ciencias Físicas y Matemáticas, Departamento de Geología, Santiago. Kelts K, Hsü KJ. 1978. Freshwater carbonate sedimentation. In Lakes: Chemistry, Geology. Physics, Lerman A (Ed.). Springer- Verlag: New York; 294–323. Kemp AES. 1996. Laminated sediments as palaeo-indicators. In Palaeoclimatology and palaeoceanography from laminated sediments. Kemp AES (Ed.). The Geological Society: London; 1–12. Kennett J, Cannariato KG, Hendy IL, Behl RJ. 2003. Methane Hydrates in Quaternary Climate Change: The Clathrate Gun Hypothesis. American Geophysical Union, Washington: DC, USA. Kirilova E, Heiri O, Enters D, Cremer H, Lotter AF, Zolitschka B, Hübener T. 2009. Climate-induced changes in the trophic status of a Central European lake. Journal of Limnology 68: 71–82. Kitchell JF, Carpenter SR. 1993. Variability in lake ecosystems: complex responses by the apical predator. In Human as Components of Ecosystems, McDonnell M, Pickett S. (Eds.). Springer Verlag: New York; 111–124. Knauth LP. 1973. Oxygen and hydrogen isotope ratios in cherts and related rocks. PhD thesis, California Institute of Technology. Kociolek JP, Spaulding SA. 2000. Freshwater diatom biogeography. Nova Hedwigia 71:223–241. Kooistra WHCF, De Stefano M, Mann DG, Medlin LK. 2003. The phylogeny of diatoms. Progress in Molecular and Subcellular Biology 33: 59–97. Kött A, Gaupp R, Wörner G. 1995. Miocene to Recent history of the western Altiplano in northern Chile revealed by lacustrine sediments of the Lauca Basin (18º15'-18º40' S/69º30'-69º 05' W). Geologische Rundschau 84: 770–780. Koutavas A, Lynch-Stieglitz J, Marchitto T, Sachs J. 2002. El Niño-like pattern in ice age tropical Pacific sea surface temperature. Science 297: 226–230. Kröger N, Poulsen N. 2008. Diatoms: from Cell Wall Biogenesis to Nanotechnology. Annual Review of Genetics 42: 83–107. Kröger N, Bergsdorf C, Sumper M. 1994. A new calcium-binding glycoprotein family constitutes a major diatom cell wall component. EMBO Journal 13: 4676–4683. Kull C, Imhof S, Grosjean M, Zech R, Veit H. 2008. Late Pleistocene glaciation in the Central Andes: temperature versus humidity control. A case study from the eastern Bolivian Andes (17ºS) and regional synthesis. Global and Planetary Change 60: 148–164. Labeyrie LD. 1974. New approach to surface seawater paleotemperatures using 18O/16O ratios in silica of diatom frustules. Nature 248: 40–42. Labeyrie L. 1984. Paléoclimatologie, des squelettes d’organisme marins pour reconstruire le climat. Echos du CEA 2: 8–11. Labeyrie LD, Juillet A. 1982. Oxygen isotopic exchangeability of diatom valve silica; interpretation and consequences for palaeoclimatic studies. Geochimica et Cosmochimica Acta 46: 967–975. Lamb AL, Leng MJ, Lamb HF, Mohammed MU. 2000. A 9000-year oxygen and carbon isotope record of hydrological change in a small Ethiopian crater lake. The Holocene 10: 167–177. Lamb AL, Leng MJ, Sloane HJ, Telford RJ. 2005. A comparison of δ18O data from calcite and diatom silica from early Holocene in a small crater lake in the tropics. Palaeogeography, Palaeoclimatology, Palaeoecology 223: 290–302. Lambert A, Giovanoli F. 1988. Records of riverborne turbidity currents and indications of slope failures in the Rhone and Lake Geneva. Limnology and Oceanography 33: 458–468. Lau KM, Weng H. 1995. Climate signal detection using wavelet transform: how to make a time series sing. Bulletin of the American Meteorological Society 76: 2391–2406. Laws EA, Bidigare RR, Popp BN. 1997. Effect of growth rate and CO 2 concentration on carbon isotopic fractionation by the marine diatom Phaeodactylum tricornutum. Limnology and Oceanography 42: 1552–1560. Leland HV, Berkas WR. 1998. Temporal variation in plankton assemblages and physicochemistry of Devils Lake, North Dakota. Hydrobiologia 377: 57–71. Lemoalle J, Dupont B. 1976. Iron-bearing oolites and the present conditions of iron sedimentation in Lake Chad. In Ores in Sediments, Amstutuz GC, Bernard AJ. (Eds.). International Union of Geological Science Vol. A3, Springer: Berlin; 167–178. Leng MJ, Barker PA. 2006. A review of the oxygen isotope composition of lacustrine diatom silica for palaeoclimate reconstruction. Earth-Science Reviews 75: 5–27. Leng MJ, Marshall JD. 2004. Palaeoclimate interpretation of stable isotope data from lake sediment archives. Quaternary Science Reviews 23: 811–831. Leng MJ, Sloane HJ. 2008. Combined oxygen and silicon isotope analysis of biogenic silica. Journal of Quaternary Science 23: 313–319. Leng MJ, Barker PA, Greenwood P, Roberts N, Reed J. 2001. Oxygen isotope analysis of diatom silica and authigenic calcite from Lake Pinarbasi, Turkey. Journal of Paleolimnology 25: 343–349. Leng MJ, Metcalfe SE, Davies SJ. 2005a. Investigating late Holocene climate variability in Central Mexico using carbon isotope ratios in organic materials and oxygen isotope ratios from diatom silica within lacustrine sediments. Journal of Palaeolimnology 34: 413–431. Leng MJ, Lamb AL, Heaton THE, Marshall JD, Wolfe BB, Jones MD, Holmes JA, Arrowsmith C. 2005b. Isotopes in lake sediments. In Isotopes in Palaeoenvironmental Research, Leng MJ (ed.). Springer: Dordrecht, Netherlands; 147–184. Leng MJ, Swann GEA, Hodson MJ, Tyler JJ, Patwardhan SV, Sloane HJ. 2009. The potential use of silicon isotope composition of biogenic silica as a proxy for environmental change. Silicon 1: 65–77. Lenters JD, Cook KH. 1997. On the origin of the Bolivian high and related circulation features of the South American climate. Journal Atmospheric Sciences 54: 656–677. Lewis WM Jr. 1983. A revised classification of lakes based on mixing. Canadian Journal of Fisheries and Aquatic Sciences 40: 1779–1787. Lindqvist JK, Lee DE. 2009. High-frequency paleoclimate signals from Foulden Maar, Waipiata Volcanic Field, southern New Zealand: An Early Miocene varved lacustrine diatomite deposit. Sedimentary Geology 222: 98–110. Livingstone DA, Melack JM. 1984. Some lakes of subsaharan Africa. In Lakes and reservoirs, Taub FB. (Ed.). Elsevier: Amsterdam; 467–497. Lotter AF, Birks HJB. 1997. The separation of the influence of nutrients and climate on the varve time-series of Baldeggersee, Switzerland. Aquatic Sciences 59: 362–375. Lowell T V, Kelly MA. 2008. Was the Younger Dryas Global? Science 321: 348–349. MacDonald JD. 1869. On the structure of the diatomaceous frustule, and its genetic cycle. Annals and Magazine of Natural History 4: 1–8. Mackay AW. 2007. The paleoclimatology of Lake Baikal: a diatom synthesis and prospectus. Earth-Science Reviews 82: 181–215. Mackay AW, Karabanov E, Leng MJ, Sloane HJ, Morley DW, Panizzo VN, Khursevich G, Williams D. 2008. Reconstructing hydrological variability in Lake Baikal during MIS 11: an application of oxygen isotope analysis of diatom silica. Journal of Quaternary Science 23: 365–374. Mann ME, Park J. 1996. Greenhouse warming and changes in the seasonal cycle of temperature: model versus observations. Geophysical Research Letters 23: 1111–1114. Bibliography 117 Mann ME, Park J, Bradley RS. 1995. Global interdecadal and century-scale climate oscillations during the past five centuries. Nature 378: 266–270. Mantua NJ, Hare S, Zhang Y, Wallace JM, Francis RC. 1997. A Pacific interdecadal climate oscillation with impacts on salmon production. Bulletin of the American Meteorological Society 78: 1069–1079. Marchant RA, Hooghiemstra H. 2004. Rapid environmental change in tropical Africa and Latin America about 4000 years before present: a review. Earth Science Reviews 66: 217–260. Marengo J. 1992. Interannual variability of surface climate in the Amazon basin. International Journal of Climatology 12: 853–863. Marengo J. 2004. Climatology of the LLJ east of the Andes as derived from NCEP reanalyses. Journal of Climate 17: 2261–2280. Marengo J. 2007. Climate change and the hydrological modes of the wet tropics. In Tropical Rainforest Responses to Climate Change, Bush MB, Flenley JR. (Eds.). Praxis: Chichester; 237–268. Marengo J, Soares W, Saulo C, Nicolini M. 2004. Climatology of the LLJ east of the Andes as derived from the NCEP reanalyses. Journal of Climate 17: 2261–2280. Margalef R. 1978. Life forms of phytoplankton as survival alternatives in an unstable environment. Oceanologica Acta 1: 493–509. Margalef R. 1983. Limnología. Ediciones Omega: Barcelona. Martin L, Bertaux J, Correge T, Ledru M-P, Mourguiart P, Sifeddine A, Soubiès F, Wirrmann D, Suguio K, Turcq B. 1997. Astronomical forcing of contrasting rainfall changes in tropical South America between 12,400 and 8800 cal yr B.P. Quaternary Research 47: 117–122 Martin JH. 1992. Iron as a limiting factor in oceanic productivity. In Primary Productivity and Biogeochemical cycles in the Sea, Falkowski PG, Woodhead AD. (eds.). New York: Plenum Press, 123–137. Martin P, Granina L, Martens K, Goddeeris B. 1998. Oxygen concentration profiles in sediments of two ancient lakes: Lake Baikal (Siberia, Russia) and Lake Malawi (East Africa). Hydrobiologia 367: 163–174. Maslin MA, Burns, SJ. 2000. Reconstruction of the Amazon basin effective moisture availability over the last 14,000 years. Science 290: 2285–2287. Maslin MA, Swann GEA. 2005. Isotopes in Marine Sediments. In Isotopes in Palaeoenvironmental Research, Leng MJ (ed.). Springer: Dordrecht, Netherlands; 227–290. Matheney RK, Knauth LP. 1989. Oxygen-isotope fractionation between marine biogenic silica and seawater. Geochimica et Cosmochimica Acta 53: 3207–3214. McCrea JM. 1950. On the isotopic chemistry of carbonates and palaeo-temperature scale. Journal of Chemical Physics 18: 849–857. McDermott F, Schwarcz H, Rowe PJ. Isotopes in Speleothems. In Isotopes in Palaeoenvironmental Research, Leng MJ (ed.). Springer: Dordrecht, Netherlands; 185–225. McGregor GR, Nieuwolt S. 1998. Tropical climatology: an introduction to the climates of the low latitudes. 2nd ed. Wiley. New York. McKenzie JA. 1985. Carbon isotopes and productivity in the lacustrine and marine environment. In Chemical Processes in Lakes, Stumm W. (Ed.). Wiley: New York; 99–118. McPhaden MJ, Zebiak SE, Glantz MH. 2006. ENSO as an integrating concept in earth science. Science 314: 1740–1745. Melander LCS. 1960. Isotope Effects on Reaction Rates. Ronald Press Co: New York. Melander L, Saunders WH. 1980. Reaction Rates of Isotopic Molecules. Wiley Interscience: New York, NY. Merkel U, Prange M, Schulz M. 2010. ENSO variability and teleconnections during glacial climate. Quaternary Science Reviews 29: 86–100. Meybeck M. 1995. Global distribution of lakes. In Physics and chemistry of lakes, Lerman A, Imboden D, Gat J (Eds.). Springer- Verlag: Berlin; 1–35. Meyers PA. 2003. Applications of organic geochemistry to paleolimnological reconstructions: A summary of examples from the Laurentian Great Lakes. Organic Geochemistry 34: 261–289. Meyers PA, Teranes JL. 2001. Sediment organic matter. In Tracking Environmental Change Using Lake Sediments, Physical and Geochemical Techniques, vol. 2. Last WM, Smol JP. (Eds.). Kluwer Academic Publishers: Dordrecht, The Netherlands; 239–270. Milligan AJ, Morel FMM. 2002. A proton buffering role for silica in diatoms. Science 297: 1848–1850. Mitchell TP, Wallace JM. 1992. The annual cycle in the equatorial convection and sea surface temperature. Journal of Climate 5: 1140–1156. Moreno A, Giralt S, Valero-Garcés BL, Sáez A, Bao R, Prego R, Pueyo JJ, González-Sampériz P, Taberner C. 2007. A 13 kyr high- resolution record from the tropical Andes: the Chungará Lake sequence (18º S, northern Chilean Altiplano). Quaternary International 161: 4–21. Morley DW, Leng MJ, Mackay AW, Sloane HJ, Rioual P, Battarbee RW. 2004. Cleaning of lake sediment samples for diatom oxygen isotope analysis. Journal of Paleolimnology 31: 391–401. Morley DW, Leng MJ, Mackay AW, Sloane HJ. 2005. Late glacial and Holocene environmental change in the Lake Baikal region documented by oxygen isotopes from diatom silica. Global and Planetary Change 46: 221–233. Moschen R, Lücke A, Schleser G. 2005. Sensitivity of biogenic silica oxygen isotopes to changes in surface water temperature and palaeoclimatology. Geophysical Research Letters 32: L07708. Mourguiart P, Ledru M-P. 2003. Last glacial maximum in an Andean cloud forest environment (Eastern Cordillera, Bolivia). Geology 31: 195–198. Moy CM, Seltzer GO, Rodbell DT, Anderson DM. 2002. Variability of El Niño/Southern Oscillation activity at millenial timescales during the Holocene epoch. Nature 420: 162–165. Mühlhauser H, Hrepic N, Mladinic P, Montecino V, Cabrera S. 1995. Water-quality and limnological features of a high-altitude Andean lake, Chungará in northern Chile. Revista Chilena de Historia Natural 68: 341–349. Muñoz A, Ojeda J, Sanchez-Valverde B. 2002. Sunspot-like and ENSO/NAO-like periodicities in lacustrine laminated sediments of the Pliocene Villarroya Basin (La Rioja, Spain). Journal of Paleolimnology 27: 453–463. Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 118 Muñoz JA, Amilibia A, Carrera N, Mon R, Roca E, Sàbat F. 2005. A geological cross-section of the Andean orogen at 25.5ºS. International Andean Geodynamics Syposium: Barcelona. Negri AJ, Adler RF, Shepherd JM, Huffman G, Manyin M, Neklin EJ. 2004. A 16-year climatology of global rainfall from SSM/ I highlighting morning versus evening differences. 13th Conference on Satellite Meteorology and Oceanography, American Meteorological Society: Norfolk, VA. Newman M, Compo GP, Alexander MA. 2003. ENSO-forced variability of the Pacific Decadal Oscillation. Journal of Climate 16: 3853–3857. Niebler HS, Hubberten HW, Gersonde R. 1999. Oxygen isotopes values of planktic foraminifera: A tool for the reconstruction of surface water stratification. In Use of proxies in paleoceanography, Fischer G, Wefer G, (Eds.). Springer-Verlag: Germany; 165–189. O’Neil JR. 1986. Theoretical and experimental aspects of isotopic fractionation. In Stable Isotopes in High Temperature Geological Processes, Valley JW, Taylor HP, O’Neil JR. (Eds.). Reviews in Mineralogy 16: 1–40. Orme AR. 2007. The tectonic framework of South America. In The physical geography of South America, Veblen TT, Young KR, Orme AR. (Eds.). Oxford University Press: Oxford; 3–22. Owen RB, Crossley R. 1992. Spatial and temporal distribution of diatoms in sediments of Lake Malawi, Central Africa and ecological implications. Journal of Paleolimnology 7: 55–71. Pace ML, Cole JJ, Carpenter SR, Kitchell JF, Hodgson JR, Van de Bogert MC, Bade DL, Kritzberg ES, Bastviken D. 2004. Whole-lake carbon-13 additions reveal terrestrial support of aquatic food webs. Nature 427: 240–243. Paduano GM, Bush MB, Baker PA, Fritz SC, Seltzer GO. 2003. A vegetation and fire history of Lake Titicaca since the Last Glacial Maximum. Palaeogeography, Palaeoclimatology, Palaeoecology 194: 259–279. Park J. 1992. Envelope estimation for quasi-periodic geophysical signals in noise: a multitaper approach. In Statistics in the Environmental and Earth Sciences, Walden AT, Guttorp P. (Eds.). Edward Arnold: London; 189–219. Pfitzer E. 1869. Über den Bau und die Zellteilung der Diatomeen. Botanische Zeitung 27: 774–776. Pilskaln CH, Johnson TJ. 1991. Seasonal signals in Lake Malawi sediments. Limnology and Oceanography 36: 544–557. Placzek C, Quade J, Patchett PJ. 2006. Geochronology and stratigraphy of late Pleistocene lake cycles on the southern Bolivian Altiplano: implications for causes of tropical climate change. GSA Bulletin 118: 515–532. Polissar PJ, Abbott MB, Shemesh A, Wolfe AP, Bradley RS. 2006. Holocene hydrologic balance of tropical South America from oxygen isotopes of lake sediment opal, Venezuelan Andes. Earth and Planetary Science Letters 242: 375–389. Popp BN, Laws EA, Bidigare RR, Dore JE, Hanson KL, Wakeham SG. 1998. Effect of phytoplankton cell geometry on carbon isotopic fractionation. Geochimica et Cosmochimica Acta 62: 69–77. Prasad S, Brauer A, Rein B, Negendank JFW. 2006. Rapid climate change during the early Holocene in western Europe and Greenland. The Holocene 16: 153–158. Ragotzkie RA. 1978. Heat budgets of lakes. Lakes, Chemistry, Geology and Physics, Lerman A. (Ed.). Springer Verlag: 1–19. Ramos VA. 1999. Los depósitos sinorogénicos terciarios de la región andina. Geología Argentina, Anales 29: 651–682. Ramos VA. 2009. Anatomy and global context of the Andes: Main geologic features and the Andean orogenic cycle. In Backbone of the Americas: Shallow Subduction, Plateau Uplift, and Ridge and Terrane Collision, Kay SM, Ramos VA, Dickinson W. (Eds.). Geological Society of America, Memoir 204: 31–65. Rasband WS. 1997-2009. ImageJ. U. S. National Institutes of Health: Bethesda, Maryland, USA; http://rsb.info.nih.gov/ij/ Ravelo C, Hillaire-Marcel C. 2007. The use of oxygen and carbon isotopes of foraminifera in paleoceanography. In Late Cenozoic Paleoceanography, Hillaire-Marcel C, de Vernal A. (Eds.). Elsevier: Amsterdam; 735–764. R Development Core Team. 2008. R: a Language and Environment for Statistical Computing. R Foundation for Statistical Computing: Vienna, Austria; http://www.R-project.org Reimer PJ, Baillie MGL, Bard E, Bayliss A, Beck JW, Bertrand CJH, Blackwell PG, Buck CE, Burr GS, Cutler KB, Damon PE, Edwards RL, Fairbanks RG, Friedrich M, Guilderson TP, Hogg AG, Hughen KA, Kromer B, McCormac G, Manning S, Ramsey CB, Reimer RW, Remmele S, Southon JR, Stuiver M, Talamo S, Taylor FW, van der Plicht J, Weyhenmeyer CE. 2004a. IntCal04 terrestrial radiocarbon age calibration, 0–26 Cal Kyr BP. Radiocarbon 46: 1029–1058. Reimer P, Brown T, Reimer R. 2004b. Discussion: reporting and calibration of post-bomb 14C data. Radiocarbon 46: 1299–1304. Renberg I. 1981. Formation, structure and visual appearance of iron-rich varved lake sediments. Verhandlungen Internationale Vereinigung für Limnologie 21: 94–101. Reynolds CS. 2006. The Ecology of Phytoplankton. Cambridge University Press: Cambridge, UK. Richardson CA. 2001. Molluscs as archives of environmental change. Oceanography and Marine Biology: an Annual Review 39: 103–164. Ricketts RD, Anderson RF. 1998. A direct comparison between the historical record of lake level and δ18O signal in carbonate sediments from Lake Turkana, Kenya. Limnology and Oceanography 43: 811–822. Riebesell U, Wolf-Gladrow D, Smetacek V. 1993. Carbon dioxide limitation of marine phytoplankton growth rates. Nature 361: 249 –251. Rietti-Shati M, Shemesh A, Karlen W. 1998. A 300-year climate record from biogenic silica oxygen isotopes in an equatorial highaltitude lake. Science 281: 980–982. Rigsby CA, Bradbury JP, Baker PA, Rollins SM, Warren MR. 2005. Late Quaternary palaeolakes, rivers, and wetlands on the Bolivian Altiplano and their palaeoclimatic implications. Journal of Quaternary Science 20: 671–691. Rings A, Lucke A, Schleser GH. 2004. A new method for the quantitative separation of diatom frustules from lake sediments. Limnology and Oceanography: Methods 2: 25–34. Rioual P, Andrieu-Ponel V, Rietti-Shati M, Battarbee RW, de Beaulieu JL, Cheddadi R, Reille M, Svobodova H, Shemesh A. 2001. High-resolution record of climate stability in France during the last interglacial period. Nature 413: 293–296. Rittenour TR, Brigham-Grette J, Mann ME. 2000. El Niño-like climate teleconnections in New England during the Late Pleistocene. Science 288: 1039–1042. Robbins EI. 1983. Accumulation of fossil fuels and metallic minerals in active and ancient rift lakes. Tectonophysics 94: 633–658. Bibliography 119 Robinson RS, Brunelle BG, Sigman DM. 2004. Revisiting nutrient utilization in the glacial Antarctic: Evidence from a new method for diatom-bound N isotopic analysis. Paleoceanography 19: PA3001. Rodbell DT, Seltzer GO. 2000. Rapid ice margin fluctuations during the Younger Dryas in the tropical Andes. Quaternary Research 54: 328–338. Rodbell DT, Seltzer GO, Anderson DM, Abbott MB, Enfield DB, Newman JH. 1999. An 15, 000-year record of El Niño-driven alluviation in southwestern Ecuador. Science 283: 516–520. Rodó X, Rodríguez-Arias MA. 2004. El Niño–Southern oscillation: absent in the early holocene? Journal of Climate 17: 423–426. Romero-Viana L, Julià R, Camacho A, Vicente E, Miracle MR. 2008. Climate signal in varve thickness: Lake La Cruz (Spain), a case study. Journal of Paleolimnology 40: 703–714. Rosenthal Y, Dahan M, Shemesh A. 2000 Southern Ocean contributions to glacial-interglacial changes of atmospheric pCO_{2}: An assessment of carbon isotope records in diatoms. Paleoceanography 15: 65–75. Rosqvist GC, Rietti-Shati M, Shemesh A. 1999. Late glacial to middle Holocene climatic record of lacustrine biogenic silica oxygen isotopes from a Southern Ocean island. Geology 27: 967–970. Rosqvist GC, Jonsson C, Yam R, Karlen W, Shemesh A. 2004. Diatom oxygen isotopes in pro-glacial lake sediments from northern Sweden: a 5000 year record of atmospheric circulation. Quaternary Science Reviews 23: 851– 859. Round FE, Crawford RM, Mann DG. 1990. The Diatoms, Biology & Morphology of the Genera. Cambridge University Press: Cambridge; 747. Rowe HD, Dunbar RB, Mucciarone DA, Seltzer GO, Baker PA, Fritz S. 2002. Insolation, moisture balance and climate change on the South American Altiplano since the Last Glacial Maximum. Climate Change 52: 175–199. Rowe HD, Guilderson TP, Dunbar RB, Southon JR, Seltzer GO, Mucciarone DA, Fritz SC, Baker PA. 2003. Late Quaternary lakelevel changes constrained by radiocarbon and stable isotope studies on sediment cores from Lake Titicaca, South America. Global and Planetary Change 38: 273–290. Ruhlemann C, Mulitza S, Muller PJ, Wefer G, Zahn R. 1999. Warming of the tropical Atlantic Ocean and slowdown of thermohaline circulation during the last deglaciation. Nature 402: 511–514. Saade A, Bowler C. 2009. Molecular tools for discovering the secrets of diatoms. Bioscience 59: 757–765. Sáez A, Cabrera L. 2002. Sedimentological and palaeohydrological responses to tectonics and climate in a small, closed, lacustrine system: Oligocene As Pontes Basin (Spain). Sedimentology 49: 1073–1094. Sáez A, Valero-Garcés BL, Moreno A, Bao R, Pueyo JJ, González-Sampériz P, Giralt S, Taberner C, Herrera C, Gibert RO. 2007. Volcanic controls on lacustrine sedimentation: the late Quaternary depositional evolution of lake Chungará (northern Chile). Sedimentology 54: 1191–1222. Sarnthein M, Winn K, Jung SJA, Duplessy JA, Labeyrie L, Erlenkeuser H, Ganssen G. 1994. Changes in east Atlantic deepwater circulation over the last 30,000 years: Eight time slice reconstructions. Paleoceanography 9: 209–267. Saulo AC, Nicolini M, Chou SC. 2000. Model characterization of the South American low-level flow during the 1997–98 spring– summer season. Climate Dynamics 16: 867–881. Schauble EA. 2004. Applying stable isotope fractionation theory to new systems. In Geochemistry of Non-Traditional Stable Isotopes, Johnson CM, Beard BL, Albarède F. (Eds.). Mineralogical Society of America Reviews in Mineralogy & Geochemistry 55: 65–111. Schelske CL, Hodell DA. 1991. Recent changes in productivity and climate of Lake Ontario detected by isotopic analysis of sediments. Limnology and Oceanography 36: 961–975. Schleser GH, Lücke A, Moschen R, Rings A. 2001. Separation of diatoms from sediment and oxygen isotope extraction from their siliceous valves — a new approach. Terra Nostra: Schriften der Alfred-Wegener-Stiftung, 6th Workshop of the European Lake Drilling Programme; 187–191. Schmidt M, Botz R, Rickert D, Bohrmann G, Hall SR, Mann S. 2001. Oxygen isotopes of marine diatoms and relations to opal- A maturation. Geochimica et Cosmochimica Acta 65: 201–211. Schneider N, Cornuelle BD. 2005. The forcing of the Pacific Decadal Oscillation. Journal of Climate 18: 4355–4373. Schneider-Mor A, Yam R, Bianchi C, Kunz-Pirrung M, Gersonde R, Shemesh A. 2005. Diatom stable isotopes, sea ice presence and sea surface temperature records of the past 640 ka in the Atlantic sector of the Southern Ocean. Geophysical Research Letters 32: L10704. Scholz CA, Johnson TC, McGill JW. 1993. Deltaic sedimentation in a rift valley lake: new seismic reflection data from Lake Malawi (Nyasa), East Africa. Geology 21: 395–398. Schwalb A. 2003. Lacustrine ostracods as stable isotope recorders of late-glacial and Holocene environmental dynamics and climate. Journal of Paleolimnology 29: 265–351. Schwalb A, Dean WE. 2002. Reconstruction of hydrological changes and response to effective moisture variations from north- central USA lake sediments. Quaternary Science Reviews 21: 1541–1554. Schwalb A, Burns SJ, Kelts K. 1999. Holocene environments from stable isotope stratigraphy of ostracods and authigenic carbonate in Chilean Altiplano lakes. Palaeogeography, Palaeoclimatology, Palaeoecology 148: 153–168. Seltzer G, Rodbell D, Burns S. 2000. Isotopic evidence for late Quaternary climatic change in tropical South America. Geology 28: 35–38. Servant M, Servant-Vildary S. 2003. Holocene precipitation and atmospheric changes inferred from river paleowetlands in the Bolivian Andes. Palaeogeography, Palaeoclimatology, Palaeoecology 194: 187–206. Shackleton NJ, Opdyke ND. 1973. Oxygen isotope and palaeomagnetic stratigraphy of equatorial Pacific core V28-238: oxygen isotope temperatures and ice volumes on a 105 and 106 year scale. Quaternary Research 3: 39–55. Shackleton NJ, Hall MA, Pate D. 1995. Pliocene stable isotope stratigraphy of ODP Site 846. In Proceedings ODP, Pisias NG, Mayer LA, Janecek TR, Palmer-Julson A, van Andel TH. (Eds.). Scientific Results 138: College Station, TX. Shemesh A, Charles CD, Fairbanks RG. 1992. Oxygen isotopes in biogenic silica: global changes in ocean temperature and isotopic composition. Science 256: 1434–1436. Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 120 Shemesh A, Macko SA, Charles CD, Rau GH. 1993. Isotopic evidence for reduced productivity in the glacial Southern-Ocean. Science 262: 407–410. Shemesh A, Burckle LH, Hays JD. 1995. Late Pleistocene oxygen-isotope records of biogenic silica from the Atlantic Sector of the Southern-Ocean. Paleoceanography 10: 179–196. Shemesh A, Rosqvist G, Rietti-Shati M, Rubensdotter L, Bigler C, Yam R, Karlen W. 2001. Holocene climatic change in Swedish Lapland inferred from an oxygen-isotope record of lacustrine biogenic silica. Holocene 11: 447–454. Shemesh A, Hodell D, Crosta X, Kanfoush S, Charles C, Guilderson T. 2002. Sequence of events during the last deglaciation in Southern Ocean sediments and Antarctic ice cores. Paleoceanography 17: 1056–1062. Shunk AJ, Driese SG, Dunbar JA. (2009). Late Tertiary paleoclimatic interpretation from lacustrine rhythmites in the Gray Fossil Site, northeastern Tennessee, USA. Journal of Paleolimnology 42: 11–24. Sicko-Goad LM, Schelske CL, Stoermer EF. 1984. Estimation of intracellular carbon and silica content of diatoms from natural assemblages using morphometric techniques. Limnology and Oceanography 29: 1170–1178. Sicko-Goad L, Stoermer EF, Kociolek JP. 1989. Diatom resting cell rejuvenation and formation: time course, species records and distribution. Journal of Plankton Research 11: 375–389. Simola H. 1992. Structural elements in varved lake sediments. Geological survey of Finland, special paper 14: 5–9. Simola H. 2007. Diatom records. Freshwater laminated sediments. In Encyclopedia of Quaternary Science, Elias SA. (Ed.).Elsevier: Amsterdam; 541–548. Singer AJ, Shemesh A. 1995. Climatically linked carbon-isotope variation during the past 430,000 years in Southern-Ocean sediments. Paleoceanography 10: 171–177. Skinner LC, Shackleton NJ, Elderfield H. 2003. Millennial-scale variability of deep-water temperature and δ18O dw indicating deepwater source variations in the Northeast Atlantic, 0–34 cal. ka BP. Geochemistry, Geophysics, Geosystems—G3: 4. Smetacek VS. 1985. Role of sinking in diatom life-history cycles: ecological, evolutionary and geological significance. Marine Biology 84: 239–251. Smith ND, Ashley GM. 1985. Proglacial lacustrine environment. In Glacial Sedimentary Enuironments, Ashley GM, Shaw J, Smith ND. (Eds.). Society of Economic Palaeontologists and Mineralogists Short Course Notes: Tulsa. Smol JP. 2008. Pollution of Lakes an Rivers. A Palaeoenvironmental Perspective. Blackwell Publishing: Malden. Solíz C, Villalba R, Argollo J, Morales MS, Christie DA, Moya J, Pacajes J. 2009. Spatio-temporal variations in Polylepis tarapacana radial growth across the Bolivian Altiplano during the 20th Century. Palaeogeography, Palaeoclimatology, Palaeoecology 281: 296–308 Sorhannus U. 2007. A nuclear-encoded small-subunit ribosomal RNA timescale for diatom evolution. Marine Micropaleontology 65: 1–12. Sterken M, Verleyen E, Sabbe K, Terryn G, Charlet F, Bertrand S, Boës X, Fagel N, De Batist M, Vyverman W. 2008. Late Quaternary climatic changes in southern Chile, as recorded in a diatom sequence of Lago Puyehue (40°402 S). Journal of Paleolimnology 39: 219–235. Stockwell RG, Mansinha L, Lowe RP. 1996. Localization of the complex spectrum: the S transform. IEEE Transactions on Signal Processing 44: 998–1001. Stoermer EF, Julius ML. 2003. Centric diatoms. In Freshwater Algae of North America, Wehr JD, Sheath RG. (eds.). San Diego: Elsevier Science (USA); 559-594. Stoermer EF, Smol JP. 1999. The Diatoms: Applications for the Environmental and Earth Sciences. Cambridge: Cambridge University Press. Strecker MR, Alonso RN, Bookhagen B, Carrapa B, Hilley GE, Sobel ER, Trauth MH. 2007. Tectonics and climate of the southern central Andes. Annual Review of Earth and Planetary Sciences 35: 747–787. Street-Perrott FA, Harrison SP. 1985. Lake levels and climate reconstruction. In Paleoclimate analysis and modeling, Hecht AD (Ed.). Wiley: New York; 291–340. Stuiver M. 1968. Oxygen-18 content of atmospheric precipitation during the last 11,000 Years in Great Lakes region. Science 162: 994–997. Sumper M, Kröger N. 2004. Silica formation in Diatoms: the function of long-chain polyamines and silaffins. Journal of Materials Chemistry 14: 2059–2065. Swann GEA, Leng MJ. 2009. A review of diatom δ18O in palaeoceanography. Quaternary Science Reviews 28: 384–398. Swann GEA, Leng MJ, Sloane HJ, Maslin MA. 2008. Isotope offsets in marine diatom δ18O over the last 200 ka. Journal of Quaternary Science 23: 389–400. Swann G, Leng MJ, Juschus O, Melles M, Brigham-Grette J, Sloane HJ. 2010. A combined oxygen and silicon diatom isotope record of Late Quaternary change in Lake El’gygytgyn, North East Siberia. Quaternary Science Reviews 29: 774–786. Swihart GH, McBay EH, Smith DH, Siefke JW. 1996. A boron isotopic study of a mineralogically zoned lacustrine borate deposit: the Kramer deposit, California, U.S.A. Chemical Geology 127: 241–250. Sylvestre F. 2009. Moisture Pattern During the Last Glacial Maximum in South America. In Past Climate Variability in South America and Surrounding Regions, Vimeux F, Sylvestre F, Khodri M. (eds.). Springer: Dordrecht, Netherlands; 3–27. Sylvestre F, Servant M, Servant-Vildary S, Causse C, Fournier M, Ybert JP. 1999. Lake-level chronology on the southern Bolivian Altiplano (18º–23ºS) during late-glacial time and the early Holocene. Quaternary Research 51: 54–66. Talbot MR. 1988. The origins of the lacustrine oil source rocks: Evidence from the lakes of tropical Africa. In Lacustrine Petroleum Source Rocks, Fleet AJ, Kelts K, Talbot MR. (Eds.). Oxford/Boston: Blackwell Scientific Publications 40; 29–43. Talbot MR. 2001. Nitrogen isotopes in palaeolimnology. In Tracking Environmental Change Using Lake Sediments, Physical and Geochemical Techniques, vol.2. Last WM, Smol JP. (Eds.). Kluwer Academic Publishers: Dordrecht, The Netherlands; 401–439 Talbot MR, Allen PA. 1996. Lakes. In Sedimentary Environments, Reading HG (Ed.). Blackwell: Oxford; 83–124. Talbot MR, Lærdal T. 2000. The late Pleistocene-Holocene palaeolimnology of Lake Victoria, East Africa, based upon elemental and isotopic analyses of sedimentary organic matter. Journal of Paleolimnology 23: 141–164. Bibliography 121 Talbot MR, Kelts K. 1990. Paleolimnological signatures from carbon and oxygen isotopic ratios in carbonates from organic carbon-rich lacustrine sediments. In Lacustrine Basin Exploration: Case Studies and Modern Analogs, Katz BJ. (Ed.). AAPG Memoir 50: 88–112. Talling JF. 1976. Depletion of carbon-dioxide from lake water by phytoplankton. Journal of Ecology 64: 79–121. Tapia PM, Fritz SC, Baker PA, Seltzer GO, Dunbar RB. 2003. A late Quaternary diatom record of tropical climatic history from Lake Titicaca (Bolivia/Peru). Palaeogeography, Palaeoclimatology, Palaeoecology 194: 139–164. Tassara A, Yañez G. 2003. Relación entre el espesor elástico de la litosfera y la segmentación tectónica del margen andino (15- 47ºS). Revista Geológica de Chile 30: 159–186. Teranes JN, McKenzie JA, Bernasconi SM, Lotter AF, Sturm M. 1999. A study of oxygen isotopic fractionation during bio- induced calcite precipitation in eutrophic Baldeggersee, Switzerland. Geochimica et Cosmochimica Acta 63: 1981–1999. Teranes JL, Bernasconi SM. 2000. The record of nitrate utilization and productivity limitation provided by 15N values in lake organic matter – A study of sediment trap and core sediments from Baldeggersee, Switzerland. Limnology and Oceanography 45: 801–813. Theissen KM, Dunbar RB, Rowe HD, Mucciarone DA. 2008. Multidecadal- to century-scale arid episodes on the Northern Altiplano during the middle Holocene. Palaeogeography, Palaeoclimatology, Palaeoecology 257: 361–376. Theriot E. 2001. Diatoms. In Encyclopedia of Life Sciences. John Wiey and Sons: Chichester. Thompson DJ. 1982. Spectrum estimation and harmonic analysis. Proceedings of the Institute of Electrical and Electronics Engineers 70: 1055–1096. Thompson L, Mosley-Thompson E, Davis ME, Lin PN, Henderson KA, Cole-Dai J, Bolzan JF, Liu KB. 1995. Late glacial stage and Holocene tropical ice core records from Huascaran, Peru. Science 269: 46–50. Thompson LG, Davis ME, Mosley-Thompson E, Sowers TA, Henderson KA, Zagorodnov VS, Lin PN, Mikhalenko VN, Campen RK, Bolzan JF, Cole-Dai J, Francou B. 1998. A 25,000-year tropical climate history from Bolivian ice cores. Science 282: 1858–1864. Thompson LG, Davis ME, Mosley-Thompson E, Lin PN, Henderson KA, Mashiotta TA. 2005. Tropical ice core records: evidence for asynchronous glaciation on Milankovitch timescales. Journal of Quaternary Science 20: 723–734. Thompson JB, Schultze-Lam S, Berveridge TJ, Des Marais DJ. 1997. Whiting events: biogenic origin due to the photosynthetic activity of cyanobacterial picoplankton. Limnology and Oceanography 42: 133–141. Thornton DCO. 2002. Diatom aggregation in the sea: mechanisms and ecological implications. European Journal of Phycology 37: 149–161. Tiedemann R, Sarnthein M, Shackleton NJ. 1994. Astronomic timescale for the Pliocene Atlantic and 18O and dust flux records of Ocean Drilling Program site 659. Paleoceanography 9: 619–638. Tiercelin JJ. 1991. Natural resources in the lacustrine facies of the Cenozoic rift basins of east Africa. In Lacustrine facies analysis, Anadón P, Cabrera L, Kelts K (Ed.). Special Publication Number 13 of the International Association of Sedimentologists. Blackwell: Oxford; 1–37. Tilman D, Kilham SS, Kilham P. 1982. Phytoplankton community ecology: the role of limiting nutrients. Annual Review of Ecology and Systematics 13: 349–372. Timmermann A, Okumura Y, An SI, Clement A, Dong B, Guilyardi E, Hu A, Jungclaus JH, Renold M, Stocker TF, Stouffer RJ, Sutton R, Xie SP, Yin J. 2007. The influence of a weakening of the Atlantic meridional overturning circulation on ENSO. Journal of Climate 20: 4899–4919. Torrence C, Compo GP. 1998. A practical guide to wavelet analysis. Bulletin of the American Meteorological Society 79: 61–78. Tranvik LJ. 1988. Availability of dissolved organic carbon for planktonic bacteria in oligotrophic lakes of differing humic content. Microbial Ecology 16: 311–322. Tranvik L, Downing JA, Cotner JB, Loiselle SA, Striegl RG, Ballatore TJ, Dillon P, Finlay K, Fortino K, Knoll LB, Kortelainen PL, Kutser T, Larsen S, Laurion I, Leech DM, McCallister SL, McKnight DM, Melack JM, Overholt E, Porter JA, Prairie Y, Renwick WH, Roland F, Sherman BS, Schindler DW, Sobek S, Tremblay A, Vanni MJ, Verschoor AM, von Wachenfeldt E, Weyhenmeyer GA. 2009. Lakes and reservoirs as regulators of carbon cycling and climate. Limnology and Oceanography 54: 2298–2314. Tyler JJ, Leng MJ, Sloane HJ, Sachse D, Gleixner G. 2008. Oxygen isotope ratios of sedimentary biogenic silica reflect the European transcontinental climate gradient. Journal of Quaternary Science 23: 341–350. Urey HC. 1947. The thermodynamic properties of isotopic substances. Journal of the Chemical Society 1947: 562–581. Urey HC, Lowenstam HA, Epstein S, McKinney CR. 1951. Measurement of palaeotemperatures and temperatures of the upper cretaceous of England, Denmark and southeastern United States. Geological Society of America Bulletin 62: 399–416. Valero-Garcés BL, Grosjean M, Schwalb A, Geyn M, Messerli B, Kelts K. 1996. Limnogeology of Laguna Miscanti: evidence for mid to late Holocene moisture changes in the Atacama altiplano. Journal of Paleolimnology 16: 1–21. Valero-Garcés BL, Grosjean M, Kelts K, Schreir H, Messerli B. 1999. Holocene lacustrine deposition in the Atacama Altiplano: facies models, climate and tectonic forcing. Palaeogeography, Palaeoclimatology, Palaeoecology 151: 101–125. Valero-Garces BL, Grosjean M, Schwalb A, Schreir H, Kelts K, Messerli B. 2000. Late Quaternary lacustrine deposition in the Chilean Altiplano (18º–28ºS). In Lake Basins through Space and Time, Gierlowski-Kordesch E, Kelts K (eds). American Association of Petroleum Geologists Studies in Geology 46: 625–636. Valero-Garcés BL, Delgado-Huertas A, Navas A, Edwards L, Schwalb A, Ratto N. 2003. Patterns of regional hydrological variability in central-southern Altiplano (18º-26ºS) lakes during the last 500 years. Palaeogeography, Palaeoclimatology, Palaeoecology 194: 319– 338. Vanormelingen P, Verleyen E, Vyverman W. 2008. The diversity and distribution of diatoms: from cosmopolitanism to narrow endemism. Biodiversity and Conservation 17: 393-405. Vaulot D. 2006. Phytoplankton. Encyclopedia of Life Sciences. Chichester: John Wiley & Sons. Velasco VM, Mendoza B. 2008. Assessing the relationship between solar activity and some large scale climatic phenomena. Advances in Space Research 42: 866–878. Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 122 Vera C, Higgins W, Amador J, Ambrizzi T, Garreaud R, Gochis D, Gutzler D, Lettenmaier D, Marengo J, Mechoso CR, Nogues- Paegle J, Silva Diaz PL, Zhang C. 2006. Towards a unified view of the American Monsoon System. Journal of Climate 19: 4977–5000. Villalba R, D’Arrigo RD, Cook ER, Wiles G, Jacoby GC. 2001. Decadal-scale climatic variability along the extra-tropical western coast of the Americas over past centuries inferred from tree-ring records. In Interhemispheric Climate Linkages, Markgraf V. (Ed.). Cambridge University Press: Cambridge, UK. Villalba R, Grosjean M, Kiefer T. 2009. Long-term multi-proxy climate reconstructions and dynamics in South America (LOTRED- SA): State of the art and perspectives. Palaeogeography, Palaeoclimatology, Palaeoecology 281: 175–179. Villwock W, Kies L, Thiedig F, Thomann R. 1985. Geologish-ökologishe Untersuchungen am Lago Chungará/ Nord Chile. Zielsetzungen und erste Ergebnisse 9: IDESIA, Chile; 21–34. Volcani BE. 1981. Cell wall formation in diatoms: morphology and biochemistry. In Silicon and Siliceous Structures in Biological Systems, Simpson X, Volcani BE. (Eds.). Springer Verlag: New York; 157–200. Vuille M. 1999. Atmospheric circulation over the Bolivian altiplano during dry and wet periods and extreme phases of the Southern Oscillation. International Journal of Climatology 19: 1579–1600. Vuille M, Keimig F. 2004. Interannual variability of summertime convective cloudiness and precipitation in the central Andes derived from ISCCP-B3 data. Journal of Climate 17: 3334–3348. Vuille M, Werner M. 2005. Stable isotopes in precipitation recording South American summer monsoon and ENSO variability: Observations and model results. Climate Dynamics 25: 401–413. Vuille M, Bradley RS, Keimig F. 2000. Interannual climate variability in the Central Andes and its relation to tropical Pacific and Atlantic forcing. Journal of Geophysical Research-Atmospheres 105: 12447–12460. Wanner H, Beer J, Bütikofer J, Crowley TJ, Cubasch U, Flückiger J, Goosse H, Grosjean M, Joos F, Kaplan JO, Küttel M, Müller SA, Prentice C, Solomina O, Stocker TF, Tarasov P, Wagner M, Widmann M. 200.: Mid- to Late Holocene climate change: an overview. Quaternary Science Reviews 27: 1791–1828. Weng C, Bush MB, Curtis JH, Kolata AL, Dillehay TD, Binford MW. 2006. Deglaciation and Holocene climate change in the western Peruvian Andes. Quaternary Research 66: 87–96. Werner D. 1977. The Biology of Diatoms. University California Press: Berkeley. Wetzel RG. 2001. Limnology. Lake and rivers ecosystems. 3rd ed. Academic Press: San Diego, USA. Willén E. 1991. Planktonic diatoms - an ecological review. Archiv für Hydrobiologie: 69–106. Whitman D, Isacks BL, Kay SM. 1996. Lithospheric structure and along-strike segmentation of the central Andean Plateau: Seismic Q, magmatism, flexture, topography and tectonics. Tectonophysics 259: 29-40. Wolfe BB, Aravena R, Abbott MB, Seltzer GO, Gibson JJ. 2001. Reconstruction of paleohydrology and paleohumidity from oxygen isotope records in the Bolivian Andes. Palaeogeography, Palaeoclimatology, Palaeoecology 176: 177–192. Wörner G, Harmon RS, Davidson J, Moorbath S, Turner DL, McMillan N, Nye C, López-Escobar L, Moreno H. 1988. The Nevados de Payachata volcanic region (18°S/69°W, N. Chile): I. Geological, geochemical, and isotopic observations. Bulletin of Volcanology 50: 287–303. Wörner G, Hammerschmidt K, Henjes-Kunst F, Wilke H. 2000. Geochronology (40Ar/39Ar, K-Ar and He-exposure ages) of Cenozoic magmatic rocks from northern Chile (18–22ºS): implications for magmatism and tectonic evolution of the central Andes. Revista Geológica de Chile 27: 205–240. Yansa CH, Dean WE, Murphy EC. 2007. Late Quaternary paleoenvironments of an ephemeral wetland in North Dakota, USA: relative interactions of groundwater hydrology and climate change. Journal of Paleolimnology 38: 441–457. 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. Zech R, Smith J, Kaplan MR. 2009 Chronologies of the Last Glacial Maximum and its Termination in the Andes (~10–55ºS) Based on Surface Exposure Dating. In Past Climate Variability in South America and Surrounding Regions, Vimeux F, Sylvestre F, Khodri M. (eds.). Springer: Dordrecht, Netherlands; 61–88. Zhou J, Lau KM. 1998. Does a Monsoon Climate Exist over South America?. Journal of Climate 11: 1020–1040. Zolitschka B. 1992. Climatic change evidence and lacustrine varves from maar lakes, Germany. Climate Dynamics 6: 229–232. Zolitschka B, Negendank JFW. 1996. Sedimentology, dating and palaeoclimatic interpretation of a 76.3 ka record from Lago Grande di Monticchio, southern Italy. Quaternary Science Reviews 15: 101–112. Zolitschka B, Brauer A, Negendank JFW, Stockhausen H, Lang A. 2000. Annually dated late Weichselian continental palaeoclimate record from the Eifel, Germany. Geology 28: 783–786. Bibliography 123 Appendices Appendix A Glossary Symbols α Fractionation factor δ Isotope composition δD Stable hydrogen isotopes; the ratio of 2H to 1H in a sample relative to a standard δ13C Stable carbon isotopes; the ratio of 13C to 12C in a sample relative to a standard δ13C carbonate Carbon isotope composition of carbonates δ13C diatom Carbon isotope composition of diatom organic cell wall inclusions δ13C DIC Carbon isotope composition of dissolved inorganic carbon δ13C DOC Carbon isotope composition of dissolved organic carbon δ13C bulk Carbon isotope composition of bulk organic matter δ15N Stable nitrogen isotopes; the ratio of 15N to 14N in a sample relative to a standard δ15N bulk Nitrogen isotope composition of bulk organic matter δ15N diatom Nitrogen isotope composition of diatom organic cell wall inclusions δ18O Stable oxygen isotopes; the ratio of 18O to 16O in a sample relative to a standard δ18O carbonate Oxygen isotope composition of carbonates δ18O diatom Oxygen isotope composition of diatom silica δ18O lakewater Oxygen isotope composition of water of the lake δ18O precipitation Oxygen isotope composition of rainfall δ30Si Stable silicon isotopes; the ratio of 30Si to 28Si in a sample relative to a standard δ30Si diatom Silicon isotope composition of diatom silica ε Isotope enrichment factors σ Probability distribution Elements H Hydrogen C Carbon N Nitrogen O Oxygen F Fluorine Na Sodium Mg Magnesium Si Silicon P Phosphorus S Sulphur Cl Chlorine K Potassium Ca Calcium Fe Iron Br Bromine Pt Platinum Chemical Formulas CO 2 Carbon dioxide CO 2(aq) CO 2 dissolved in water CH 4 Methane O 2 Oxygen in its molecular form OH- Hydroxide ion H 2 O Distilled water H 2 O 2 Hydrogen peroxide HCO 3 - Bicarbonate ion HCl Hydrochloric acid HNO 3 Nitric acid N 2 Nitrogen gas NO 2 - Nitrite ion Glossary 127 NO 3 - Nitrate ion NH 4 + Ammonium cation PO 4 3- Ortophosphate ion SiO 2 Silicon dioxide or silica Si(OH) 4 Silicic acid SO 4 - Sulfate ion CaCO 3 Calcium carbonate ClF 3 Chlorine trifluoride BrF 5 Bromine pentafluoride Acronyms and abrevations AAO Antarctic Oscillation AMS Accelerator Mass Spectrometry B-A Bølling-Allerød chronozone BFC Diatomite Control Sample BP Before Present BSi Biogenic Silica CSIC Consejo Superior de Investigaciones Científicas DIC Dissolved Inorganic Carbon DOC Dissolved Organic Carbon EL Evaporation Line ENSO El Niño-Southern Oscillation GISP Greenland Ice Sheet Precipitation GMWL Globale Meteroric Water Line IAEA International Atomic Energy Agency IBiS Isotopes in Biogenic Silica ICP Inductively Coupled Plasma ICTJA Institut de Ciències de la Terra-'Jaume Almera' IPE Instituto Pirenaico de Ecología IRMS Isotope ratio mass spectrometry ITCZ InterTropical Convergence Zone LEC Lancaster Environmental Center LGM Last Glacial Maximun LML Local Meteoric Line LRC Limnological Research Center MSCL Multi-Sensor Core Logger MTM Multi-Taper Method NAO North Atlantic Oscillation NBS National Bureau of Standards; National Institute of Standards and Technology NERC Natural Environment Research Council NIGL NERC Isotope Geosciences Laboratory OM Organic Matter OES Optical Emission Spectroscopy PCA Principal Component Analysis PDO Pacific Decadal Oscillation P/E Precipitation/Evaporation ratio PhD Doctor of Philosophy RDA Redundancy Analysis RML Regional Meteoric Line SASM South American Summer Monsoon SDV Silica Deposition Vesicle SEM Scanning Electron Microscopes SLAP Standard Light Antarctic Precipitation SPCZ South Pacific Convergence Zone SPLITT Gravitational split-flow lateral-transport thin SST Sea Surface Temperature Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 128 SWF Stepwise Fluorination TC Total Carbon TDS Total Dissolved Solids TF Time-Frequency TOC Total Organic Carbon UB Universitat de Barcelona UCN Universidad Católica del Norte UDC Universidade da Coruña UK United Kingdom USA United States of America VPDB Vienna Peedee Belemnite VSMOW Vienna Standard Mean Ocean Water XRD X-Ray Diffraction XRF X-Ray Fluorescence YD Younger Dryas chronozone 3D Three Dimensions aff affinis asl above sea level ca circa cal calibrated m/e mass-to-charge ratio n number of samples na not available sd standard deviation spp species Units ppm parts per million ‰ per mil; one part per thousand % percentage wt% weight percentage cps counts per second μm micrometre mm millimetre cm centimetre m metre km kilometre km2 square kilometre ml millilitre Hm3 cubic hectometre mm yr-1 millimetre per year l s-1 litre per second m3yr-1 cubic metre per year l s-1 litre per second g l-1 grams per litre mg cm-2 yr-1 milligrams per square metres and year million valves g-1 million valves per gram º degree ºC degree Celsius h hour Myrs million years μS cm-1 microsiemens per centimetre Wm-2 Watts per square metres Hz hertz Glossary 129 Appendix B Final isotope data Sample Identifier Extraction Number (FF____) Yield % del 18 O V- SMOW Raw Run Correction del 18 O V-SMOW Run Corrected Within Run Reproducibility (1 sigma) Between Run Reproducibility (1 sigma) Comments TA-02 FF24958 71 +41,56 -1,27 40,3 TA-04 FF24957 71,5 +41,34 -1,27 40,1 TA-05 FF24960 71 +42,21 -1,27 40,9 TA-08 FF24961 62,9 +42,41 -1,27 41,1 TA-09 FF24962 72,1 +42,62 -1,27 41,3 TA-12 FF24963 55,9 +49,64 -1,27 48,4 SAMPLE WET- REPEATED TA-12/X FF24968 68 +43,50 -1,56 41,9 TA-13 FF24966 71,6 +41,12 -1,27 39,8 TA-15 FF24969 69,4 +42,35 -1,56 40,8 TA-18 FF24970 70 +43,17 -1,56 41,6 TA-20 FF24972 62 +41,64 -1,56 40,1 TA-22 FF24944 65,7 +40,25 -1,47 38,8 DROPPED - AIR IN VESSEL ANALYSIS POOR TA-22/X FF24964 71,5 +41,38 -1,27 40,1 TA-23 FF24946 66,1 +41,91 -1,47 40,4 TA-25 FF24947 67,8 +42,24 -1,47 40,8 TA-27 FF24948 66,2 +41,67 -1,47 40,2 TA-31 FF24949 69,2 +41,08 -1,47 39,6 TA-34 FF24951 68,8 +41,49 -1,47 40,0 TA-36 FF24952 64,7 +41,06 -1,47 39,6 TA-37 FF24954 68,5 +40,56 -1,47 39,1 TA-40 FF24973 67,0 +42,12 -1,56 40,6 TA-41 FF24974 74,4 +40,82 -1,56 39,3 TA-41/B FF24978 59,7 +40,94 -1,56 39,4 MEAN 39,3 0,1 TB-02 FF25072 64,7 +40,93 -1,31 39,6 TB-04 FF25073 69,0 +41,00 -1,31 39,7 TB-06 FF25074 69,1 +40,26 -1,31 38,9 TB-07 FF25080 68,3 +40,86 -1,72 39,1 TB-09 FF25081 67,0 +39,70 -1,72 38,0 TB-09/B FF25086 69,4 +39,65 -1,72 37,9 MEAN 38,0 0,03 TB-12 FF25083 73,6 +40,73 -1,72 39,0 TB-14 FF25084 64,7 +41,06 -1,72 39,3 TB-16 FF25085 +40,73 -1,72 39,0 TB-20 FF25088 66,7 +39,19 -1,43 37,8 TB-21 FF25089 71,8 +40,37 -1,43 38,9 TB-24 FF25090 72,2 +40,46 -1,43 39,0 TB-25 FF25092 70,3 +40,11 -1,43 38,7 TB-25/B FF25096 72,6 +39,93 -1,43 38,5 MEAN 38,6 0,13 TB-27 FF25093 70,6 +40,12 -1,43 38,7 TB-29 FF25106 71,0 +37,50 -1,52 36,0 TB-32 FF25097 73,7 +38,10 -1,43 36,7 TB-33 FF25098 73,9 +39,11 -1,43 37,7 TB-36 FF25094 69,0 +40,91 -1,43 39,5 TC-001 FF24975 71,7 +40,26 -1,56 38,70 TC-004 FF24976 69,7 +38,32 -1,56 36,76 TC-005 FF24985 73,6 +38,69 -1,61 37,08 TC-005/X FF24993 72,7 +38,30 -1,21 37,09 MEAN 37,08 0,01 TC-007 FF24981 65,8 +39,62 -1,61 38,01 TC-009 FF24982 73,1 +38,55 -1,61 36,94 Isotope data 133 TC-010 FF24983 72,7 +39,58 -1,61 37,97 TC-011 FF25828 72,7 +37,94 -1,07 36,87 TC-012 FF24986 60,2 +37,24 -1,61 35,63 TC-013 FF24987 71,6 +39,96 -1,61 38,35 TC-014 FF25829 73,0 +37,34 -1,07 36,27 TC-015 FF25830 67,2 +38,93 -1,07 37,86 TC-015/B FF25835 72,0 +39,22 -1,07 38,15 MEAN 38,01 0,21 TC-016 FF24988 63,6 +39,46 -1,61 37,85 TC-017 FF24989 68,9 +39,54 -1,61 37,93 TC-018 FF25832 67,6 +40,74 -1,07 39,67 TC-019 FF24994 73,5 +39,56 -1,21 38,35 TC-020 FF25833 71,7 +38,44 -1,07 37,37 TC-021 FF24995 73,2 +40,00 -1,21 38,79 TC-021/X FF25834 71,5 +38,59 -1,07 37,52 TC-021/X FF26278 61,8 +38,62 -1,09 37,53 MEAN 37,53 0,00 TC-022 FF25836 69,6 +39,14 -1,07 38,07 TC-023 FF25838 71,7 +38,36 -1,07 37,29 TC-024 FF24996 75,8 +39,86 -1,21 38,65 TC-025 FF25840 71,7 +38,38 -1,02 37,36 TC-026 FF24998 77,2 +38,12 -1,21 36,91 TC-027 FF24999 71,7 +39,00 -1,21 37,79 TC-028 FF25841 70,2 +37,88 -1,02 36,86 TC-029 FF25000 77,3 +39,56 -1,21 38,35 TC-030 FF25842 74,4 +39,75 -1,02 38,73 TC-031 FF25843 74,9 +39,51 -1,02 38,49 TC-032 FF25001 72,0 +39,56 -1,21 38,35 TC-032 FF25845 70,5 +38,92 -1,02 37,90 TC-032/B FF25849 74,6 +39,08 -1,02 38,06 MEAN 38,10 0,23 TC-033 FF25002 77,5 +38,15 -1,21 36,94 TC-034 FF25846 77,3 +38,60 -1,02 37,58 TC-035 FF25004 75,0 +37,85 -1,33 36,52 TC-036 FF25847 75,2 +38,64 -1,02 37,62 TC-037 FF25006 68,7 +37,29 -1,33 35,96 TC-038 FF25848 77,4 +38,69 -1,02 37,67 TC-039 FF25007 69,1 +39,65 -1,33 38,32 TC-039/X FF25101 69,7 +39,52 -1,52 38,00 MEAN 38,16 0,22 TC-040 FF25853 72,7 +37,28 -0,77 36,51 TC-041 FF25854 64,7 +38,69 -0,77 37,92 TC-042 FF25008 71,5 +38,88 -1,33 37,55 TC-042/B FF25011 72,9 +38,14 -1,33 36,81 MEAN 37,18 0,53 TC-042/X FF26277 70,4 +36,87 -1,09 35,78 TC-042/XX FF26298 68,8 +36,42 -0,88 35,54 MEAN 35,66 0,17 TC-043 FF25009 70,6 +38,82 -1,33 37,49 TC-044 FF25855 65,6 +36,95 -0,77 36,18 TC-044/B FF25858 65,1 +37,27 -0,77 36,50 MEAN 36,34 0,11 TC-046 FF25012 71,4 +36,86 -1,33 35,53 TC-046 FF258566 64,2 +36,17 -0,77 35,40 MEAN 35,46 0,09 TC-047/X FF26008 74,7 +37,54 -0,6 36,94 TC-047 FF25859 67,0 +37,94 -0,77 37,17 MEAN 37,05 0,16 TC-049 FF25014 74,6 +38,90 -1,33 37,57 TC-050 FF25860 ? +38,12 -0,77 37,35 TC-051 FF25861 71,0 +38,97 -0,77 38,20 TC-052 FF25016 70,6 +38,12 -1,5 36,62 TC-053 FF25862 69,4 +38,68 -0,77 37,91 TC-054 FF25017 70,1 +39,20 -1,5 37,70 TC-055 FF25864 77,9 +37,66 -0,36 37,30 TC-056 FF25019 68,8 +39,34 -1,5 37,84 TC-057 FF25020 68,4 +39,22 -1,5 37,72 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 134 Isotope data 135 TC-058 FF25021 68,8 +39,21 -1,5 37,71 TC-059 FF25022 67,1 +38,13 -1,5 36,63 TC-059/B FF25026 68,7 +38,70 -1,5 37,20 MEAN 36,92 0,40 TC-059/X FF26279 71,7 +38,71 -1,09 37,61 MEAN 37,15 0,49 TC-060 FF25024 72,5 +38,80 -1,5 37,30 TC-061 FF25025 68,5 +38,66 -1,5 37,16 TC-062 FF25866 74,3 +36,89 -0,36 36,53 TC-063 FF25028 76,7 +37,78 -1,28 36,50 TC-064 FF25867 71,9 +37,54 -0,36 37,18 TC-065 FF25868 75,3 +37,75 -0,36 37,39 TC-066 FF25029 73,0 +39,69 -1,28 38,41 TC-066/B FF25034 73,8 +40,18 -1,28 38,90 MEAN 38,65 0,35 TC-066 FF25869 79,0 +38,19 -0,36 37,83 MEAN 38,38 0,53 TC-067 FF25030 71,2 +39,44 -1,28 38,16 TC-068 FF25871 75,1 +38,25 -0,36 37,89 TC-068/B FF25874 74,0 +38,44 -0,36 38,08 MEAN 37,99 0,13 TC-069 FF25872 79,6 +36,27 -0,36 35,91 TC-069/X FF26009 ? +36,42 -0,6 35,82 MEAN 35,86 0,06 TC-070 FF25032 73,9 +38,26 -1,28 36,98 TC-071 FF25033 73,8 +39,98 -1,28 38,70 TC-072 FF26010 77,3 +35,66 -0,6 35,06 TC-073 FF25888 68,2 +38,21 -1,02 37,19 TC-074 FF25035 70,8 +39,77 -1,28 38,49 TC-075 FF25037 79,0 +39,28 -1,28 38,00 TC-076 FF25038 74,4 +39,29 -1,28 38,01 TC-076 FF25889 69,0 +39,56 -1,02 38,54 MEAN 38,28 0,38 TC-077 FF26046 75,9 +38,61 -0,8 37,81 TC-077/B FF26050 78,5 +38,48 -0,8 37,68 MEAN 37,75 0,09 TC-078 FF25040 74,2 +38,69 -1,39 37,30 TC-079 FF25892 69,8 +38,60 -1,02 37,58 TC-079/B FF25897 66,0 +38,85 -1,02 37,83 MEAN 37,70 0,17 TC-080 FF25041 73,7 +40,05 -1,39 38,66 TC-081 FF25053 68,8 +38,98 -1,39 37,59 TC-082 FF25893 67,1 +38,83 -1,02 37,81 TC-083 FF25043 70,6 +40,31 -1,39 38,92 TC-084 FF25894 72,3 +41,15 -1,02 40,13 TC-085 FF25695 68,9 +39,36 -1,02 38,34 TC-086 FF25045 74,0 +40,02 -1,39 38,63 BATCH#10 BFC DATA POOR USE AVERAGE RUN CORRECTION TC-086/B FF25050 67,2 +39,63 -1,39 38,24 MEAN 38,44 0,28 TC-087 FF24046 77,5 +40,30 -1,39 38,91 TC-088 FF25047 73,3 +39,41 -1,39 38,02 TC-089 FF25048 75,6 +39,29 -1,39 37,90 TC-090 FF25054 69,3 +40,58 -1,34 39,24 TC-091 FF25055 70,0 +40,14 -1,34 38,80 AVERAGE RUN CORRECTION = 28.88 - 30.22 = -1.34 TC-091/X FF25102 69,9 +40,03 -1,52 38,51 MEAN 38,66 0,20 TC-092 FF25056 64,5 +39,01 -1,34 37,67 TC-093 FF25057 69,1 +40,58 -1,34 39,24 TC-094 FF25898 68,4 +39,01 -1,02 37,99 TC-095 FF25900 67,0 +38,82 -0,82 38,00 TC-096 FF25058 66,4 +39,96 -1,34 38,62 TC-097 FF25076 67,8 +38,76 -1,72 37,04 TC-098 FF25901 71,7 +37,78 -0,82 36,96 TC-099 FF25903 69,2 +38,40 -0,82 37,58 TC-100 FF25077 67,1 +40,03 -1,72 38,31 TC-101 FF25079 68,2 +39,15 -1,72 37,43 TC-102 FF25905 70,1 +39,36 -0,82 38,54 TC-103 FF25906 72,1 +39,26 -0,82 38,44 MEAN 38,03 0,58 TC-103/X FF26274 67 +38,49 -1,09 37,39 MEAN 37,51 0,16 TC-104 FF25909 59,8 +38,83 -1,13 37,70 TC-104/X FF26012 74,2 +38,04 -0,6 37,44 MEAN 37,57 0,18 TC-105 FF25066 69,2 +39,81 -1,31 38,50 TC-106 FF26013 73,7 +37,99 -0,6 37,39 TC-107 FF25067 71,0 +39,02 -1,31 37,71 TC-107/B FF25070 67,1 +38,81 -1,31 37,50 MEAN 37,60 0,15 TC-108 FF25068 69,8 +32,38 -1,31 31,07 TC-108/X FF25103 71,5 +32,73 -1,52 31,21 MEAN 31,14 0,10 TC-109 FF25911 72,1 +36,47 -1,13 35,34 TC-110 FF25071 67,4 +37,57 -1,31 36,26 TC-110/X FF25105 67,5 +36,92 -1,52 35,40 MEAN 35,83 0,60 TC-110/X FF26276 68,8 +36,58 -1,09 35,49 MEAN 35,45 0,06 TC-111 FF26048 77,8 +37,85 -0,8 37,05 3-060 FF25993 76,8 +41,51 -0,77 40,7 3-070 FF25995 69,1 +40,87 -0,77 40,1 3-080 FF25996 74,7 +40,10 -0,77 39,3 3-090 FF25997 74,6 +40,28 -0,77 39,5 3-095 FF25944 69,0 +37,77 -1,42 36,3 3-100 FF25845 71,4 +38,08 -1,42 36,7 3-110 FF25946 74,7 +39,92 -1,42 38,5 3-110/B FF25949 66,6 +40,30 -1,42 38,9 MEAN 38,7 3-115 FF25848 76,3 +38,95 -1,42 37,5 3-120 FF25951 69,6 +40,24 -1,42 38,8 3-130 FF25952 71,7 +39,98 -1,42 38,6 3-135 FF25953 69,3 +40,82 -1,42 39,4 3-140 FF25954 68,5 +40,46 -1,42 39,0 3-145 FF25956 73,4 +40,46 -0,99 39,5 3-149 FF25957 74,1 +40,25 -0,99 39,3 4-001 FF25958 76,8 +40,59 -0,99 39,6 4-005 FF25959 74,2 +40,23 -0,99 39,2 4-010 FF25961 66,3 +40,70 -0,99 39,7 4-015 FF25962 69,6 +40,82 -0,99 39,8 4-020 FF25964 76,8 +41,25 -0,99 40,3 4-025 FF25971 77,2 +41,11 -0,99 40,1 4-025/B FF25976 76,8 +40,85 -0,99 39,9 MEAN 40,0 0,2 4-030 FF25972 78,4 +41,10 -0,99 40,1 4-040 FF25973 73,0 +41,07 -0,99 40,1 4-045 FF25974 74,1 +40,71 -0,99 39,7 4-050 FF25977 71,8 +39,30 -0,99 38,3 4-060 FF25978 74,0 +40,06 -0,99 39,1 4-070 FF25980 69,1 +38,86 -1,24 37,6 4-080 FF25982 74,0 +39,44 -1,24 38,2 4-085 FF25983 73,4 +39,83 -1,24 38,6 4-090 FF25984 71,8 +39,28 -1,24 38,0 4-090/B FF25987 72,6 +40,12 -1,24 38,9 MEAN 38,5 0,6 4-100 FF25985 69,0 +37,06 -1,24 35,8 4-110 FF25986 69,0 +38,92 -1,24 37,7 4-120 FF26052 74,8 +36,27 -0,8 35,5 4-130 FF25989 70,2 +35,94 -1,24 34,7 4-140 FF25990 72,8 +39,01 -1,24 37,8 4-147 FF25992 72,7 +38,73 -0,77 38,0 5-001 FF25913 65,2 +39,16 -1,13 38,0 5-010 FF25914 75,0 +40,38 -1,13 39,3 5-015 FF25915 71,2 +40,08 -1,13 39,0 5-020 FF25918 74,0 +39,17 -1,13 38,0 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 136 Isotope data 137 5-020 FF25918 74,0 +39,17 -1,13 38,0 5-030 FF25920 67,3 +38,19 -1,23 37,0 5-040 FF25922 76,5 +38,13 -1,23 36,9 5-050 FF25923 73,4 +39,29 -1,23 38,1 5-060 FF25924 67,2 +38,89 -1,23 37,7 5-060/X FF26049 77,0 +38,57 -0,8 37,8 MEAN 37,7 0,1 5-070 FF25925 70,6 +38,79 -1,23 37,6 5-080 FF25927 69,1 +39,18 -1,23 37,9 6-001 FF25928 74,0 +36,46 -1,23 35,2 6-002 FF25929 67,8 +38,81 -1,23 37,6 6-010 FF25932 71,8 +37,44 -0,52 36,9 6-020 FF25933 69,5 +39,13 -0,52 38,6 6-030 FF25935 70,8 +39,09 -0,52 38,6 6-050 FF25936 70,8 +38,51 -0,52 38,0 6-050/B FF25842 69,8 +38,91 -0,52 38,4 MEAN 38,2 0,3 6-060 FF25938 74,8 +38,53 -0,52 38,0 6-062 FF25939 71,8 +38,50 -0,52 38,0 6-070 FF25940 74,9 +37,44 -0,52 36,9 6-076 FF25941 71,9 +37,75 -0,52 37,2 TEPHRA 2-007-T FF25998 80,1 +8,31 -0,77 7,5 2-007-T/B FF26002 82,5 +8,14 -0,77 7,4 MEAN 7,5 0,1 2-041-T FF25999 73,2 +8,33 -0,77 7,6 2-121-T FF26000 79,7 +7,93 -0,77 7,2 2-080-T FF26004 81,2 +7,84 -0,6 7,2 3-040-T FF26005 80,5 +8,15 -0,6 7,6 BFC FF24945 +30,35 BFC FF24959 +30,49 BFC FF24962 +29,78 BFC FF24965 +30,17 BFC FF24977 +30,22 BFC FF24971 +30,65 MEAN 30,28 1s 0,3 FF25065/BFC 76,1 +30,42 FF25069/BFC 76,6 +29,95 -1,31 RUN CORRECTION IS OFFSET OFMEAN OF THE BFC STDS FROM THE PREFFERED VALUE FF25078/BFC 75,8 +30,58 FF25082/BFC 76,3 +30,61 -1,72 FF25091/BFC 78,8 +30,39 FF25095/BFC 78,9 +30,22 -1,43 FF25104/BFC 78,2 +30,40 -1,52 PREFERED VALUE BFC = 28.88 WRT NBS 28 MEAN +30,37 1s 0,2 BFC FF24971 74,0 +30,65 FF24977 79,2 +30,22 FF24984 73,2 +30,72 FF24990 75,4 +30,26 FF24992 79,0 +29,50 FF24997 80,3 +30,11 FF25005 78,0 +30,21 FF25010 82,9 +28,85 FF25018 78,1 +30,41 FF25023 78,0 +30,35 FF25031 78,9 +30,27 FF25036 81,5 +30,05 FF25044 78,4 +30,32 FF25049 78,8 +30,22 FF25065 76,1 +30,42 FF25069 76,6 +29,95 FF25078 75,8 +30,58 FF25104 78,2 +30,40 FF25831 78,9 +29,81 FF25837 77,9 +30,09 FF25844 81,0 +29,68 FF25850 81,4 +30,12 FF25852 76,3 +29,38 FF25857 76,3 +29,91 FF25865 84,1 +28,75 FF25870 78,8 +29,72 FF25891 74,0 +29,72 FF25896 75,9 +30,07 FF25904 80,7 +29,70 FF25908 78,3 +29,60 FF25917 76,2 +30,41 FF26007 80,1 +29,15 FF26014 79,6 +29,81 FF26047 80,7 +29,64 FF26041 79,6 +29,71 FF26275 79,8 +30,11 FF26283 80,4 +29,83 MEAN +29,98 1s 0,5 BFC FF25908 78,3 +29,60 FF25917 76,2 +30,41 FF25930 76,3 +30,11 FF25934 78,9 +28,86 FF25937 77,4 +29,93 FF25947 78,4 +30,12 FF25950 76,8 +30,47 FF25960 81,3 +29,84 FF25963 79,4 +30,09 FF25875 81,5 +30,13 FF25968 80,4 +29,60 FF25988 78,8 +30,16 FF25981 79,4 +30,08 FF25994 79,6 +29,69 FF26001 78,4 +29,60 FF26007 80,1 +29,15 FF26053 83,0 +30,02 FF26054 77,9 +29,68 FF26055 78,1 +29,90 FF26056 80,2 +29,11 MEAN 29,83 1s 0,4 Carbon Samples Sample depth depth Age d13C %C %N C/N 11-3-060 351,3 -351,3 -7357 -25,0 0,09 0,13 0,67 11-3-070 365,8 -365,8 -7493 -25,7 0,04 0,13 0,34 11-3-080 421,7 -421,7 -8016 -25,0 11-3-090 445,7 -445,7 -8265 -23,2 0,20 0,15 1,38 11-3-095 453,2 -453,2 -8354 -24,6 0,54 0,18 3,06 11-3-100 460,7 -460,7 -8444 -25,3 0,42 0,21 1,98 11-3-110 473,5 -473,5 -8595 -26,5 0,17 0,16 1,10 11-3-115 479,9 -479,9 -8671 -26,2 0,24 0,07 3,23 11-3-120 486,3 -486,3 -8747 -25,5 0,29 0,07 3,85 11-3-130 499,1 -499,1 -8913 -25,8 0,26 0,08 3,49 11-3-135 506,0 -506,0 -9008 -26,4 0,07 0,10 0,74 11-3-140 513,0 -513,0 -9103 -24,3 0,44 0,08 5,23 11-3-145 521,0 -521,0 -9211 -29,8 0,08 0,11 0,72 11-3-149 527,4 -527,4 -9298 -24,3 0,41 0,11 3,68 11-4-001 531,0 -531,0 -9348 -24,5 0,22 0,07 3,33 11-4-005 539,2 -539,2 -9459 -26,0 0,17 0,08 2,29 11-4-010 549,5 -549,5 -9598 -26,7 0,26 0,11 2,47 11-4-015 559,7 -559,7 -9720 -27,5 0,09 0,06 1,71 11-4-020 569,9 -569,9 -9842 -25,7 0,17 0,07 2,38 Ultra-high resolution environmental and climatic reconstruction using oxygen and carbon iostopes of diatom frustles 138 11-4-025 576,0 -576,0 -9915 -24,0 0,19 0,04 5,11 11-4-030 582,1 -582,1 -9987 -23,6 0,27 0,04 6,83 11-4-040 592,3 -592,3 -10108 -23,3 0,23 0,06 3,93 11-4-040 (II) -24,8 0,11 0,04 11-4-045 597,4 -597,4 -10168 -22,6 0,29 0,04 6,77 11-4-050 602,5 -602,5 -10228 -22,9 0,42 0,07 5,97 11-4-060 -24,9 11-4-060 (II) 612,6 -612,6 -10349 -25,7 0,24 0,05 4,49 11-4-070 622,8 -622,8 -10432 -26,8 0,20 0,05 3,74 11-4-080 633,0 -633,0 -10502 -27,1 0,20 0,05 3,74 11-4-085 638,0 -638,0 -10538 -26,8 0,26 0,05 5,59 11-4-090 643,1 -643,1 -10573 -25,9 0,16 0,05 3,35 11-4-100 653,3 -653,3 -10643 -24,5 0,24 0,05 5,02 11-4-110 663,5 -663,5 -10714 -27,7 0,21 0,04 5,12 11-4-120 673,6 -673,6 -10784 -26,8 0,37 0,06 6,53 11-4-130 683,8 -683,8 -10855 -27,3 0,37 0,06 6,22 11-4-140 693,9 -693,9 -10925 -27,3 0,18 0,05 3,83 11-4-147 701,1 -701,1 -10966 -28,0 0,10 0,03 3,85 11-5-001 705,1 -705,1 -10988 -25,4 0,21 0,05 4,10 11-5-010 714,3 -714,3 -11037 -27,0 0,08 0,09 0,95 11-5-015 719,4 -719,4 -11064 -26,0 0,25 0,05 4,57 11-5-020 724,4 -724,4 -11091 -26,5 0,24 0,11 2,15 11-5-030 734,6 -734,6 -11145 -28,6 0,21 0,03 6,84 11-5-030 -26,7 11-5-040 744,8 -744,8 -11205 -27,9 0,06 0,10 0,54 11-5-050 754,9 -754,9 -11279 -26,8 0,04 0,12 0,35 11-5-060 765,1 -765,1 -11354 -27,7 0,02 0,11 0,16 11-5-070 774,2 -774,2 -11422 -27,3 0,08 0,05 1,55 11-5-080 775,3 -775,3 -11429 -26,7 0,08 0,06 1,44 11-6-001 790,5 -790,5 -11558 -28,9 0,19 0,07 2,76 11-6-002 791,7 -791,7 -11568 -26,6 0,09 0,06 1,53 11-6-010 799,7 -799,7 -11654 -26,9 0,07 0,06 1,14 11-6-020 809,8 -809,8 -11760 -27,1 0,08 0,05 1,46 11-6-030 820,0 -820,0 -11867 -27,2 0,07 0,06 1,15 11-6-040 830,2 -830,2 -11986 11-6-050 840,3 -840,3 -12105 -27,5 0,02 0,06 0,33 11-6-060 850,5 -850,5 -12232 -27,4 0,30 0,06 5,32 11-6-060 (II) -29,2 11-6-062 852,5 -852,5 -12257 -28,6 0,01 0,06 0,11 11-6-070 860,7 -860,7 -12359 -30,3 0,16 0,03 5,75 11-6-076 863,7 -863,7 -12396 -29,6 0,04 0,06 0,65 %C %N C/N BROC1 41,4 4,5 9,2 BROC1 38,9 4,2 9,3 BROC1 39,6 4,4 9,0 BROC1 40,5 4,0 10,1 BROC1 40,4 4,2 9,7 BROC1 39,1 4,1 9,5 BROC1 40,4 3,9 10,3 BROC1 41,7 4,3 9,6 BROC1 41,7 4,8 8,7 BROC1 41,0 5,1 8,1 Isotope data 139 Appendix C Original papers The palaeohydrological evolution of Lago Chungara´ (Andean Altiplano, northern Chile) during the Lateglacial and early Holocene using oxygen isotopes in diatom silica ARMAND HERNA´NDEZ,1* ROBERTO BAO,2 SANTIAGO GIRALT,1 MELANIE J. LENG,3,4 PHILIP A. BARKER,5 ALBERTO SA´EZ,6 JUAN J. PUEYO,6 ANA MORENO,7,8 BLAS L. VALERO-GARCE´S7 and HILARY J. SLOANE3 1 Institute of Earth Sciences ’Jaume Almera’ (CSIC), Barcelona, Spain 2 Faculty of Sciences, University of A Corun˜a, A Corun˜a, Spain 3 NERC Isotope Geosciences Laboratory, British Geological Survey, Nottingham, UK 4 School of Geography, University of Nottingham, Nottingham, UK 5 Department of Geography, Lancaster Environment Centre, Lancaster University, Lancaster, UK 6 Faculty of Geology, University of Barcelona, Barcelona, Spain 7 Pyrenean Institute of Ecology (CSIC), Zaragoza, Spain 8 Limnological Research Center, University of Minnesota, Minneapolis, Minnesota, USA Herna´ndez, A., Bao, R., Giralt, S., Leng, M. J., Barker, P. A., Sa´ez, A., Pueyo, J. J., Moreno, A., Valero-Garce´s, B. L. and Sloane, H. J. 2008. The palaeohydrological evolution of Lago Chungara´ (Andean Altiplano, northern Chile) during the Lateglacial and early Holocene using oxygen isotopes in diatom silica. J. Quaternary Sci., Vol. 23 pp. 351–363. ISSN 0267-8179. Received 30 July 2007; Revised 21 December 2007; Accepted 22 December 2007 ABSTRACT: Oxygen isotopes of diatom silica and petrographical characterisation of diatomaceous laminated sediments of Lago Chungara´ (northern Chilean Altiplano) have allowed us to establish its palaeohydrological evolution during the Lateglacial–early Holocene (ca. 12 000–9400 cal. yr BP). These laminated sediments are composed of light and dark pluriannual couplets of diatomaceous ooze formed by different processes. Light sediment laminae accumulated during short-term diatom blooms whereas dark sediment laminae represent the baseline limnological conditions during several years of deposition. Oxygen isotope analysis of the dark diatom laminae show a general d18O enrichment trend during the studied period. Comparison of these d18Odiatom values with the previously published lake-level evolution suggests a correlation between d18Odiatom and the precipitation:evaporation ratio, but also with the evolution of other local hydrological factors as changes in the groundwater outflow as well as shifts in the surface:volume ratio of Lago Chungara´. The lake expanded (probably increasing this ratio) during the rising lake-level trend due to changes in its morphology, enhancing evaporation. Furthermore, the lake’s hydrology was probably modified as the groundwater outflow became sealed by sediments, increasing lake water residence time and potential evaporation. Both factors could cause isotope enrichment. # Natural Environment Research Council (NERC) copyright 2008. Reproduced with the permission of NERC. Published by John Wiley & Sons, Ltd. KEYWORDS: diatom ooze; laminated sediments; oxygen isotopes; rhythmites; Holocene; Andean Altiplano. Introduction Oxygen isotopes of diatom silica have been widely used in palaeoenvironmental reconstructions from lake sediments in the last decade (see Leng and Barker, 2006, for a comprehen- sive review). Using oxygen isotope ratios in palaeoenviron- mental reconstruction is, however, not easy, because the sedimentary record can be influenced by a wide range of interlinked environmental processes ranging from regional climate change to local hydrology. The oxygen isotopic composition of diatom silica depends on the isotope composition of the water when the skeleton of the siliceous micro-organisms is secreted, and also on the ambient water temperature (Shemesh et al., 1992). Therefore, knowledge of all the environmental factors that may have influenced the isotope composition of the lakewater is vital for the interpretation of the JOURNAL OF QUATERNARY SCIENCE (2008) 23(4) 351–363  Natural Environment Research Council (NERC) copyright 2008. Reproduced with the permission of NERC. Published by John Wiley & Sons, Ltd. Published online in Wiley InterScience (www.interscience.wiley.com) DOI: 10.1002/jqs.1173 *Correspondence to: A. Herna´ndez, Institute of Earth Sciences ’Jaume Almera’ (CSIC), C/Lluı´s Sole´ i Sabarı´s s/n, E-08028 Barcelona, Spain. E-mail: ahernandez@ija.csic.es Contract/grant sponsor: Spanish Ministry of Science and Education; contract/ grant numbers: BTE2001-3225, BTE2001-5257-E, CGL2004-00683/B, TECGL2007-60932/BTE. d18Odiatom signal (Leng et al., 2005a). One of these environ- mental factors is evaporation, which has a major influence on the isotope composition of any standing water body (Leng and Marshall, 2004). The d18O record can therefore be used, at least in closed lakes, as an indicator of changes in the precipitation to evaporation ratio (P/E) related to climatic changes (Leng and Marshall, 2004). Yet, before any palaeoclimatic interpretation of the isotope records from a lake is considered, other local palaeohydrological intervening factors from the basin need to be taken into account (Sa´ez and Cabrera, 2002; Leng et al., 2005a). The sedimentary records of high-altitude Andean Altiplano lakes are good candidates for carrying out oxygen isotope studies to reconstruct the late Quaternary palaeoclimatology of the region. They preserve an excellent centennial- to millennial-scale record of effective moisture fluctuations and source changes during the Lateglacial and Holocene, although the interpretation is not always straightforward (Abbot et al., 1997; Argollo and Mourguiart, 2000; Valero-Garce´s et al., 2000, 2003; Grosjean et al., 2001; Baker et al., 2001a, 2001b; Tapia et al., 2003; Fritz et al., 2004, 2006; Placzek et al., 2006). The d18O analyses of carbonates, cellulose and biogenic silica have successfully been used to reconstruct the hydrological responses to climate change in different Andean lacustrine systems (Schwalb et al., 1999; Seltzer et al., 2000; Abbott et al., 2000, 2003; Wolfe et al., 2001; Polissar et al., 2006). Up to now, only stable isotopes in carbonates have been examined in Lago Chungara´ (Valero-Garce´s et al., 2003), although its sedimentary record is made up of rich diatomac- eous ooze ideal for diatom silica oxygen isotope studies. Lago Chungara´ currently behaves as a closed lake, without any surface outlet, and evaporation is the dominant water loss process (Herrera et al., 2006); however, it has shown a complex depositional history since the Lateglacial (Sa´ez et al., 2007) and the relative role of other factors (groundwater versus evapor- ation) should be evaluated. Here we examine a high-resolution d18O diatom silica record of three selected sections belonging from the Lateglacial to early Holocene (ca. 12,000–9400 cal. yr BP) from Lago Chungara´. We emphasise the role that some local factors such as sedimentary infill and palaeohydrology can play on the interpretation of the d18O diatom silica record and therefore the need to discriminate between the climatic and local environ- mental signals. The Lago Chungara´ Geology, climate and limnology Lago Chungara´ (188150 S, 698100 W, 4520 m a.s.l.) is located at the NE edge of Lauca Basin, in the Chilean Altiplano. It lies in a highly active tectonic and volcanic context (Clavero et al., 2002). The lake sits in the small hydrologically closed Chungara´ Sub-Basin, which was formed as a result of a debris avalanche during the partial collapse of the Parinacota Volcano, damming the former Lauca River (Fig. 1(A)). Lago Chungara´ and Lagunas Cotacotani were formed almost immediately. The collapse post-avalanche event has been dated and the ages range between 18 000 cal. yr BP, using He-exposure techniques (Wo¨rner et al., 2000; Hora et al., 2007), and 11 155–13 500 14C yr BP, employing radiocarbon dating methods (Francis and Wells, 1988; Baied and Wheeler, 1993; Amman et al., 2001). In these cases the authors dated lacustrine sediments from Lagunas Cotacotani. In addition, Clavero et al. (2002, 2004) dated palaeosol horizons by radiocarbon and proposed a maximum age of 8000 14C yr BP for the collapse. Lago Chungara´ is situated in the arid Central Andes, in a region dominated by tropical summer moisture (Garreaud et al., 2003). The isotope composition of rainfall (Aravena et al., 1999; Herrera et al., 2006) and the synoptic atmospheric precipitation patterns (Ruttlant and Fuenzalida, 1991) indicate that the main moisture source comes from the Atlantic Ocean via the Amazon Basin. During the summer months (DJFM) weak easterly flow prevails over the Altiplano as a consequence of the southward migration of the subtropical jet stream and the establishment of the Bolivian high-pressure system (Garreaud et al., 2003). This narrow time window defines the wet season in the Altiplano (Valero-Garce´s et al., 2003). Mean annual rainfall in the Chungara´ region is about 350mm  yr1, but the actual range is variable (100–750mm  yr1). Mean tempera- ture is 4.28C and the potential evaporation was estimated at over 4750mm  yr1 (see references in Valero-Garce´s et al., 2000). In this region, a significant fraction of the inter-annual variability of summer precipitation is currently related to the El Nin˜o Southern Oscillation (ENSO) (Vuille, 1999). El Nin˜o years seem to be recorded in the Sajama and Quelcaya ice cores by significant decreases in snow accumulation (Thompson et al., 1986; Vuille, 1999). Instrumental data from the Chungara´ region show a reduction of the precipitation during moderate to intense El Nin˜o years. However, there is no direct relationship between the relative El Nin˜o strength and the amount of rainfall reduction (for further details see Valero-Garce´s et al., 2003). Rainfall isotope composition in this region is characterised by a large variability in d18O (between þ1.2 and 21.1% SMOW) and of dD (between þ22.5 and 160.1% SMOW). The origin of the lightest isotope values are the strong kinetic fractionation in the air masses from the Amazon. The altitudinal isotopic gradient of d18O in the Chungara´ region is very high (between þ0.76%/100 m and þ2.4%/100m) compared with other worldwide regions (Herrera et al., 2006). Lago Chungara´ has an irregular shape with a maximum length of 8.75 km, maximum water depth of 40m, a surface area of 21.5 km2 and a volume of 400 106m3 (Mu¨hlhauser et al., 1995; Herrera et al., 2006) (Fig. 1(B)). The western and northern lake margins are steep, formed by the eastern slopes of Ajoya and Parinacota volcanoes. The eastern and southern margins are gentle, formed by the distal fringe of recent alluvial fans and the River Chungara´ valley (Sa´ez et al., 2007). At present, the main inlet to the lake is the Chungara´ River (300–460 L  s1), although secondary streams enter the lake in the southwestern margin. Themain water loss is by evaporation (3.107m3  yr1) but it has been estimated that groundwater outflow from Lago Chungara´ to Lagunas Cotacotani is about 6–7.106m3  yr1 (Dorador et al., 2003). The calculated residence time for the lake’s water is approximately 15 yr (Herrera et al., 2006). The lake is polymictic, oligomesotrophic to meso-eutrophic (Mu¨hlhauser et al., 1995), contains 1.2 g  L1 of total dissolved solids, its conductivity ranges between 1500 and 3000mS  cm1 (Dorador et al., 2003) and the water chemistry is of Na–Mg–HCO3–SO4 type. Tempera- ture profiles measured in November 2002 showed a gradient from the lake surface (9.1–12.18C) to the lake bottom (6.2–6.48C at 35m of water depth), with a thermocline (0.5–0.68C) located at about 19 m of water depth. Oxygen ranged from 11.9–12.5 ppm (surface) to 7.6 ppm (bottom) and the pH oscillated between 8.99 (surface) and 9.30 (bottom). Lake water is enriched by evaporation with regard to rainfall and spring waters. The mean values of d18O and dD are1.4% SMOW and 43.4% SMOW, respectively (Herrera et al.,  Natural Environment Research Council (NERC) copyright 2008. Reproduced with the permission of NERC. Published by John Wiley & Sons, Ltd. J. Quaternary Sci., Vol. 23(4) 351–363 (2008) DOI: 10.1002/jqs 352 JOURNAL OF QUATERNARY SCIENCE 2006). Primary productivity is mainly governed by diatoms and chlorophyceans (Dorador et al., 2003). Macrophyte commu- nities in the littoral zone form dense patches that contribute to primary productivity. Seasonal measurements of conductivity, nitrate, phosphate and chlorophyll reveal these changes in productivity and in the composition of algal communities are mainly due to changes in water temperature and salinity (Dorador et al., 2003). The absence of raised lacustrine deposits around the lake margins suggests that the current level of the lake is at its highest since lake formation (Sa´ez et al., 2007). Previous work and sedimentary sequence In November 2002 15 sediment cores (6.6 cm inner diameter and up to 8m long) were recovered from Lago Chungara´ using a raft equipped with a Kullenberg system. All cores were cut in 1.5m sections and physical properties (GRAPE density, p-wave velocity and magnetic susceptibility) were measured in the laboratory using a GEOTEKTM multi-sensor core logger (MSCL) at 1 cm intervals. Afterwards, the cores were split into two halves, scanned using a DMT colour scanner, and the textures, colours and sedimentary structures were described. Smear slides were prepared for the description of the sediment composition and to estimate the biogenic, clastic and endogenic mineral content. After a detailed lithological correlation of the cores (Sa´ez et al., 2007), cores 10 and 11 located offshore were selected for conducting the palaeoenvironmental reconstruction. A com- posite core recording the whole sedimentary infill (minimum thickness of 10m) of the offshore zone was constructed from the detailed description and correlation of cores 10 and 11. From here on all core depths are referred to this composite core. Figure 1 (A) Location of Lago Chungara´ on NE edge of Lauca Basin. (B) Bathymetric map of Lago Chungara´ showing the main morphological units of the lake floor and position of the sediment cores  Natural Environment Research Council (NERC) copyright 2008. Reproduced with the permission of NERC. Published by John Wiley & Sons, Ltd. J. Quaternary Sci., Vol. 23(4) 351–363 (2008) DOI: 10.1002/jqs THE PALAEOHYDROLOGICAL EVOLUTION OF LAGO CHUNGARA´ 353 From the bottom to the top of the core, two sedimentary units (units 1 and 2) were identified and correlated mainly using tephra keybeds. These lithological units were subdivided into two subunits (subunits 1a, 1b, 2a and 2b). Basal unit 1a ranges between 0.58m and 2.56m of thickness and is made up of finely laminated green andwhitish diatomaceous ooze. Unit 1b (from 0.62m to 1.87m thick) is composed of laminated and massive brown diatomaceous ooze with carbonate-rich intervals. Unit 2a (between 1.56m and 3.44m thick) is made up of a brown diatomaceous ooze with tephra layers and carbonate-rich intervals. The sediments of the uppermost unit 2b range from 0.86m to 3m in thickness and they are composed of dark grey to black diatomaceous ooze with abundant tephra layers (for further details see Moreno et al., 2007, and Sa´ez et al., 2007). The cores have been analysed for a number of proxies including X-ray fluorescence (XRF), X-ray diffraction (XRD), total organic and inorganic carbon (TOC and TIC), pollen, diatoms and total biogenic silica (Moreno et al., 2007; Sa´ez et al., 2007). The chronological model for the sedimentary sequence of Lago Chungara´ is based on 17 AMS 14C dates of bulk organic matter and aquatic plant macrofossils, and one 238U/230Th date from carbonates. The radiocarbon dates were performed in the Poznan Radiocarbon Laboratory (Poland), whereas the 238U/230Th sample was analysed by high-resolution inductively coupled plasma–infrared mass spectrometry (ICP-IRMS) at the University of Minnesota (Edwards et al., 1987; Cheng et al., 2000; Shen et al., 2002). The present-day reservoir effect was determined by dating the dissolved inorganic carbon (DIC) of the lake water at the Beta Analytics Inc. laboratory (USA). The real reservoir effect of the lake was calculated by correcting the DIC radiocarbon date for the effects of thermonuclear bomb tests (Hua and Barbetti, 2004). The calibration of radiocarbon dates was performed using CALIB 5.02 software and the INTCAL98 curve (Stuiver et al., 1998; Reimer et al., 2004). The software described in Heegaard et al. (2005) was used to construct the final age–depth model (see Moreno et al., 2007, and Giralt et al., 2007, for details). Materials and methods Three intervals from unit 1 were selected and sampled for thin-section study and d18O diatom silica analysis. Interval 1 (located at the subunit 1a, between 831 cm and 788 cm core depth) is made up of finely laminated green and whitish sediments. Interval 2 (between 605 cm and 622 cm core depth) is found in the transition between subunits 1a and 1b and is made up of laminated green and pale-brown diatomaceous ooze. Interval 3 (located at subunit 1b, between 537 cm and 574 cm core depth) is made up of laminated dark-brown and white diatomaceous ooze with carbonates. The selection criteria of these three intervals are discussed below. The chronological model defines the corresponding age of the three intervals. Interval 1 was deposited between 11 990 and 11 530 cal. yr BP, interval 2 between 10 430 and 10 260 cal. yr BP and interval 3 between 9890 and 9430 cal. yr BP. Each interval was continuously covered by thin sections. Thin sections of 120mm 35mm (30mm in thickness), with an overlap of 1 cm at each end, were obtained after freeze-drying and balsam-hardening. Detailed petrographical descriptions and lamina thickness measurements were performed using a Zeiss Axioplan 2 Imaging petrographic microscope. Several samples were also selected for observation with a Jeol JSM-840 electronmicroscope in order to complement the petrographical study. Each lamina of the three intervals was sampled with a blade for isotope analysis. A total of 190 samples (111 samples from interval 1, 37 samples from interval 2 and 42 samples from interval 3) were obtained. However, a selection of 37 samples from dark laminae were selected for d18Odiatom analyses to investigate the baseline hydrological evolution of Lago Chungara´. These dark laminae would represent a normal annual cycle of the lake with alternating phases of stratification andmixing. These conditions would lead to the development of a complex diatom community, among other algal groups (Herna´ndez et al., 2007). Analysis of the oxygen isotope composition of diatom silica from these 37 samples requires that thematerial is almost pure diatomite (Juillet-Leclerc, 1986), so a meticulous protocol involving chemical attack, sieving, settling and laminar flow separation was performed. Specifi- cally, our samples were treated following the method proposed by Morley et al. (2004), with some variations (Fig. 2(A)). The first stage (chemical attack) followed the standard method in order to remove the carbonates (10% HCl) and organic matter (hydrogen peroxide) (Battarbee et al., 2001), but also included a further step using concentrated HNO3 in order to remove any remaining organic matter. The second stage (sieving at 125mm) allowed us to eliminate resistant charcoal and terrigenous particles. The 63mm and 38mm sieves allowed us to obtain a fraction of quasi-monospecific diatoms (Cyclostephanos andinus) in most of the samples. The third stage was an alternative approach to heavy liquid separation. Gravitational split-flow thin fractionation (SPLITT) was employed at Lancaster University (UK) (Rings et al., 2004; Leng and Barker, 2006). The SPLITT technique was only applied to the most problematic samples which contained remaining difficult-to-separate clay and fine tephra particles. In the final step, the purified diatom samples were dried at 408C for 24–48h. After the cleaning process six samples were checked with XRD, total carbon (TC) analysis and scanning electron microscopy (SEM) observations. This checking process revealed that the samples did not contain significant terrigenous matter. The TC values were below 0.5%wt and the terrigenous content (clays or tephra) was less than 1%wt (Fig. 2(B)). Although a large number of diatoms were broken during the cleaning process, this did not affect the final isotope data. We therefore assume that the d18O values of the purified samples retained climatic and hydrological information (Morley et al., 2004; Leng and Barker, 2006). Oxygen extraction for isotope analyses followed the classical step-wise fluorination method (Matheney and Knauth, 1989). The method involved three steps. First, the hydrous layer was stripped by outgassing in nickel reaction tubes at room temperature. Second, a pre-fluorination clean-up step involved a stoichiometric deficiency of reagent, bromine pentafluoride (BrF5), heated at 258C for several minutes. The final step was a full reaction at 4508C for 12 h with an excess of BrF5. The oxygen liberated was converted to CO2 by exposure to hot graphite (following Clayton and Mayeda, 1963). The oxygen yield was monitored, for every sample, by comparison with the calculated theoretical yield for SiO2. The intervals examined here hadmean yields of 69–70%of their theoretical yield based on silica. This fact suggests that around 30% of the material, including hydroxyl and loosely bonded water (both OH and H2O), was removed during prefluorination. A random selection of five samples was analysed in duplicate, giving a reproducibility between 0.01% and 0.6% (1s). The standard laboratory quartz and a diatomite control sample (BFC) had a mean reproducibility over the period of analysis of 0.2%. The CO2 was analysed for 18O/16O using a FinniganTM Matt 253 mass spectrometer. The results were calibrated versus  Natural Environment Research Council (NERC) copyright 2008. Reproduced with the permission of NERC. Published by John Wiley & Sons, Ltd. J. Quaternary Sci., Vol. 23(4) 351–363 (2008) DOI: 10.1002/jqs 354 JOURNAL OF QUATERNARY SCIENCE Figure 2 (A) Diagram showing the three-stage cleaning method for concentrating diatoms for oxygen isotope analysis (modified from Morley et al., 2004). (B) SEM images of two samples before cleaning (left) and after cleaning (right)  Natural Environment Research Council (NERC) copyright 2008. Reproduced with the permission of NERC. Published by John Wiley & Sons, Ltd. J. Quaternary Sci., Vol. 23(4) 351–363 (2008) DOI: 10.1002/jqs THE PALAEOHYDROLOGICAL EVOLUTION OF LAGO CHUNGARA´ 355 NBS-28 quartz international standard. Data are reported in the usual delta form (d) as per mil (%) deviations from V-SMOW. The fluorination process and the 18O/16O ratios measured were carried out at the NERC Isotope Geosciences Laboratory, British Geological Survey (UK). Results: petrography and isotope composition of diatoms Smear slide, SEM and several analyses (XRD, TC, biogenic silica) of the lake sediments before they were prepared for isotope analysis showed that the samples were composed of both amorphous and crystalline material. The amorphous fraction comprises biogenic silica (between 47% and 58% weight), organic matter and volcanic glass. The crystalline fraction represented <10% of the sediments. Interval 1 (11 990–11 530 cal. yr BP) Diatom concentration ranges from 108.3 to 633.8 million valves g1. The interval is dominated by euplanktonic diatoms ranging from 79.1% to 93.9% of the diatom assemblage. The thicknesses of the laminae are between 0.9 and 10.3mm (Fig. 3(A)). Smear slide, thin section and SEM observations showed that light laminae were quasi-monospecific layers of large Cyclostephanos andinus (diameter> 50mm) (Fig. 4(D)). The upper contact of the light laminae with the dark laminae is transitional, showing an increase in diatom diversity with subdominant tychoplanktonic (Fragilaria spp.) and benthic diatoms (mainly Cocconeis spp., Achnanthes spp., Navicula spp. and Nitzschia spp.) (Fig. 4(C)), whereas the lower contact is abrupt (Fig. 4(A)). Diatom valves show good preservation with no preferred orientation in the lower part, but increasingly orientated upwards. The content of the organic matter also increases upwards. Dark laminae comprise a more diverse mixture of diatoms, including the euplanktonic (those having a strict planktonic character) smaller Cyclostephanos andinus (diameter< 50mm) than those found in light laminae, and diatoms of the Cyclotella stelligera complex, as well as tychoplanktonic (those usually having a benthic life form but which can occasionally be facultatively planktonic) and benthic diatoms (bottom-dwelling forms) (Fig. 4(B)). These dark laminae are also enriched in organic matter, probably from diatoms and other algal groups. Up to 41 light and dark laminae couplets were defined. The thickness of these couplets ranges between 4.2mm and 22.5mm and, according to the chronological model, they are pluriannual (mean about 10 yr). The rhythmite starts with the dominance of light laminae, progressively changing to a dominance of dark laminae. The d18Odiatom values of the purified diatoms in interval 1 range from þ35.5% to þ39.2% (Fig. 3(A)). Higher d18Odiatom occur in the lower part of the interval (around 822 cm of core depth). There is an upward decreasing trend (1.9% 100 yr1) attaining a minimum of þ35.5% around 803 cm depth. This stretch is followed by an increasing shift of 2.9% 100 yr1 towards the upper part of the interval, where a relative maximum of þ38.8% is reached at 793 cm depth. The uppermost two samples show a light depletion. The mean d18Odiatom value of this interval is þ37.8 0.85%. Interval 2 (10 430–10 260 cal. yr BP) Diatom concentration ranged from 95.2 to 218 million valves g1 in interval 2. Almost 94% of the diatom assemblages of this interval were made up of euplanktonic diatoms. Benthic taxa show the minimum values for the three analysed intervals. The thickness of diatomaceous ooze laminae ranged from 1.8mm to 16mm (Fig. 3(B)). Light laminae were dominated by large Cyclostephanos andinus (diameter> 50mm), with some tychoplanktonic (Fragilaria spp.) and benthic diatoms, as well as minor amounts of siliciclasts and organic matter. Dark laminae are composed of a mixture of small and large Cyclostephanos andinus valves, with more abundant tycho- planktonic and benthic diatoms (as well as organic matter) compared to light laminae. Diatom valves are not so well preserved as in interval 1, sometimes showing a high degree of fragmentation and a preferred orientation. The contact between the laminae is similar to those found in interval 1. Clear couplets were only observed in the upper two-thirds of the interval and only 10 couplets could be identified (Fig. 3(B)). They are pluriannual (mean couplet represents about 10 yr of sedimentation) and their thicknesses range between 5.5 and 19mm. Light laminae were more abundant in the upper part of interval 2, whereas dark laminae are more abundant in the lower part. The d18Odiatom curve shows a clear increasing trend during this interval (Fig. 3(B)). The lowest d18Odiatom value (þ36%) was recorded at the bottom of the interval (617 cm depth) and the maximum at the two uppermost samples (þ39.7% and þ39.6%; 606–605 cm of core depth). The magnitude of the increasing trend is much higher between the two lowermost samples (18.5% 100 yr1) than for the rest of the interval (0.6% 100 yr1). The mean d18Odiatom value of this interval is þ38.7 1.4%. Interval 3 (9890–9430 cal. yr BP) Diatom concentration ranges between 163.8 and 255.8 million valves g1 for interval 3. Euplanktonic diatoms (68.6–98.1%) also dominate this interval, and have the minimum values for the three intervals. On the contrary, benthic diatoms show moderate values (up to 31.4%), being the highest for the three intervals. Light diatomaceous ooze laminae ranged between 0.9 and 12.3mm in thickness (Fig. 3(C)) and they comprise Cycloste- phanos andinus (diameter> 50mm), increasing upwards in both taxonomic diversity and organic matter content. The lower contact with dark laminae shows an abrupt change in diatom size, whereas the upper one is gradual. Diatom valves show good preservation with no orientation in the lower part but are preferentially oriented upwards. Dark laminae comprise a mixture of smaller Cyclostephanos andinus (diame- ter< 50mm), with subdominant tychoplanktonic and benthic diatoms, as well as a high organic matter content. The lower contact is gradual whereas the upper one abrupt. Up to 18 light and dark pluriannual couplets were defined (mean couplet represent around 12 yr). These couplets are 3–18mm thick. The rhythmite starts with light laminae, progressively changing to dark laminae. The d18Odiatom curve for interval 3 (Fig. 3(C)) shows an overall continuous increasing trend of 0.9% 100 yr1 from þ39.1% (570 cm of core depth) to þ41.3% (548 cm of core depth). Superimposed over the general trend are short-term fluctuations. The mean d18Odiatom value of this interval is þ40.1 0.77%.  Natural Environment Research Council (NERC) copyright 2008. Reproduced with the permission of NERC. Published by John Wiley & Sons, Ltd. J. Quaternary Sci., Vol. 23(4) 351–363 (2008) DOI: 10.1002/jqs 356 JOURNAL OF QUATERNARY SCIENCE Fi gu re 3 D ig it al im ag es o ft h e th re e in te rv al s se le ct ed ac co rd in g to it s d ep th an d ti m es ca le .T h e id en ti fi ed co u p le ts an d th e d1 8 O va lu es fr o m d ia to m si li ca h av e b ee n p lo tt ed fo r in te rv al 1 (A ), in te rv al 2 (B )a n d in te rv al 3 (C ). St ip p le d li n e sh o w s m ea n va lu es  Natural Environment Research Council (NERC) copyright 2008. Reproduced with the permission of NERC. Published by John Wiley & Sons, Ltd. J. Quaternary Sci., Vol. 23(4) 351–363 (2008) DOI: 10.1002/jqs THE PALAEOHYDROLOGICAL EVOLUTION OF LAGO CHUNGARA´ 357 Figure 4 Rhythmite type showing thickness, colour, ecological succession and temporal scale. (A) SEM image showing the contact between dark (bottom) and light lamina (top). (B) Petrographical microscope image of the dark lamina. (C) SEM image showing the transitional contact between light (bottom) and dark lamina (top). (D) Petrographical microscope image of the light lamina. See text for details. This figure is available in colour online at www.interscience.wiley.com/journal/jqs  Natural Environment Research Council (NERC) copyright 2008. Reproduced with the permission of NERC. Published by John Wiley & Sons, Ltd. J. Quaternary Sci., Vol. 23(4) 351–363 (2008) DOI: 10.1002/jqs 358 JOURNAL OF QUATERNARY SCIENCE The three intervals have different d18Odiatom averages displaying a progressive low-frequency enrichment from interval 1 (þ37.8 0.85%) to interval 3 (þ40.1 0.77%). The overall isotopic enrichment is 2.1% throughout these intervals. Discussion The sedimentary model of diatom rhythmites Laminated diatomaceous oozes in the sedimentary record of Lago Chungara´ comprise variable-thickness couplets of alter- nating light and dark laminae. These couplets display different features (colour and mean thickness) in the three intervals described here, although they exhibit similar diatom assem- blages and textural characteristics, and therefore it is assumed that their formation is by similar environmental processes. Rhythmite types have been established (Fig. 4); light laminae are formed almost exclusively by diatom skeletons of a quasi-monospecific assemblage of Cyclostephanos andinus, while dark laminae, with a high organic matter content, comprise a mixture of a more diverse diatom assemblage, including the euplanktonic Cyclostephanos andinus, although diatoms of the Cyclotella stelligera complex are co-dominant taxa. Subdominant groups are some tychoplanktonic (Fragilaria spp.) and benthic taxa (Cocconeis spp., Achnanthes spp., Navicula spp., Nitzschia spp.). Each couplet was deposited during time intervals ranging from 4 to 24 yr according to our chronological model. Couplets are therefore not a product of annual variations in sediment supply but due to some kind of pluriannual processes. The good preservation and size of diatom valves in the light laminae suggest accumulation during short-term extraordinary diatom blooms, perhaps of only days to weeks in duration. These diatom blooms could have been triggered by climatically driven strong nutrient inputs to the lake and/or to nutrient recycling under extreme turbulent conditions and mixing affecting the whole water column. On the contrary, the baseline conditions are represented by the dark laminae. Each of these laminae is made up of the remains (organic matter and diatom skeletons) of a diverse planktonic community deposited throughout several years under different water column mixing regimes. The preserved remains are therefore a reflection of different stages in the phytoplankton succession throughout several years (Reynolds, 2006). Lake level and d18Odiatom changes A preliminary lake-level reconstruction of Lago Chungara´ was undertaken employing the variations of euplanktonic diatoms, Botryococcus and macrophyte remains (see Sa´ez et al., 2007). This reconstruction shows a general deepening trend during the Lateglacial and early Holocene. This overall increase in lake level is punctuated by one deepening (D1; Fig. 5) and by two shallowing episodes (S1 and S2; Fig. 5). According to Sa´ez et al. (2007) the three selected intervals described here represent two different lacustrine conditions. Intervals 1 and 3 are likely shallower episodes that occurred in different climatic periods, whereas interval 2 occurred during a period between two shallow intervals, and likely with higher lake-level conditions. However, the resolution of the lake-level reconstruction provided by Sa´ez et al. (2007) does not preclude the occurrence of shallowing episodes other than those previously detected. The isotope analyses presented here of these three intervals have allowed us to characterise the hydrological evolution of the lake for these different lacustrine conditions during the Lateglacial and early Holocene. Dark laminaewere selected for d18Odiatoms analyses to investigate the baseline hydrological evolution of Lago Chungara´. These dark laminae would represent a normal annual cycle of the lake with alternating phases of stratification and mixing. These conditions would lead to the development of a complex diatom community among other algal groups (Herna´ndez et al., 2007). The d18Odiatom variation can result from a variety of processes (Jones et al., 2004; Leng et al., 2005b) but for closed lakes, particularly in arid regions where water loss is mainly through evaporation, measured d18Owater values are alwaysmore enriched than those of ambient precipitation since the oxygen lighter isotope (16O) is preferentially lost via evaporation. Under these circum- stances, the d18Odiatom record can be used as an indicator of Figure 5 Lake-level evolution curve based on biological indicators (modified from Sa´ez et al., 2007). Deepening–shallowing episode (D1) and shallowing–deepening episodes (S1 and S2) are indicated. The lake followed an overall deepening trend (see Sa´ez et al., 2007, for further details). Shaded bands mark the three studied intervals. On the right corresponding mean values of d18O from diatom silica of the studied intervals are shown  Natural Environment Research Council (NERC) copyright 2008. Reproduced with the permission of NERC. Published by John Wiley & Sons, Ltd. J. Quaternary Sci., Vol. 23(4) 351–363 (2008) DOI: 10.1002/jqs THE PALAEOHYDROLOGICAL EVOLUTION OF LAGO CHUNGARA´ 359 changes in the precipitation:evaporation ratio (P/E) related to climatic changes (Leng and Marshall, 2004). Lago Chungara´ is a hydrologically closed lake and its main water loss is currently via evaporation, meaning that changes in d18O values should be directly related to shifts in P/E. The lake-level change from the deeper-water conditions recorded during the sedimentation of interval 2 to the shallower conditions that occurred during the deposition of interval 3, according to the Sa´ez et al. (2007) reconstruction, is compatible with the observed increase in d18O values. However, the isotope values and the lake-level reconstruction do not agree over the transition from interval 1 to interval 2. The isotope values suggest a reduced P/E (shallower) stage, whereas several proxy indicators suggest deeper conditions (Fig. 5). A possible explanation for this could involve shifts in d18O related to other environmental circumstances, such as variations in the morphometrical parameters and changes in the groundwater outflow. Changes in the surface:volume ratio and in ground- water outflow of Lago Chungara´ from the Lateglacial to early Holocene are the factors likely to account for most of the shifts found in the d18O values. Besides fluctuations in P/E, another factor to take into account is basin morphology. During the lake’s evolution the lake’s surface:volume ratio would have changed. A tentative palaeobathymetric reconstruction of Lago Chungara´ based on the lake-level curve from Sa´ez et al. (2007) (Fig. 6) shows that during the Lateglacial the lake only occupied the present central plain area. The rise in the lake level during the early Holocene, although punctuated by some oscillations, flooded the extensive eastern and southern margins of the basin. Under these circumstances, the lake underwent a significant increase in its surface area (Fig. 6). Because the eastern margin is much shallower than the central plain (Fig. 1), the whole lake’s surface:volume ratio would have significantly increased, and also concurrently the relative importance of evaporation. Thus the observed d18O high values of interval 3 could be explained Figure 6 Hydrological evolution of the Lago Chungara´ in the Lateglacial–early Holocene. North–South cross-section of the lake (left) and water lake surface area (right) for the sedimentation of interval 1 (11 990–11 530 cal. yr BP (A)), interval 2 (10 430–10 260 cal. yr BP (B)) and interval 3 (9890–9431 cal. yr BP (C))  Natural Environment Research Council (NERC) copyright 2008. Reproduced with the permission of NERC. Published by John Wiley & Sons, Ltd. J. Quaternary Sci., Vol. 23(4) 351–363 (2008) DOI: 10.1002/jqs 360 JOURNAL OF QUATERNARY SCIENCE not only by the shallowing trend from interval 2 to interval 3, but also by the increase in the lake’s surface:volume ratio between both intervals. There are no signs of subaerial exposure in the recovered sediments of the eastern platform, which indicates that lake water level did not drop significantly afterwards. Although the lake was deeper during interval 3 than during interval 1, the mean isotope value is higher during interval 3. This fact could be explained by the increase in the surface:volume ratio and by the reduction of groundwater losses. Hence the morphology of the lake, and not only water depth, must be considered as a key factor in any interpretation of the d18Odiatom in terms of changes in P/E. Furthermore, changes in the groundwater fluxes in Lago Chungara´ could have been a significant factor in the shifts found in the d18O values from the Lateglacial to early Holocene. The groundwater outflow from the lake during the Lateglacial was probably higher than during the Holocene. This condition would progressively change with the sedimen- tary infill of the basin. Drainage, through the breccia barrier, would progressively become less efficient as the groundwater outflows silted up (Leng et al., 2005a). Thus, evaporative losses would have predominated over groundwater during the early Holocene. This highlights the fact that stable isotopes would not, in this case, have a direct correspondence with changes in the lake water level. In summary, the relative increase in evaporation due to the increase in the lake’s surface:volume ratio between the studied intervals could have played a significant role. Superimposed onto this situation, the increase in d18O values from the Lateglacial (when the lake was at its shallowest) to the early Holocene (when the overall deepening trend started) is also likely to have been related to a change to a predominantly evaporative lake as the lake’s bottom became more imperme- able due to sediment basin sealing. Conclusions The thin section study of diatomaceous laminated sediments shows that the rhythmites are made up of light quasi- monospecific lamina of the euplanktonic diatom Cyclostepha- nos andinus and a pluriannual dark lamina rich in organic matter and a mixture of a more diverse diatom assemblage. The formation of light laminae is apparently related to short-term (days to weeks) diatom blooms, whereas dark laminae represent baseline conditions lasting several years. The oxygen isotope record of the dark laminae diatoms of Lago Chungara´ indicates a progressive d18O enrichment from the Lateglacial to early Holocene. Besides changes in the P/E ratio, two other factors could have governed shifts in the Lago Chungara´ d18O record. The basin’s stepped morphology forced the expansion of the lake towards the eastern and southern shallow margins during the rising trend. These changes could have caused an increase in the lake’s surface:volume ratio, thus enhancing the evaporation which caused isotope enrichment during the early Holocene. In addition, the hydrology of the lake was probably modified during the Lateglacial to early Holocene transition as the lake’s groundwater outflow became progressively sealed by sediments, thereby increasing lake water residence time and potential evaporation. In summary, changes in the groundwater:evaporation loss ratio and changes in the lake’s extent caused isotope enrichment during the Lateglacial and early Holocene. Previous work has focused on issues of diagenesis, contamination and host–water interactions that can all influence d18Odiatom, whereas local hydrological factors have been largely neglected. These results point to the complex interplay among the different factors which intervene in the diatom oxygen isotope record of closed lakes and how interpretation needs to be adapted to the different evolutionary stages of the lake’s ontogeny. This study highlights the importance of reconstructing local palaeohydrology as this may be only indirectly related to palaeoclimate. Acknowledgements The Spanish Ministry of Science and Education funded the research at Lago Chungara´ through the projects ANDESTER (BTE2001-3225), BTE2001-5257-E, LAVOLTER (CGL2004-00683/BTE) and GEOBILA (CGL2007-60932/BTE). The Limnological Research Center (University of Minnesota, USA) provided the technology and expertise to retrieve the cores. NERC (UK) funded the isotope analysis. We are grateful to CONAF (Chile) for the facilities provided in Chungara´. We thank Michael Ko¨hler (GFZ-Potsdam) for the thin sections preparation. SarahMetcalfe and Antje Schwalb are thanked for their reviews. References Abbott MB, Seltzer GO, Kelts K, Southon J. 1997. Holocene paleohy- drology of the tropical Andes from Lake Records. Quaternary Research 47: 70–80. Abbott MB, Wolfe PW, Aravena R, Wolfe AP, Seltzer GO. 2000. Holocene hydrological reconstructions from stable isotopes and paleolimnology, Cordillera Real, Bolivia. Quaternary Science Reviews 19: 1801–1820. Abbott MB, Wolfe BB, Wolfe AP, Seltzer GO, Aravena R, Mark BG, Polissar PJ, Rodwell DT, Rowe HD, Vuille M. 2003. Holocene paleohydrology and glacial history of the central Andes using multi- proxy lake sediment studies. Palaeogeography, Palaeoclimatology, Palaeoecology 194: 123–138. Amman C, Jenny B, Kammer K, Messerli B. 2001. Late Quaternary glacier response to humidity changes in the arid Andes of Chile (18-29- S). Palaeogeography, Palaeoclimatology, Palaeoecology 172: 313–326. Aravena R, Suzuki O, Pen˜a H, Pollastri A, Fuenzalida H, Grilli A. 1999. Isotopic composition and origin of the precipitation in northern Chile. Applied Geochemistry 14: 411–422. Argollo J, Mourguiart P. 2000. Late Quaternary climate history of the Bolivian Altiplano. Quaternary International 72: 37–51. Baied CA, Wheeler JC. 1993. Evolution of High Andean Puna ecosys- tem environment, climate and culture change over the last 12 000 years in the Central Andes.Mountain Research andDevelopment 13: 145–156. Baker PA, Seltzer GO, Fritz SC, Dunbar RB, Grove MJ, Tapia PM, Cross SL, Rowe HD, Broda JP. 2001a. The history of South American tropical precipitation for the past 25 000 years. Science 291: 640–643. Baker PA, Rigsby CA, Seltzer GO, Fritz SC, Lowenstein TK, Bacher NP, Veliz C. 2001b. Tropical climate changes at millennial and orbital timescales on the Bolivian Altiplano. Nature 409: 698–701. Battarbee RW, Jones VJ, Flower RJ, Cameron NG, Bennion H, Carvalho L, Juggins S. 2001. Diatoms. In Tracking Environmental Change Using Lake Sediments, Smol JP, Birks HJB, Last WM (eds). Kluwer Academic: Dordrecht, Netherlands; 155–202. Cheng H, Edwards RL, Hoff J, Gallup CD, Richards DA, Asmeron Y. 2000. The half-lives of uranium-234 and thorium-230. Chemical Geology 169: 17–33. Clavero JE, Sparks SJ, Huppert HE. 2002. Geological constraints on the emplacement mechanism of the Parinacota debris avalanche, north- ern Chile. Bulletin of Volcanology 64: 40–54. Clavero JE, Sparks SJ, Polanco E, Pringle M. 2004. Evolution of Par- inacota volcano, Central Andes, northern Chile. Revista Geolo´gica Chile 31: 317–347.  Natural Environment Research Council (NERC) copyright 2008. Reproduced with the permission of NERC. Published by John Wiley & Sons, Ltd. J. Quaternary Sci., Vol. 23(4) 351–363 (2008) DOI: 10.1002/jqs THE PALAEOHYDROLOGICAL EVOLUTION OF LAGO CHUNGARA´ 361 Clayton RN, Mayeda TK. 1963. The use of bromine pentafluoride in the extraction of oxygen from oxide and silicates for isotope analysis. Geochimica et Cosmochimica Acta 27: 43–52. Dorador C, Pardo R, Vila I. 2003. Variaciones temporales de para´me- tros fı´sicos, quı´micos y biolo´gicos de un lago de altura: el caso del Lago Chungara´. Revista Chilena de Historia Natural 76: 15–22. Edwards RL, Chen JH, Wasserburg GJ. 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. Francis PW, Wells G. 1988. Landsat thematic mapper observations of debris avalanche deposits in the Central Andes. Bulletin of Volca- nology 50: 258–278. Fritz SC, Baker PA, Lowenstein TK, Seltzer GO, Rigsby CA, Dwyer GS, Tapia PM, Arnold KK, Ku TL, Luo S. 2004. Hydrologic variation during the last 170 000 years in the southern hemisphere tropics of South America. Quaternary Research 61: 95–104. Fritz SC, Baker PA, Tapia P, Garland J. 2006. Spatial and temporal variation in cores from Lake Titicaca, Bolivia/Peru during the last 13 000 years. Quaternary International 158: 23–29. Garreaud RD, Vuille M, Clement AC. 2003. The climate of the Alti- plano: observed current conditions and mechanisms of past changes. Palaeogeography, Palaeoclimatology, Palaeoecology 194: 5–22. Giralt S, Moreno A, Bao R, Sa´ez A, Prego R, Valero BL, Pueyo JJ, Gonza´lez-Sampe´riz P, Taberner C. 2007. Statistical approach to disentangle environmental forcings in a lacustrine record: the Lago Chungara´ case (Chilean Altiplano). Journal of Palaeolimnology (in press). DOI: 10.1007/s10933-007-9151-9. Grosjean M, van Leeuwen JFN, van der Knaap WO, Geyh MA, Ammann B, Tanner W, Messerli B, Nu´n˜ez L, Valero-Garce´s BL, Veit H. 2001. A 22 000 14C year BP sediment and pollen record of climate change from Laguna Miscanti (231S), northern Chile. Global and Planetary Change 28: 35–51. Heegaard E, Birks HJB, Telford RJ. 2005. Relationships between cali- brated ages and depth in stratigraphical sequences: an estimation procedure by mixed-effect regression. The Holocene 15: 612– 618. Herna´ndez A, Bao R, Giralt S, Leng MJ, Barker PA, Pueyo JJ, Sa´ez A, Moreno A, Valero-Garce´s B, Sloane HJ. 2007. A high-resolution study of diatom oxygen isotopes in a Late Pleistocene to early Holocene laminated record from Lake Chungara´ (Andean Altiplano, Northern Chile). Geochimica et Cosmochimica Acta 71: A398. Herrera C, Pueyo JJ, Sa´ez A, Valero-Garce´s BL. 2006. Relacio´n de aguas superficiales y subterra´neas en el a´rea del lago Chungara´ y lagunas de Cotacotani, norte de Chile: un estudio isoto´pico. Revista Geolo´gica de Chile 33: 299–325. Hora J, Singer B,Wo¨rner G. 2007. Volcan eruption and evaporative flux on the thick curst of the Andean Central Volcanic Zone: 40Ar/39Ar constrains from Volca´n Parinacota, Chile. Geological Survey of America Bulletin 119: 343–362. Hua Q, Barbetti M. 2004. Review of tropospheric bomb C-14 data for carbon cycle modeling and age calibration purposes. Radiocarbon 46: 1273–1298. Jones V, Leng MJ, Solovieva N, Sloane H, Tarasov P. 2004. Holocene climate on the Kola Peninsula: evidence from the oxygen isotope record of diatom silica. Quaternary Science Reviews 23: 833–839. Juillet-Leclerc A. 1986. Cleaning process for diatomaceous samples. In 8th Diatom Symposium, Ricard M (ed.). Koeltz Scientific: Koenig- stein, Germany; 733–736. Leng MJ, Barker PA. 2006. A review of the oxygen isotope composition of lacustrine diatom silica for palaeoclimate reconstruction. Earth- Science Reviews 75: 5–27. Leng MJ, Marshall JD. 2004. Palaeoclimate interpretation of stable isotope data from lake sediment archives. Quaternary Science Reviews 23: 811–831. Leng MJ, Lamb AL, Heaton THE, Marshall JD, Wolfe BB, Jones MD, Holmes JA, Arrowsmith C. 2005a. Isotopes in lake sediments. In Isotopes in Palaeoenvironmental Research, Leng MJ (ed.). Springer: Dordrecht, Netherlands; 147–184. Leng MJ, Metcalfe SE, Davies SJ. 2005b. Investigating late Holocene climate variability in Central Mexico using carbon isotope ratios in organic materials and oxygen isotope ratios from diatom silica within lacustrine sediments. Journal of Palaeolimnology 34: 413–431. Matheney RK, Knauth LP. 1989. Oxygen-isotope fractionation between marine biogenic silica and seawater. Geochimica et Cosmochimica Acta 53: 3207–3214. Moreno A, Giralt S, Valero-Garce´s BL, Sa´ez A, Bao R, Prego R, Pueyo JJ, Gonza´lez-Sampe´riz P, Taberner C. 2007. A 13 kyr high-resolution record from the tropical Andes: the Chungara´ Lake sequence (188 S, northern Chilean Altiplano). Quaternary International 161: 4–21. Morley DW, LengMJ, Mackay AW, Sloane HJ, Rioual P, Battarbee RW. 2004. Cleaning of lake sediment samples for diatom oxygen isotope analysis. Journal of Paleolimnology 31: 391–401. Mu¨hlhauser H, Hrepic N, Mladinic P, Montecino V, Cabrera S. 1995. Water-quality and limnological features of a high-altitude Andean lake, Chungara´ in northern Chile. Revista Chilena deHistoria Natural 68: 341–349. Placzek C, Quade J, Patchett PJ. 2006. Geochronology and stratigraphy of late Pleistocene lake cycles on the southern Bolivian Altiplano: implications for causes of tropical climate change.GSA Bulletin 118: 515–532. Polissar PJ, Abbott MB, Shemesh A, Wolfe AP, Bradley RS. 2006. Holocene hydrologic balance of tropical South America from oxygen isotopes of lake sediment opal, Venezuelan Andes. Earth and Pla- netary Science Letters 242: 375–389. Reimer PJ, Baillie MGL, Bard E, Bayliss A, Beck JW, Bertrand CJH, Blackwell PG, Buck CE, Burr GS, Cutler KB, Damon PE, Edwards RL, Fairbanks RG, Friedrich M, Guilderson TP, Hogg AG, Hughen KA, Kromer B, McCormac G, Manning S, Ramsey CB, Reimer RW, Remmele S, Southon JR, Stuiver M, Talamo S, Taylor FW, van der Plicht J,Weyhenmeyer CE. 2004. IntCal04 terrestrial radiocarbon age calibration, 0-26cal kyr BP. Radiocarbon 46: 1029–1058. Reynolds CS. 2006. The Ecology of Phytoplankton. Cambridge Uni- versity Press: Cambridge, UK. Rings A, Lucke A, Schleser GH. 2004. A new method for the quanti- tative separation of diatom frustules from lake sediments. Limnology and Oceanography: Methods 2: 25–34. Ruttlant J, Fuenzalida H. 1991. Synoptic aspects of the central Chile rainfall variability associated with the Southern Oscillation. Inter- national Journal of Climatology 111: 63–76. Sa´ez A, Cabrera L. 2002. Sedimentological and palaeohydrological responses to tectonics and climate in a small, closed, lacustrine system: Oligocene As Pontes Basin (Spain). Sedimentology 49: 1073–1094. Sa´ez A, Valero-Garce´s BL, Moreno A, Bao R, Pueyo JJ, Gonza´lez- Sampe´riz P, Giralt S, Taberner C, Herrera C, Gibert RO. 2007. Volcanic controls on lacustrine sedimentation: the late Quaternary depositional evolution of lake Chungara´ (northern Chile). Sedimen- tology 54: 1191–1222. Schwalb A, Burns SJ, Kelts K. 1999. Holocene environments from stable isotope stratigraphy of ostracods and authigenic carbonate in Chilean Altiplano lakes. Palaeogeography, Palaeoclimatology, Palaeoecol- ogy 148: 153–168. Seltzer GO, Rodbell DT, Burns S. 2000. Isotopic evidence for late Quaternary climatic change in tropical South America. Geology 28: 35–38. Shemesh A, Charles CD, Fairbanks RG. 1992. Oxygen isotopes in biogenic silica: global changes in ocean temperature and isotopic composition. Science 256: 1434–1436. Shen CC, Edwards RL, Cheng H, Dorale JA, Thomas RB, Moran SB, Edmonds HN. 2002. Uranium and thorium isotopic and concen- tration measurements by magnetic sector inductively coupled plasma mass spectrometry. Chemical Geology 185: 165–178. Stuiver M, Reimer PJ, Bard E, Beck JW, Burr GS, Hughen KA, Kromer B, McCormac G, van der Plicht J, Spurk M. 1998. INTCAL98 radio- carbon age calibration, 24000-0 cal BP. Radiocarbon 40: 1041–1083. Tapia PM, Fritz SC, Baker PA, Seltzer GO, Dunbar RB. 2003. A late Quaternary diatom record of tropical climatic history from Lake Titicaca (Bolivia/Peru). Palaeogeography, Palaeoclimatology, Palaeoecology 194: 139–164. Thompson LG, Mosley-Thompson E, DansgaardW, Grootes PM. 1986. The Little Ice Age as recorded in the stratigraphy of the tropical Quelccaya Ice Cap. Science 234: 361–364. Valero-Garces BL, Grosjean M, Schwalb A, Schreir H, Kelts K, Messerli B. 2000. Late Quaternary lacustrine deposition in the Chilean Alti-  Natural Environment Research Council (NERC) copyright 2008. Reproduced with the permission of NERC. Published by John Wiley & Sons, Ltd. J. Quaternary Sci., Vol. 23(4) 351–363 (2008) DOI: 10.1002/jqs 362 JOURNAL OF QUATERNARY SCIENCE plano (188–288 S). In Lake Basins through Space and Time, Gierlowski-Kordesch E, Kelts K (eds). American Association of Petroleum Geologists Studies in Geology, no. 46; 625– 636. Valero-Garce´s BL, Delgado-Huertas A, Navas A, Edwards L, Schwalb A, Ratto N. 2003. Patterns of regional hydrological variability in central-southern Altiplano (188–268 S) lakes during the last 500 years. Palaeogeography, Palaeoclimatology, Palaeoecology 194: 319– 338. Vuille M. 1999. Atmospheric circulation over the Bolivian altiplano during dry and wet periods and extreme phases of the Southern Oscillation. International Journal of Climatology 19: 1579– 1600. Wolfe BB, Aravena R, Abbott MB, Seltzer GO, Gibson JJ. 2001. Reconstruction of paleohydrology and paleohumidity from oxygen isotope records in the Bolivian Andes. Palaeogeography, Palaeocli- matology, Palaeoecology 176: 177–192. Wo¨rner G, Hammerschmidt K, Henjes-Kunst F, Wilke H. 2000. Geo- chronology (40Ar/39Ar, K-Ar and He-exposure ages) of Cenozoic magmatic rocks from northern Chile (18–228 S): implications for magmatism and tectonic evolution of the central Andes. Revista Geolo´gica de Chile 27: 205–240.  Natural Environment Research Council (NERC) copyright 2008. Reproduced with the permission of NERC. Published by John Wiley & Sons, Ltd. J. Quaternary Sci., Vol. 23(4) 351–363 (2008) DOI: 10.1002/jqs THE PALAEOHYDROLOGICAL EVOLUTION OF LAGO CHUNGARA´ 363 ORIGINAL PAPER ENSO and solar activity signals from oxygen isotopes in diatom silica during late glacial-Holocene transition in Central Andes (18S) Armand Herna´ndez • Santiago Giralt • Roberto Bao • Alberto Sa´ez • Melanie J. Leng • Philip A. Barker Received: 8 April 2009 / Accepted: 1 February 2010  Springer Science+Business Media B.V. 2010 Abstract The late glacial-Holocene transition from the Lago Chungara´ sedimentary record in northern Chilean Altiplano (18S) is made up of laminated sediments composed of light-white and dark-green pluriannual couplets of diatomaceous ooze. Light-white sediment laminae accumulated during short-term extraordinary diatom blooms whereas dark-green sediment laminae represent the baseline limnological conditions during several years of deposition. Diatom oxygen isotope ratios (d18Odiatom) from 40 consecutive dark-green laminae, ranging from 11,990 to 11,450 cal year BP, show that a series of decadal-to-centennial dry–wet oscillations occurred. Dry periods are marked by relatively high isotope values whereas wet episodes are indicated by lower values. This interpretation agrees with the reconstructions of terrigenous inputs and regional effective moisture availability carried out in the lake but there is a systematic temporal disagreement between them owing to the non-linear response of the lacustrine ecosystem to environmental forcings. Furthermore, the d18Odiatom record tracks effective moisture changes at a centennial scale. Three major phases have been established (11,990–11,800, 11,800–11,550, and 11,550–11,450 cal year BP). Each phase is defined by an increasing isotope trend followed by a sudden depletion. In addition, several wet and dry events at a decadal scale are superim- posed onto these major trends. Spectral analyses of the d18Odiatom values suggest that cycles and events could have been triggered by both El Nin˜o-Southern Oscillation (ENSO) and solar activity. Significant ENSO frequencies of 7–9 years and 15–17 years, and periodicities of the solar activity cycles such as 11 years (Schwabe), 23 years (Hale) and 35 years (Bru¨ckner) have been recognised in the oxygen isotope time series. Time–frequency analysis shows that although solar and ENSO forcing were present at the onset of the Holocene, they were more intense during the late glacial period. The early Holocene might have been mainly governed by La Nin˜a-like A. Herna´ndez (&)  S. Giralt Institute of Earth Sciences Jaume Almera-CSIC, C/Lluı´s Sole´ i Sabarı´s s/n, 08028 Barcelona, Spain e-mail: ahernandez@ija.csic.es R. Bao Faculty of Sciences, University of A Corun˜a, Campus da Zapateira s/n, 15701 A Corun˜a, Spain A. Sa´ez Faculty of Geology, University of Barcelona, C/Martı´ Franque`s s/n, 08028 Barcelona, Spain M. J. Leng NERC Isotope Geosciences Laboratory, British Geological Survey, Nottingham NG12 5GG, UK M. J. Leng School of Geography, University of Nottingham, Nottingham NG7 2RD, UK P. A. Barker Lancaster Environment Centre, Lancaster University, Lancaster LA1 4YQ, UK 123 J Paleolimnol DOI 10.1007/s10933-010-9412-x conditions that correspond to wet conditions over the Andean Altiplano. Keywords Lake  Oxygen isotopes  Late glacial  Holocene  Andean altiplano  ENSO  Solar activity  Diatoms Introduction The study of Andean Altiplano lacustrine records plays a prominent role for interpreting the Quaternary palaeoclimatic history of the South American tropics and therefore for understanding the function of the tropics in the Earth’s climate system (Grosjean et al. 2001; Valero-Garce´s et al. 2003; Placzek et al. 2006) (Fig. 1). For this reason, studies on the sedimentary records from this area have increased in the last few decades. Most of these studies have focussed on the reconstruction of climatic events at millennial time scales, especially since the Last Glacial Maximum (Baker et al. 2001a). There is a general consensus that orbital forces are the main factor triggering the climatic conditions at a millennial-scale (Rowe et al. 2002; Placzek et al. 2006), and are therefore respon- sible for those climatic events. Superimposed onto this long term variability, changes in the hydrologic balance at a sub-millennial scale in the Andean Altiplano, have been attributed to the variability of the Pacific Sea Surface Temperatures (SSTs) and the strength of the zonal winds (Rowe et al. 2002; Garreaud et al. 2003). Both factors are controlled by El Nin˜o-Southern Oscillation (ENSO) and changes in the solar activity (Theissen et al. 2008). A number of studies have detected multidecadal- to centennial- scale hydrological balance shifts, suggesting that these relationships have been active since, at least, the mid-Holocene (Valero-Garce´s et al. 2003; The- issen et al. 2008). Diatom oxygen isotopes (d18Odiatom) are increas- ingly being used for palaeoenvironmental reconstruc- tions in lacustrine sedimentary records (Rietti-Shati et al. 1998; Barker et al. 2001). However, application of this proxy to high-resolution centennial to millennial lacustrine records is still in its infancy (Barker et al. 2007). d18Odiatom in decadal-to-centennial resolution palaeoclimatic reconstructions has not been utilised, mainly due to the difficulty in obtaining high resolution samples from sites with sufficient variation in d18Odiatom (outside of analytical error) that can be characterised at this fine temporal scale. Additional difficulties in using d18Odiatom are related to the difficulty in obtaining monospecific diatom samples in order to eliminate any species-specific effect variability, to acquire the nec- essary amount of sample from these short periods of time, and to have pure diatom samples, since significant contaminants can produce excursions in d18Odiatom that are similar to those produced by climate variations (Brewer et al. 2008). The diatomaceous ooze from Lago Chungara´ has previously been the subject of a preliminary diatom oxygen isotope study at low resolution. This earlier study was aimed at three non-consecutive stretches of the sedimentary record, and did not include all the dark-green laminae (Herna´ndez et al. 2008). For the present study we have analysed 40 consecutive dark- green laminae, corresponding to the late glacial and early Holocene, which represent a continuous record of the background limnological conditions (Herna´n- dez et al. 2008) (see the sedimentary model in the sedimentary sequence and rhythmite type section below). The excellent preservation and high diatom content of the record of Lago Chungara´ allow a detailed study of the regional moisture balance at decadal and centennial timescales. Fig. 1 Location of Lago Chungara´ on a South America rainfall rate map (mm/year) simplified from Negri et al. (2004). Main atmospheric systems are indicated. ICTZ: Intertropical Convergence Zone, SPCZ: South Pacific Conver- gence Zone J Paleolimnol 123 Here, we present the decadal-to-centennial time scale moisture balance reconstruction for the Andean Altiplano during the late glacial-Holocene transition (11,990–11,450 cal year BP) based on high-resolution analysis of d18Odiatom. This analysis was performed on successive and continuous 40 dark-green laminae of lacustrine sediments present in a core located in the offshore zone of Lago Chungara´. In order to support the interpretation, isotope data are compared with the reconstructions of the terrigenous inputs and inferred regional effective moisture in the Lago Chungara´ performed in the same core by Giralt et al. (2008). Lago Chungara´ setting Geology and hydrology Lago Chungara´ (188150S, 698100W, 4,520 m a.s.l.) is located in the Chilean Altiplano (Central Andes) lying in a highly active tectonic and volcanic context (Hora et al. 2007). The lake sits in the small hydrologically closed Chungara´ sub-Basin which was formed as a result of a debris avalanche during the partial collapse of the Parinacota Volcano, damming the former Lauca River (Fig. 2). Lago Chungara´ and Lagunas Cotaco- tani were formed almost immediately. However, the age of this collapse is not well constrained, with estimates ranging from 13,000 to 20,000 year BP (Hora et al. 2007). The lake has an irregular shape with a maximum length of 8.75 km, maximum water depth of 40 m, a surface area of 21.5 km2 and a volume of 400 9 106 m3 (Mu¨hlhauser et al. 1995; Herrera et al. 2006) (Fig. 2a). At present, the main inlet to the lake is the Chungara´ River (300–460 l s-1) although secondary streams enter to the lake in the south-western margin. Evapo- ration is the main water loss (3 9 107 m3 year-1), and groundwater outflow from Lago Chungara´ to Lagunas Cotacotani (6–7 9 106 m3 year-1, Dorador et al. 2003) represents about 20% of the total outflow. The calcu- lated residence time for the lakewater is approximately 15 years (Herrera et al. 2006). Water inputs to the lake have, on average, the following composition: 42 ppm HCO3 -, 3 ppm Cl-, 17 ppm SO4 2-, 7 ppm Na?, 4 ppm Mg2?, 8 ppm Ca2?, 3 ppm K? and 22 ppm Si. The Mg/Ca ratio of water inputs ranges from 0.22 to 0.71, depending on the local lithology of the catchment (Herrera et al. 2006; Sa´ez et al. 2007). Isotope data, temperature, pH, O2 and conductivity profiles of the lake water and the water inputs to the lake are shown in Table 1. The lake can be considered as polymictic and oligo- to meso-eutrophic (Mu¨hlhauser et al. 1995). The d18O and dD composition of the lake water reveal that it diverges from the Global Meteoric Water Line Fig. 2 a Location of the Chungara´-Cotacotani lake district (modified from Google Earth). b Cross section of sediments infilling Lago Chungara´. Position of core 11 is shown; note that the position of the core is projected in its equivalent position at the lake central plain. Arrows indicate major hydrological inputs and sedimentary contributions to the lake. Simplified from Sa´ez et al. (2007) J Paleolimnol 123 Table 1 Isotopic and other chemical and physical data analysed from water samples collected in the Lago Chungara´ (Modified from Herrera et al. 2006) Site Depth (m) Date d18O dD d13C Conductivity (lS cm-1) pH Temp (8C) O2 Lake Chungara´ 0 Nov-02 -1.52 -42.0 4.96 1,444 9.30 9.1 12.10 Chungara´ 20 Nov-02 -1.62 -42.6 4.26 1,433 9.17 7.2 7.80 Chungara´ 0 Jan-02 -3.85 -48.8 NA 1,262 9.63 17 NA Chungara´ 2 Nov-02 -1.54 -44 NA 1,423 9.23 10.2 9.10 Chungara´ 2 Nov-02 -1.58 -44.7 4.62 1,451 9.23 9.9 9.30 Chungara´ 5 Nov-02 -1.53 -43.2 5.52 1,462 9.17 9.5 8.90 Chungara´ 20 Nov-02 -1.59 -44.4 5.34 1,455 9.05 7.1 7.60 Chungara´ 0 Nov-02 -1.49 -44,2 5.43 1,473 9.20 12.1 11.90 Chungara´ 5 Nov-02 NA NA NA 1,461 NA 9.5 NA Chungara´ 10 Nov-02 NA NA NA 1,459 NA 9 NA Chungara´ 15 Nov-02 NA NA NA 1,458 NA 8.1 NA Chungara´ 20 Nov-02 NA NA NA 1,457 NA 7.6 NA Chungara´ 25 Nov-02 NA NA NA 1,457 NA 7 NA Chungara´ 31 Nov-02 -1.61 -45.8 -3.55 NA 9.02 NA NA Chungara´ 2 Nov-02 -1.50 -44.3 5.97 1,464 9.20 9.9 NA Chungara´ 21 Nov-02 -1.76 -41.7 3.68 1,473 9.08 7.6 NA Chungara´ 0 Nov-02 NA NA NA 1,464 NA 10.4 NA Chungara´ 5 Nov-02 NA NA NA 1,461 NA 9.6 NA Chungara´ 10 Nov-02 NA NA NA 1,461 NA 9.3 NA Chungara´ 15 Nov-02 NA NA NA 1,461 NA 9.1 NA Chungara´ 20 Nov-02 NA NA NA 1,456 9.15 7.2 NA Chungara´ 25 Nov-02 NA NA NA 1,456 NA 7 NA Chungara´ 30 Nov-02 NA NA NA 1,455 NA 6.8 NA Chungara´ 35 Nov-02 NA NA NA 1,454 NA 6.4 NA Chungara´ 0 Nov-02 -1.56 -39.9 3.21 1,439 9.08 9 11.20 Chungara´ 15 Nov-02 -1.54 -43.4 5.17 1,438 8.99 8.3 8.30 Chungara´ 33 Nov-02 -1.58 -43.4 4.56 1,435 9.10 6.2 6.20 Chungara´ 0 Nov-02 -2.05 -44.4 7.03 1,388 9.22 10.7 NA Chungara´ 0 Jan-02 -3.39 -46.8 6.40 1,400 9.19 11.4 NA Chungara´ 0 Jun-06 NA NA NA 1,463 9.42 5 10.50 Chungara´ 12 Jan-02 NA NA 8.10 NA NA NA NA Chungara´ 0 Jun-06 NA NA NA 1,493 9.70 2 9.30 Chungara´ 0 Jun-06 NA NA NA 1,418 9.48 9 13.00 Chungara´ 0 Jun-06 NA NA NA 220 9.70 5 25.00 Springs Bofedal 0 Jan-02 -14.98 -116 -2.20 52 7.15 7.9 NA Bofedal 0 Nov-02 -16.49 -116.6 -7.04 48.7 7.07 7.8 NA Mal paso 0 Jan-02 -15.58 -120.5 NA 46.8 7.31 9.7 NA Mal paso 0 Nov-02 -17.04 -121.9 -8.97 51.2 6.97 10/11.6 NA Mal paso 0 Jan-04 -17.3 -121.3 NA 122 7.71 10.8 NA Ajata 0 Jan-02 -14.07 -106.2 NA 59.8 7.80 7.6 NA J Paleolimnol 123 (GMWL) and the Regional Meteoric Line (RML). This divergence can be attributed to the enrichment of the lake water by evaporation with regard to rainfall and springwater (Fig. 3a; Herrera et al. 2006). The mean lake water values of d18O and dD (January 2002 to January 2004) are -1.4% SMOW and -43.4% SMOW, respectively (Table 1). Climate Climate in the Chungara´-Cotacotani lake district is dominated by arid conditions due to the influence of the South Pacific Anticyclone (Fig. 1). Modern mean annual temperature at Lago Chungara´ is ?4.2C. Annual rainfall ranges from 100 to 750 mm year-1 (mean 411 mm year-1), losing 1,200 mm year-1 via evapo- ration, which exceeds precipitation (Fig. 3b; Valero- Garce´s et al. 2000). The lake region shows pronounced seasonal contrasts due to the dominance of the tropical summer moisture (Garreaud et al. 2003), known as the South American Summer Monsoon (SASM) (Vuille and Werner 2005). Regional moisture originates from the tropical Atlantic Ocean and is transported to the Altiplano throughout Amazonia during the summer months (DJFM). During these months weak easterly flow prevails over the Altiplano as a consequence of the southward migration of the subtropical jet stream and the establishment of the Bolivian high-pressure system (Garreaud et al. 2003). This climatic situation defines this time window as the wet season in the Altiplano accounting for more than 70% of the annual precipi- tation (Fig. 3b). The SASM is a major component of the climate system over tropical and subtropical South America during the austral summer and is remotely forced by tropical Pacific SSTs (Vuille and Werner 2005). Thus, the inter-annual to decadal climate variability is currently related to ENSO-like variations over the Pacific basin (Valero-Garce´s et al. 2003). Instrumental data from the Chungara´ area show a reduction of the precipitation during moderate to intense El Nin˜o years (Fig. 3c). However, there is no direct relationship between the relative El Nin˜o strength and the amount of rainfall reduction (Valero-Garce´s et al. 2003). Furthermore, on longer timescales, it is speculated that changes in tropical- Atlantic meridional SST gradients also force precip- itation variability on the Altiplano (Baker et al. 2001b). Rainfall isotope composition in Central Andes (Fig. 3d) is characterised by a large variability in d18O (between ?1.2 and -21.1%) and ofdD (between ?22.5 and -160.1%). The origin of the lightest oxygen isotope values is the strong fractionation in the air masses from the Amazon and is directly related to higher rainfall intensity (‘amount effect’) (Herrera et al. 2006). However, the rainfall oxygen isotope composition in the Chungara´-Cotacotani lake district only oscillates by 6%, between -14 and -20%, with a mean value of -14.3% (Fig. 3a, d). Sedimentary sequence and rhythmite type The sedimentary infill of Lago Chungara´ was charac- terised by the lithological description of fifteen lake cores obtained in 2002 (Sa´ez et al. 2007) and by seismic imagery obtained in 1993 (Valero-Garce´s et al. 2000; Sa´ez et al. 2007). From the bottom to the top of core 11, Table 1 continued Site Depth (m) Date d18O dD d13C Conductivity (lS cm-1) pH Temp (8C) O2 Ajata 0 Nov-02 -15.28 -205.9 -6.86 50 7.46 5.3 NA Canal 0 Nov-02 -17.12 -121.7 -10.14 66.7 8.09 10 NA Canal 0 Jan-04 -17.26 -119.3 NA 380 7.62 9.4 NA Colada Ajata 0 Jan-02 -14.88 -111 -2.80 155.1 5.82 4.9 NA Colada Ajata 0 Nov-02 -16.23 -112.1 -2.69 136.1 5.94 4.5 NA River Chungara´ 0 Jan-02 -14.82 -111.8 NA 241 9.28 16 NA Chungara´ 0 Nov-02 -16.09 -113.1 -1.15 223 8.99 17 NA Chungara´ 0 Jan-04 -16.27 -114.1 NA 316 8.02 14.2 NA NA not available data J Paleolimnol 123 two sedimentary units (units 1 and 2) were identified and correlated over the offshore zone of the lake mainly using tephra keybeds (Figs. 2a, 2b). Unit 1 is made up of diatomaceous ooze with variable types and quantities of carbonates (calcite, aragonite) and amorphous organic matter. It is continuous across the lake, although thickest in the NW sector of the central plain and thins towards the south and west, probably overlapping the Miocene substrate. Unit 1 occurs in the central plain and the sharply rising flank of the lake (Fig. 2b) and is divided in two subunits. Subunit 1a is composed by light-white and dark-green diatomaceous ooze couplets and a rhythmite type was defined (Herna´ndez et al. 2008). Light-white laminae are formed by the skeletons of the diatom Cyclostephanos andinus (Theriot, Carney, and Richerson) Tapia, Theriot, Fritz, Cruces and Riv. Dark-green laminae, with higher organic matter content, are made up by a mixture of diatoms, including the euplanktonic Cyclo- stephanos andinus, although diatoms of the Cyclotella stelligera complex are co-dominant taxa. Subdominant groups are some tychoplanktonic (Fragilaria spp.) and benthic taxa (Cocconeis spp., Achnanthes spp., Navic- ula spp., Nitzschia spp.). Subunit 1b is composed of centimetre- to decimetre-thick laminated brown dia- tomaceous ooze and endogenic carbonates that occur in low concentrations. Unit 2 is about 6 m-thick and grades laterally to the west and south into alluvial and Fig. 3 a Isotope values (d18O/dD) from rainfall, Lago Chungara´, Rı´o Chungara´ and studied area springs. GMWL: Global Meteoric Water Line; LML: Local Meteoric Line; EL: Evaporation Line. Note the isotopic enrichment of the lake water by evaporation with regard to rainfall and springwater. b Mean monthly rainfall (mm) and temperature (C) at Chungara´ meteorological station (18.17S, 69.08W, 4,500 m a.s.l.). Note the seasonality of both parameters. c Annual rainfall in the Chungara´ region from 1962 to 1994. The arrows indicate strong El Nin˜o years. Modified from Valero-Garce´s et al. (2003). d Isotope composition (d18O/dD) of rainfall samples from La Paz (Bolivia) obtained by the International Atomic Energy Agency (IAEA) since 1995 until 1998 (blue squares and black triangles), and the samples of Lago Chungara´ and the very close Lagunas de Cotacotani obtained from Herrera et al. (2006) (red circles). LML: Local Meteoric Line (dD = 7.9d18O ? 14). Note the variability in d18O and of dD values J Paleolimnol 123 deltaic deposits, and towards the east into macrophyte, organic-rich facies (Fig. 2b). It is mainly composed of massive to slightly banded diatomaceous ooze inter- bedding with 13 tephra layers. Unit 2 is also divided in two subunits. Subunit 2a is composed of brownish-red massive to slightly banded sapropelic diatomaceous ooze with common calcitic crystals (silt grain-sized) and carbonate-rich layers. Subunit 2b consists of dark- grey diatomaceous ooze with frequent macrophyte remains alternating with massive black tephra layers, mainly composed of plagioclase, glass and mafic minerals (Sa´ez et al. 2007; Moreno et al. 2007). The chronological model for the sedimentary sequence of Lago Chungara´ is based on 17 14C AMS dates obtained from bulk organic matter from the central plain cores and aquatic organic macrofos- sils picked from littoral cores, and one 238U/230Th date from carbonate. Details on the construction of the chronological framework are discussed elsewhere (Giralt et al. 2008). According to the chronological model, the studied interval records the late glacial- Holocene transition (11,990–11,450 cal year BP) and each couplet was deposited during time intervals ranging from 4 to 24 years. Light-white sediment laminae accumulated during short term diatom blooms (occurring from days to weeks) whereas dark-green sediment laminae represent the baseline limnological conditions during several years of deposition (Herna´ndez et al. 2008). Previous work has characterised the surface and underground waters from the Chungara´ and Cotaco- tani lake district (Herrera et al. 2006), as well as the sediments of Lago Chungara´. X-Ray Fluorescence (XRF), X-Ray Diffraction (XRD), Total Carbon and Total Organic Carbon (TC and TOC), Total Biogenic Silica (TBSi), pollen and diatom analyses were performed. These multiproxy studies have allowed us to establish the sedimentary, hydrological and environmental evolution of the Lago Chungara´ at different time scales (Sa´ez et al. 2007; Moreno et al. 2007; Giralt et al. 2008). Methods An interval of 46.5 cm from the laminated dark-green and light-white sediments of Subunit 1a (deposited between 11,990 and 11,450 cal year BP) was selected and sampled, lamina by lamina, from core 11 (Fig. 2). All the dark-green laminae (40 samples) of this interval were selected for d18Odiatom analyses to carry out a very high-resolution study of the baseline environmental evolution of Lago Chungara´ (sampling ranging from 4.1 to 24.4 years; average temporal resolution is ca. 12 years) during the late glacial and early Holocene transition period. Of these, 22 samples were also previously used in a lower resolution study (Herna´ndez et al. 2008). The thickness of the dark- green laminae sampled ranges between 2 and 9 mm. Analysis of the oxygen isotope composition of diatom silica requires the material to be almost pure diatomite (Juillet-Leclerc 1986). Our samples were treated following the method proposed by Morley et al. (2004) with some variations (Herna´ndez et al. 2008). The samples were treated to remove organics and carbonate, then sieved at 125 lm to eliminate resistant charcoal and terrigenous particles. The 63- and 38-lm sieves were used to obtain a diatom concentrate made up almost exclusively by large valves of the centric diatom Cyclostephanos andinus, eliminating in the samples any species-specific effect variability (Fig. 4). Gravity settling in a water column during the sieving process also helped to remove any remaining tephra and clay particles. The Gravitational Split-flow Thin Fractionation (SPLITT) was then applied to the most problematic samples, at Lancaster University (UK), as an alternative approach to heavy liquid separation (Rings et al. 2004; Leng and Barker Fig. 4 Diatom-rich sediment from Lago Chungara´ after the cleaning process. Large Cyclostephanos andinus valves are the unique component J Paleolimnol 123 2006). Finally, the purified diatom samples were dried at 40C between 24 h and 48 h. After the cleaning process, all the samples were checked under the light microscope and some of them also with XRD and scanning electron microscope (SEM), as well as analysed for TC to verify that they did not contain any significant amount of terrigenous matter (Fig. 4). For oxygen isotope analysis, a stepwise fluorina- tion method was applied to 5–10 mg of the purified diatoms in order to strip the frustule hydrous layer before a full reaction with BrF5 (Leng and Barker 2006). The oxygen liberated was then converted to CO2 using the method of Clayton and Mayeda (1963), measured by IRMS and normalised against NBS standards. A random selection of 7 samples were analysed in duplicate or triplicate giving a mean reproducibility value of \0.2% (1r), only one sample gave a reproducibility value of 0.4% (1r). The fluorination process and the 18O/16O ratios measured were carried out at the NERC Isotope Geosciences Laboratory, British Geological Survey (UK). We employed two methods of spectral analyses to examine any periodic components in the d18Odiatom values: Multi-Taper Method (MTM) and Time– Frequency analysis. These two spectral analyses allowed us to examine statistically significant signals in the time series in both the frequency and time domains. MTM provided both a means of spectral estimation and signal reconstruction for time series with spectra that contain both singular and continuous components (Theissen et al. 2008). Time–Frequency (TF) analysis is a hybrid tool between the Fourier Transform and wavelets that intends to use a localised spectrum. For that, this analysis does not use a fixed-size Gaussian window but a Gaussian window that adapts to the spectrum (Stockwell et al. 1996). All the statistical treatments of the datasets were performed using the R software package (R Devel- opment Core Team 2008). Results Oxygen isotopes The d18Odiatom record (Fig. 5d) shows both short- term (decadal) and long-term oscillations (centennial time scales) ranging from ?35% to ?39.2% (mean = ?37.4 ± 0.8%). From the bottom to the top, the studied record can be subdivided into three phases. These intervals correspond to three enrich- ment/depletion phases (Fig. 5c). Each phase starts with a continuous centennial isotope enrichment which abruptly ends with a sharp depletion: Phase 1. Lower interval (11,990–11,800 cal year BP). It shows the maximum and minimum d18Odiatom values (?39.2% and ?35.1% respec- tively, with a mean value of ?37.7 ± 1%) throughout the whole record. It starts with an increasing trend of *3.3%/100 year which fin- ishes at 11,860 cal year BP. This trend is followed by a shift to lighter values of *8.1%/100 year with a sharp final decrease in the d18Odiatom values of 3.5% in less than 10 years, acquiring the minimum value for the whole record at ca. 11,800 cal year BP. Both trends are interrupted by ca. 5–20 years depletion/enrichment excursions ranging between ± 0.9 and ± 1.7%. Phase 2. Middle interval (11,800–11,550 cal year BP). This section (mean value ?37.3 ± 0.7%) starts with an enrichment trend showing an upwards gradient of *1.3%/100 year which fin- ishes at 11,570 cal year BP with a ?38.3% d18Odiatom value. This trend is however punctuated by one sudden rise (?2.3%) and up to four minor depletions (ranging from -0.6 to -1.3%) of the d18Odiatom values on a 40–55 years basis. The enrichment trend is followed by a shift of *9.1%/100 year to lighter values reaching a minimum value of ?36.2%. Phase 3. Upper interval (11,550–11,450 cal year BP). This interval (mean value ?37.2 ± 0.7%) also starts with an enrichment trend but, because the section only comprises three samples, this enrichment has not been estimated. This trend is also followed by depletion of 1.3% in 10 years. Spectral analyses of the diatom oxygen isotope record Multi-taper analysis (MTM) performed on the d18Odiatom values shows a number of clear periodic- ities (Fig. 6a). Almost all identified periodicities (7.2, 8.9, 11.1, 13, 18.6, 22.3 and 39.4 years) exceed the 99% confidence interval whereas only two (3.7 and J Paleolimnol 123 8 years) lie between 95 and 99% confidence interval (Fig. 6a). Most of the sub-decadal identified frequen- cies are close to the minimum temporal resolution of the sampling (4.1 years), which explains in great part the weaker intensity of the short periodicities between 3 and 8 years. Therefore, only the most significant frequencies and above the minimum temporal sampling resolution have been taken into account in the discussion. Time–Frequency (TF) analysis reveals the stron- gest energy for the lower values of frequency, mainly focussed on the 35-years cycles, whereas it decreases towards higher frequency values, i. e., the higher periodicities (Fig. 6b). This fact can mostly be explained by the decadal sampling resolution, making periodicities lower than 10 years less significant. Additionally, TF analysis indicates that the highest energies of the significant frequencies are located in the late glacial period between ca. 11,950 and 11,700 cal year BP, decreasing just from the onset of the Holocene until, at least, approximately 11,550 cal year BP (Fig. 6b). TF diagram also high- lights that the identified frequencies did not have the same intensity (energy) during all the studied period. For instance, the shortest significant periodicity observed in the MTM (7.2 years) was mainly active during the first 150 years of the record, whereas it was only active during three short time windows in the following 500 years. A similar pattern is also observed for the rest of the significant periodicities (8.9, 11.1, 13, 18.6, 22.3 and 39.4 years). The maximum energy areas correspond to depletions in the d18Odiatom values, i.e. 11,800 and 11,550 cal year BP (Fig. 6b). Fig. 5 d18Odiatom data for the period 11,990–10,475 cal year BP from Lago Chungara´, compared with other paleoenviron- mental records of the lake. a Planktonic diatoms percent abundance curve for the whole Lago Chungara´ sequence (Sa´ez et al. 2007). b d18Odiatom data of non-consecutive dark-green laminae from three intervals of the record (Herna´ndez et al. 2008). c Photography of laminated sediments corresponding to the sampled interval of subunit 1a in core 11. d Oxygen isotope enrichment/depletion phases, in the studied interval, inter- preted from the data. e d18Odiatom data from the present study and interpretation in terms of wet and dry conditions. The values correspond to 40 consecutive dark-green laminae throughout the whole selected interval. f Terrigenous input variations derived from the first eigenvector of Principal Component Analysis (PCA) on magnetic susceptibility, X-Ray Fluorescence (XRF), X-Ray Diffraction (XRD), Total Carbon and Total Organic Carbon (TC and TOC), Total Biogenic Silica (TBSi) (Giralt et al. 2008). g Effective moisture availability variations from the second eigenvector of the mentioned PCA (Giralt et al. 2008). Correlation lines correspond to the main oxygen isotope depletion peaks. Note that the main trends of the three curves are similar but there is a systematic temporal disagreement between them J Paleolimnol 123 Discussion Controlling factors of d18Odiatom in Lago Chungara´ d18Odiatom in lake sediments is controlled by the oxygen isotope composition of the lake water (d18Olakewater), temperature, and the possible disequilibrium by vital effects or diagenesis (Leng and Barker 2006). We discount vital effects and diagenesis as analyses were made on near-monospecific diatom samples and pres- ervation of the diatom frustules is excellent (Fig. 4). d18Olakewater depends on the balance between the isotope composition of water inputs (including the Fig. 6 a Multi-taper analysis of the d18Odiatom values. The 90, 95 and 99% confidence levels are indicated and significant periodicities are shown. Note that periodicities with more than 99% of significance are shown in black and those with more than 95% significance in blue. b Time–Frequency analysis of the d18Odiatom values. Pink indicates high energy whereas blue displays low energy areas. Energies below 0.03 were clipped in order to facilitate understanding of the graph. Red and blue horizontal bands mark different frequency bands of the ENSO and solar activity forcings. Yellow vertical bands show zones with d18Odiatom shifts and their corresponding power values for each frequency. A weakening pattern in ENSO and solar activity energies can be observed through the late glacial-early Holocene transition J Paleolimnol 123 source and amount of precipitation, surface runoff and groundwater inflow) and outputs (evaporation and groundwater loss) in the lake. The measured d18O of the inputs (springs, Rı´o Chungara´ and rainfall) in the Lago Chungara´ is homogeneous, giving values close to the isotope composition of precipitation (d18Oprecipitation) (Fig. 3a). d 18Oprecipitation is a func- tion of the isotope composition of the moisture source and air-mass trajectory, but in the Lago Chungara´ there are no changes in the moisture source compo- sition since the air masses always come from the Atlantic Ocean throughout the Amazon basin (Gros- jean et al. 1997). During moisture transport from the Atlantic to the lake area, three processes are directly responsible for the low and variable values of the present d18Oprecipitation throughout the Andean Alti- plano (Aravena et al. 1999). These processes include interaction of the air masses within the Amazon basin, an altitude effect due to the ascent of the air masses along the eastern slope of the Andes, and the convective nature of the storms in the Altiplano region. Nevertheless, in the Lago Chungara´ region the values obtained for the measured d18Oprecipitation are relatively stable with almost all values around -14 and -20% (Figs. 3a, 3d; Herrera et al. 2006), whereas d18Olakewater is much higher (Fig. 3a). This result is in accordance with a d18Olakewater enrichment via evaporation. Thus, any isotopic variation of d18Olakewater will be more related to changes in the amount of precipitation (‘‘amount effect’’) and evap- oration rather than to the variability of d18Oprecipitation. Evaporation enriches d18O of lake water by 14% relative to the inlets (precipitation, springs and river) in the present day (Figs. 3a, 3b; Table 1). During the late glacial and early Holocene the water residence time of the lake was shorter than present because of the different palaeohydrological context, but even so it can be considered closed for that period (Herna´n- dez et al. 2008). Accordingly, the variations in the d18Odiatom must be mainly derived from changes in the d18Olakewater resulted from shifts in the balance between the precipitation and the evaporation (P-E), rather than dominated by temperature. However, two factors should be considered in the interpretation of the d18Odiatom values in terms of temperature oscillations. The first factor is related to d18Oprecipitation that correlates directly with changes in the air tempera- ture. The global relationship between changes in d18Oprecipitation with air temperature is commonly referred to as the ‘Dansgaard relationship’, and it implies changes between ?0.2 and ?0.7%/8C (Dansgaard 1964). The second is the lakewater temperature dependence of oxygen isotope fraction- ation between diatom silica and the lake water (Brandriss et al. 1998). Nevertheless, the fraction- ation factor value of this temperature dependence is still controversial. Published fractionation factors range from -0.2% and -0.5%/C (Brandriss et al. 1998; Moschen et al. 2005). The two temperature factors have opposing effects on d18Odiatom but, owing to its larger variability, the effects of the first factor (air temperature) usually dominate over the second. However, even in the case of the largest change due to the Dansgaard relation- ship, its magnitude will be greatly damped by the effect of the isotope fractionation between diatom silica and lake water. Moreover, it is known that most of the tropical rainfall isotope datasets exhibit a far stronger correlation with total precipitation than with air temperature (Leng et al. 2005), indicating in the Lago Chungara´ case a magnification of the P-E balance in wetter periods. Hence, we can assume that in the Lago Chungara´ the effects of precipitation variability and tempera- ture oscillations in the d18Odiatom values will be small in comparison to evaporative concentration, as pointed by other authors for closed lakes in general (Gat 1980; Gasse and Fontes 1992). Variations of the precipitation-evaporation balance in the lake Oxygen isotopes have widely been used to carry out lake level reconstructions and to establish consequent palaeoclimatic interpretations (Barker et al. 2001; Valero-Garce´s et al. 2003). There is a relationship between lake level change and the P-E balance for Lago Chungara´ during the late glacial and early Holocene, but this dependency is hampered by local palaeohydrological factors such as changes in the groundwater outflow and shifts in the lake surface/volume ratio which produce a background long term enrichment trend (Herna´ndez et al. 2008). This effect is however negligible when considering isotopic changes at a decadal to centen- nial time scale. Both present (Fig. 3) and past (Thompson et al. 1998) rainfall isotope values in J Paleolimnol 123 the Lago Chungara´ region are much lighter than those measured for the lakewater, and the magnitude of the long-term enrichment trend is very small compared to them. Therefore, depletions of d18Odiatom would directly be related to wet episodes in the Andean Altiplano, whereas exceptionally high values, which stand out over the general enrichment trend, would indicate arid episodes. The observed d18Odiatom enrichment trends agree with periods where light-white laminae are more common, whereas depletion episodes coincide with poorly developed and less abundant light-white laminae (Figs. 5c, 5d). These light-white laminae are most likely the result of exceptional periods of mixing of the shallow water column during low- stands, which recycle nutrients from the hypolimnion and therefore trigger extraordinary diatom blooms (Herna´ndez et al. 2007). This interpretation is also supported by terrigenous input and regional effective moisture reconstructions previously performed on the Lago Chungara´ sedimentary record (Giralt et al. 2008) (Figs. 5f, 5g). These reconstructions were carried out by applying multivariate statistical anal- yses (Cluster, Redundancy Analysis (RDA) and Principal Component Analysis (PCA)) to magnetic susceptibility, XRF, XRD, TC, TOC, TBSi and grey- colour curve data. The terrigenous inputs curve was derived from the first eigenvector of the PCA, whereas the regional effective moisture reconstruc- tion was obtained from the second eigenvector. For the lower part of Chungara sequence (Unit 1), the more positive values of the terrigenous inputs were interpreted, as increasing erosion rate of catchment volcanic sediments, suggesting humid conditions. Similarly, the effective moisture availability proxy depends on the P-E balance, with positive values corresponding to drier conditions (Giralt et al. 2008). The comparison of the three proxies (Figs. 5e, 5f, 5g) shows that the hydrological response of the diatom silica oxygen isotopes (a biological proxy) and of the other two reconstructions to the environmental vari- ations is not the same. The main trends in the three curves (Figs. 5e, 5f, 5g) are similar but there is a systematic temporal disagreement (ranging between ca. 5 and 50 cal year BP) between the terrigenous inputs and the regional effective moisture availability (which both react first) and the d18Odiatom (reacting afterwards). This time lag between the two proxies highlights the complex and non-linear response of the lacustrine ecosystem to environmental forcings (Fritz 2008). After rainfall the increased runoff and input of terrigenous material is almost immediate. On the contrary, the oxygen isotope homogenisation of the lakewater which later will be incorporated on the diatom frustule, has a delayed time of response. This depends on the epilimnion water residence time and, furthermore, whether the lake is hydrologically closed or not. Hence, the observed time lag can be showing these different responses of the system to the same forcing. However, we cannot discount the poorly understood concept of silica maturation, where pores in the silica matrix close through early diagenesis creating differ- ences in the d18O between living diatoms and sediment assemblages (Schmidt 2001) and therefore a lag in the d18Odiatom record. At centennial scale, the Lago Chungara´ isotopic values show a general pattern of increasing d18Odiatom (Fig. 5b), with an enhanced enrichment period at the bottom, but interrupted by three major depletion events. The depletion events, accentuated by the ‘‘amount effect’’, correspond to heavy rainfall condi- tions, whereas enriched values would indicate excep- tionally dry conditions favouring the evaporation. This interpretation is reinforced by the terrigenous input and effective regional moisture availability independent reconstructions. Three wet/dry phases have been identified in the d18Odiatom record (Fig. 5d). Phase 1 (11,990– 11,800 cal year BP) shows a significantly increased gradient in d18Odiatom suggesting that dominantly dry climate conditions played a key role triggering this isotope enrichment. Because of this drier situation the lake level would be lower, as also indicated by the important development and major presence of light- white laminae in this part of the interval. Three low- intensity and short-term wet episodes punctuate the established late glacial arid period (Fig. 5e). These episodes can also be recognised and correlated with events of increased terrigenous inputs and effective moisture availability (Figs. 5e, 5f, 5g). The much weaker isotope enrichment for phase 2 (11,800 and 11,550 cal year BP) can be mainly ascribed to the general low magnitude palaeohyd- ological background trend towards heavier isotope conditions of the late glacial-early Holocene transi- tion (Herna´ndez et al. 2008). This fact, together with the poorer development and minor presence of the J Paleolimnol 123 light-white laminae with respect to the previous interval, suggests that the enrichment via evaporation was much less important than during the sedimenta- tion of phase 1, corresponding to a more humid period. Furthermore, the terrigenous inputs and effective regional moisture availability curves show relatively wetter conditions for this period (Figs. 5f, 5g). This trend is also punctuated by a sudden rise in the lowest part of the interval indicating a short dry event and slight depletions in d18Odiatom indicating wet decadal-scale events (Fig. 5e). In phase 3 any clear trend is difficult to identify (Fig. 5e). Although the d18Odiatom record seems to show a new trend towards drier conditions after the sudden wet event dated at 11,550 cal year BP, the lack of suitable samples has hampered any firm conclusions. Long-term, centennial- to millennial-scale palaeoclimatic implications There are many Late Quaternary palaeoclimatic reconstructions from the Andean Altiplano region (Sylvestre et al. 1999; Rigsby et al. 2005) but the climatic context for the late glacial-Holocene transi- tion still remains unclear. Some authors have defined a cold period (12,600–11,500 cal year BP) coincident with the Northern hemisphere’s Younger Dryas event (Baker et al. 2001b). The wet (‘‘Coipasa phase’’, Thompson et al. 1998; Placzek et al. 2006) or dry (Maslin and Burns 2000; Weng et al. 2006) character of this event remains controversial. On the contrary, other authors consider this period just the final part of the deglaciation towards the present interglacial (‘‘Tican˜a phase’’, Sylvestre et al. 1999), as part of a long-term dry pattern (Rowe et al. 2002; Abbott et al. 2003). The previous lake level reconstruction, mainly based on the abundance of planktonic diatoms, shows a shallowing followed by a long term rising trend for the interval presented here (Sa´ez et al. 2007). Additionally, recent data on the Lago Chungara´ record, mainly based on XRF core scanner analysis, has established the late glacial to Holocene transition as a relatively wet period (Giralt et al. 2008). The centennial scale d18Odiatom record is congruent with the lake level reconstruction performed by Sa´ez et al. (2007) which represents the palaeoclimatic evolution related to the major lake level variations (Fig. 5a). The non-continuous isotopic data (Fig. 5b) also displays a persistent, but minor, background isotope enrichment trend. This enrichment is related to changes in the lake morphology due to shifts in its surface/volume ratio, as well as changes in the groundwater outflow during the lake ontogeny (Herna´ndez et al. 2008). In any case, the new d18Odiatom data presented here highlights that the glacial-interglacial transition in the central Andean Altiplano was punctuated by abrupt and high-fre- quency centennial climatic variability. Short-term, decadal- to centennial-scale palaeoclimatic implications Millennial-scale shifts in the Atlantic-Amazon-Alti- plano hydrologic system have been attributed to orbitally induced changes in solar insolation, coupled with long-term changes in the ENSO variability (Rowe et al. 2002; Abbott et al. 2003; Servant and Servant-Vildary 2003). However, higher-resolution changes are not directly related to orbitally induced insolation forcing (Abbott et al. 2003). The interan- nual climate variability in the Andean Altiplano is most likely related to changes in the Pacific Tropical SSTs, and the sign and strength of the zonal winds above the Altiplano (Garreaud et al. 2003). Both factors would affect the strength and position of the Bolivian high and, hence, the moisture distribution over the region. The main force controlling the SSTs is the ENSO variability, involving dry or wet situations in the Altiplano during El Nin˜o- or La Nin˜a-like conditions respectively (Garreaud et al. 2003; Vuille and Werner 2005). This is consistent with instrumental data from the Chungara´ area where precipitation is reduced during moderate to intense El Nin˜o years (1965, 1972, 1983, and 1992) (Fig. 3c). Additionally, the sign and strength of the zonal winds above the Altiplano would be modulated by decadal and multidecadal variations in solar activity, possibly related to the mode of the ENSO system (Theissen et al. 2008). Although ENSO modulation by solar activity has been suggested (Velasco and Mendoza 2008), no clear relationship has been demonstrated between both forcings. Nevertheless, there is broad agreement that ENSO events are the main control governing the moisture distribution in the Altiplano (Servant and Servant-Vildary 2003), and that dec- adal-scale changes in the effective moisture could be J Paleolimnol 123 related to the solar activity during the mid-Holocene (Theissen et al. 2008). The results presented here would suggest a similar pattern during the late glacial-Holocene transition over the Andean Altiplano (Fig. 6b). The identified frequencies can be attributed to different periodicities of the solar activity cycles such as Schwabe 11 years (identified as 11.1 and 13 years), Hale 23 years (22.3 years) and Bru¨ckner 35 years (39.4 years), and of the ENSO frequency (main frequency at 7–9 years (7.2 and 8.9 years) and its decadal fre- quency 15–17 years (18.6 years)). The influence of solar activity and ENSO variability on the isotope record is supported by the fact that several period- icities concordant with both forces were identified. The time–frequency analysis suggests that the driest period (11,950–11,800 cal year BP) was ruled by high solar activity, mainly represented by a Bru¨ckner cycle, and strong ENSO-like conditions. The ENSO and solar activity signals remain present for the early Holocene period (between 11,750 until 11,500 cal year BP), although they show a weakening pattern through this period (Fig. 5b). This fact is congruent with the progressive weakening of the ENSO suggested by other authors for the late glacial-Holocene transition (Rodbell et al. 1999; Moy et al. 2002; Rodo´ and Rodrı´guez-Arias 2004). In Lago Chungara´, the onset of the Holocene was characterised by minor d18Odiatom enrichment by evaporation and by the occurrence of multi-decadal weak depletions that would be governed by the more humid La Nin˜a-like conditions. This would agree with previous observations that suggest a reduction in the El Nin˜o intensity within the region during the early-Holocene in favour of long-term La Nin˜a-like conditions in the tropical Pacific (Betancourt et al. 2000; Koutavas et al. 2002). Conclusions The late glacial to Holocene transition from the Lago Chungara´ record is made up of laminated diatom-rich sediments which provide excellent material for the application of oxygen isotope analysis in biogenic silica. d18Odiatom data have for the first time provided palaeoclimatic reconstruction at decadal-to-centen- nial resolution. The well-laminated nature of these sediments allowed a lamina by lamina continuous sampling, giving one of the highest resolution records available for d18Odiatom. It has also revealed impor- tant insights into the usefulness of this method, as well as provided decisive palaeoenvironmental infor- mation for this critical period. d18Odiatom from dark-green diatom laminae repre- sent the baseline in the environmental evolution of Lago Chungara´, and show decadal to centennial variability in the moisture conditions of the Andean Altiplano. The isotopic record displays a persistent background isotope enrichment trend related to changes in the lake morphology and groundwater outflow during the late glacial and early Holocene. Overprinted onto this long-term (centennial to mil- lennial) trend there are cyclically short-term (decadal to centennial) shifts which are not related to changes in temperature or isotopic composition of the source of precipitation, but to the P-E balance variability in the Altiplano. The record shows two major isotope depletions, occurring at a centennial time scale (11,800 and 11,550 cal year BP) indicating a long-term increase in moisture conditions, and one major isotope enrichment above the background levels that occurred between 11,990 and 11,800 cal year BP indicating a short dry phase during the late glacial. Minor depletions at a decadal time scale are associ- ated with weaker rainfall short-term events. The comparison with terrigenous input and effective moisture availability reconstructions previously per- formed for Lago Chungara´ shows agreement, but includes a systematic time lag (up to 50 years) among these proxies and d18Odiatom. This is mainly due to the time necessary to change the d18Olakewater values and its subsequent incorporation into the diatom frustules, but other factors should not be completely disre- garded. The time lag highlights the fact that not all the proxies react at the same time to environmental forcing and this needs to be more often recognised in high resolution palaeolimnological reconstructions. Sub-millennial shifts in the hydrological balance of Lago Chungara´ are hypothesised to be the result of changes in the strength and position of the Bolivian High. Spectral analyses of d18Odiatom suggest that these changes in the atmospheric conditions over the Altiplano during the wet events were triggered by both ENSO and solar activity. The change from the late glacial dry period to a wetter early Holocene period confirms a weakening of El Nin˜o intensity in J Paleolimnol 123 the Andean Altiplano region in favour of La Nin˜a-like conditions found elsewhere. Nested upon the under- lying climate dynamics are the different cyclicities of solar activity (Schwabe, Hale and Bru¨ckner) that were active during different time windows. There is undoubtedly an interaction between these and ENSO at the decadal and greater scales and it is likely that apparent solar forcing of the Lago Chungara´ record is transmitted via ENSO modulation of the South American monsoon. The complexity of Andean Altiplano palaeoenvironmental conditions, and the absence of other high resolution studies for this time interval, does not allow us to establish any clear conclusion on the existence of significant climatic events synchronous to the Younger Dryas in the northern hemisphere. While many studies have dem- onstrated ENSO-like forcing during the glacial-inter- glacial transition, this highly resolved record is one of the few that preserves key ENSO frequencies, there- fore further implicating this major climatic process with events governing the transition to the Holocene. Acknowledgments The Spanish Ministry of Science and Inno- vation funded the research at Lago Chungara´ through the projects ANDESTER (BTE2001-3225), Complementary Action (BTE2001- 5257-E), LAVOLTER (CGL2004-00683/BTE), GEOBILA (CGL2007-60932/BTE) and CONSOLIDER-Ingenio 2010 GRACCIE (CSD2007-00067). A. Herna´ndez have benefited from a FPI grant from The Spanish Ministry of Science and Innovation. The Limological Research Center (USA) provided the technology and expertise to retrieve the cores. We are grateful to CONAF (Chile) for the facilities provided in Chungara´. The NIGL (UK) funded the isotope analysis, and Hilary Sloane is specially thanked for assistance with the diatom oxygen isotope measurements. References Abbott MB, Wolfe BB, Wolfe AP, Seltzer GO, Aravena R, Mark BG, Polissar PJ, Rodwell DT, Rowe HD, Vuille M (2003) Holocene paleohydrology and glacial history of the central Andes using multiproxy lake sediment studies. Palaeogeogr Palaeoclimatol Palaeoecol 194:123–138. doi: 10.1016/s0031-0182(03)00274-8 Aravena R, Suzuki O, Pen˜a H, Pollastri A, Fuenzalida H, Grilli A (1999) Isotopic composition and origin of the precipi- tation in northern Chile. Appl Geochem 14:411–422 Baker PA, Seltzer GO, Fritz SC, Dunbar RB, Grove MJ, Tapia PM, Cross SL, Rowe HD, Broda JP (2001a) The history of South American tropical precipitation for the past 25, 000 years. Science 291:640–643 Baker PA, Rigsby CA, Seltzer GO, Fritz SC, Lowenstein TK, Bacher NP, Veliz C (2001b) Tropical climate changes at millennial and orbital timescales on the Bolivian Alti- plano. Nature 409:698–701 Barker PA, Street-Perrott FA, Leng MJ, Greenwood PB, Swain DL, Perrott RA, Telford RJ, Ficken KJ (2001) A 14 ka oxygen isotope record from diatom silica in two alpine tarns on Mt Kenya. Science 292:2307–2310 Barker PA, Leng MJ, Gasse F, Huang Y (2007) Century-to- millennial scale climatic variability in Lake Malawi revealed by isotope records. Earth Planet Sci Lett 261:93– 103. doi:10.1016/j.epsl.2007.06.010 Betancourt JL, Latorre C, Rech JA, Quade J, Rylander KA (2000) A 22,000-year record of monsoonal precipitation from Northern Chile’s Atacama Desert. Science 289: 1542–1546 Brandriss ME, O’Neil JR, Edlund MB, Stoermer EF (1998) Oxygen isotope fractionation between diatomaceous silica and water. Geochim Cosmochimi Acta 62:1119–1125 Brewer TS, Leng MJ, Mackay AW, Lamb AL, Tyler JJ, Marsh NG (2008) Unravelling contamination signals in biogenic silica oxygen isotope composition: the role of major and trace element geochemistry. J Quat Sci 23:321–330. doi: 10.1002/jqs.1171 Clayton RN, Mayeda TK (1963) The use of bromine penta- fluoride in the extraction of oxygen from oxide and sili- cates for isotope analysis. Geochim Cosmochim Acta 27:43–52 Dansgaard W (1964) Stable isotopes in precipitation. Tellus 16:436–468 Dorador C, Pardo R, Vila I (2003) Variaciones temporales de para´metros fı´sicos, quı´micos y biolo´gicos de un lago de altura: el caso del Lago Chungara´. Rev Chil Hist Nat 76:15–22 Fritz SC (2008) Deciphering climatic history from lake sedi- ments. J Paleolimnol 39:5–16. doi:10.1007/s10933-007- 9134-x Garreaud RD, Vuille M, Clement AC (2003) The climate of the Altiplano: observed current conditions and mechanisms of past changes. Palaeogeogr Palaeoclimatol Palaeoecol 194:5–22. doi:10.1016/S0031-0182(03)00269-4 Gasse F, Fontes JC (1992) Climatic changes in northwest Africa during the last deglaciation (16–7 ka BP). NATO ASI Series 12. Kluwer, Dordrecht, pp 295–325 Gat JR (1980) Isotope hydrology of very saline lakes. In: Nissenbaum A (ed) Hypersaline brines and evaporitic environments. Elsevier, Amsterdam, pp 1–8 Giralt S, Moreno A, Bao R, Sa´ez A, Prego R, Valero BL, Pueyo JJ, Gonza´lez-Sampe´riz P, Taberner C (2008) Sta- tistical approach to distangle environmental forcings in a lacustrine record: the Lago Chungara´ case (Chilean Alti- plano). J Palaeolimnol 40:195–215. doi:10.1007/s10933- 007-9151-9 Grosjean M, Valero-Garce´s B, Geyh MA, Messerli B, Schreier H, Kelts K (1997) Mid and late holocene limnogeology of laguna del negro francisco, northern chile, and its paleo- climatic implications. The Holocene 7:151–159 Grosjean M, van Leeuwen JFN, van der Knaap WO, Geyh MA, Ammann B, Tanner W, Messerli B, Nu´n˜ez L, Valero- Garce´s BL, Veit H (2001) A 22,000 14C year BP sedi- ment and pollen record of climate change from Laguna Miscanti (23S), Northern Chile. Glob Planet Change 28:35–51 J Paleolimnol 123 Herna´ndez A, Bao R, Giralt S, Leng MJ, Barker PA, Pueyo JJ, Sa´ez A, Moreno A, Valero-Garce´s B, Sloane HJ (2007) A high-resolution study of diatom oxygen isotopes in a Late Pleistocene to Early Holocene laminated record from Lake Chungara´ (Andean Altiplano, Northern Chile). Geochim Cosmochim Acta 71:A398 Herna´ndez A, Bao R, Giralt S, Leng MJ, Barker PA, Sa´ez A, Pueyo JJ, Moreno A, Valero-Garce´s BL, Sloane HJ (2008) The palaeohydrological evolution of Lago Chungara´ (Andean Altiplano, northern Chile) during the Lateglacial and early Holocene using oxygen isotopes in diatom silica. J Quat Sci 23:351–363. doi:10.1002/ jqs.1173 Herrera C, Pueyo JJ, Sa´ez A, Valero-Garce´s BL (2006) Rela- cio´n de aguas superficiales y subterra´neas en el a´rea del lago Chungara´ y lagunas de Cotacotani, norte de Chile: un estudio isoto´pico. Rev Geol Chile 33:299–325 Hora J, Singer B, Wo¨rner G (2007) Volcano eruption and evaporative flux on the thick curst of the Andean Central Volcanic Zone: 40Ar/39Ar constrains from Volca´n Pari- nacota, Chile. Geol Surv Am Bull 119:343–362. doi: 10.1130/B25954.1 Juillet-Leclerc A (1986) Cleaning process for diatomaceous samples. In: Ricard M (ed) 8th diatom symposium. Koeltz Scientific Books, Koenigstein, pp 733–736 Koutavas A, Lynch-Stieglitz J, Marchitto T, Sachs J (2002) El Nin˜o-like pattern in ice age tropical Pacific sea surface temperature. Science 297:226–230 Leng MJ, Barker PA (2006) A review of the oxygen isotope composition of lacustrine diatom silica for palaeoclimate reconstruction. Earth Sci Rev 75:5–27. doi:10.1016/ j.earscirev.2005.10.001 Leng MJ, Lamb AL, Heaton THE, Marshall JD, Wolfe BB, Jones MD, Holmes JA, Arrowsmith C (2005) Isotopes in lake sediments. In: Leng MJ (ed) Isotopes in palaeoen- vironmental research. Springer, Dordrecht, pp 147–184 Maslin MA, Burns SJ (2000) Reconstruction of the Amazon Basin effective moisture availability over the past 14,000 years. Science 290:2285–2287 Moreno A, Giralt S, Valero-Garce´s BL, Sa´ez A, Bao R, Prego R, Pueyo JJ, Gonza´lez-Sampe´riz P, Taberner C (2007) A 13 kyr high-resolution record from the tropical Andes: The Chungara´ Lake sequence (18S, northern Chilean Altiplano). Quat Int 161:4–21. doi:10.1016/j.quaint.2006. 10.020 Morley DW, Leng MJ, Mackay AW, Sloane HJ, Rioual P, Battarbee RW (2004) Cleaning of lake sediment samples for diatom oxygen isotope analysis. J Paleolimnol 31:391–401 Moschen R, Lu¨cke A, Schleser G (2005) Sensitivity of bio- genic silica oxygen isotopes to changes in surface water temperature and palaeoclimatology. Geophys Res Lett 32:L07708. doi:10.1029/2004GL022167 Moy CM, Seltzer GO, Rodbell DT, Anderson DM (2002) Variability of El Nin˜o/Southern Oscillation activity at millenial timescales during the Holocene epoch. Nature 420:162–165 Mu¨hlhauser H, Hrepic N, Mladinic P, Montecino V, Cabrera S (1995) Water-quality and limnological features of a high- altitude andean lake, Chungara´ in northern chile. Rev Chil Hist Nat 68:341–349 Negri AJ, Adler RF, Shepherd JM, Huffman G, Manyin M, Neklin EJ (2004) A 16-year climatology of global rainfall from SSM/I highlighting morning versus evening differ- ences. 13th Conference on Satellite Meteorology and Oceanography. American Meteorological Society, Nor- folk, VA P6. 16 Placzek C, Quade J, Patchett PJ (2006) Geochronology and stratigraphy of late Pleistocene lake cycles on the southern Bolivian Altiplano: Implications for causes of tropical climate change. Geol Soc Am Bull 118:515–532. doi: 10.1130/B25770.1 Rietti-Shati M, Shemesh A, Karlen W (1998) A 3000-year climatic record from biogenic silica oxygen isotopes in an equatorial high-altitude lake. Science 281:980–982 Rigsby CA, Bradbury JP, Baker PA, Rollins SM, Warren MR (2005) Late Quaternary palaeolakes, rivers, and wetlands on the Bolivian Altiplano and their palaeoclimatic implications. J Quat Sci 20:671–691. doi:10.1002/jqs.986 Rings A, Lucke A, Schleser GH (2004) A new method for the quantitative separation of diatom frustules from lake sediments. Limnol Oceanogr Methods 2:25–34 Rodbell DT, Seltzer GO, Anderson DM, Abbott MB, Enfield DB, Newman JH (1999) An 15, 000-year record of El Nin˜o-driven alluviation in southwestern Ecuador. Science 283:516–520 Rodo´ X, Rodrı´guez-Arias MA (2004) El Nin˜o–Southern oscillation: absent in the early holocene? J Clim 17:423– 426 Rowe HD, Dunbar RB, Mucciarone DA, Seltzer GO, Baker PA, Fritz S (2002) Insolation, moisture balance and cli- mate change on the South American Altiplano since the Last Glacial Maximum. Clim Change 52:175–199 Sa´ez A, Valero-Garce´s BL, Moreno A, Bao R, Pueyo JJ, Gonza´lez-Sampe´riz P, Giralt S, Taberner C, Herrera C, Gibert RO (2007) Volcanic controls on lacustrine sedi- mentation: The late Quaternary depositional evolution of lake Chungara´ (Northern Chile). Sedimentology 54:1191– 1222. doi:10.1111/j.1365-3091.2007.00878.x Servant M, Servant-Vildary S (2003) Holocene precipitation and atmospheric changes inferred from river paleowet- lands in the Bolivian Andes. Palaeogeogr Palaeoclimatol Palaeoecol 194:187–206 Stockwell RG, Mansinha L, Lowe RP (1996) Localization of the complex spectrum: the S transform. IEEE TSP 44:998–1001 Sylvestre F, Servant M, Servant-Vildary S, Causse C, Fournier M, Ybert J-P (1999) Lake-level chronology on the southern Bolivian Altiplano (188–238S) during late-gla- cial time and the early Holocene. Quat Res 51:54–66 R Development Core Team (2008) R: a language and envi- ronment for statistical computing. R Foundation for sta- tistical computing, Vienna, Austria ISBN 3-900051-07-0. http://www.R-project.org Theissen KM, Dunbar RB, Rowe HD, Mucciarone DA (2008) Multidecadal- to century-scale arid episodes on the Northern Altiplano during the middle Holocene. Palaeo- geogr Palaeoclimatol Palaeoecol 257:361–376. doi: 10.1016/j.palaeo.2007.09.011 Thompson LG, Davis ME, Mosley-Thompson E, Sowers TA, Henderson KA, Zagorodnov VS, Lin PN, Mikhalenko VN, Campen RK, Bolzan JF, Cole-Dai J, Francou B J Paleolimnol 123 (1998) A 25, 000-year tropical climate history from Bolivian ice cores. Science 282:1858–1864 Valero-Garce´s BL, Grosjean M, Schwalb A, Schreir H, Kelts K, Messerli B (2000) Late Quaternary lacustrine deposi- tion in the Chilean Altiplano (18–28S). In: Gierlowski- Kordech E, Kelts K (eds) Lake basins through space and time. Studies in Geology 46. Am Assoc Petr Geol, pp 625–636 Valero-Garce´s BL, Delgado-Huertas A, Navas A, Edwards L, Schwalb A, Ratto N (2003) Patterns of regional hydro- logical variability in central-southern Altiplano (188- 268S) lakes during the last 500 years. Palaeogeogr Pal- aeoclimatol Palaeoecol 194:319–338 Velasco VM, Mendoza B (2008) Assessing the relationship between solar activity and some large scale climatic phenomena. Adv Space Res 42:866–878. doi:10.1016/ j.asr.2007.05.050 Vuille M, Werner M (2005) Stable isotopes in precipitation recording South American summer monsoon and ENSO variability: observations and model results. Clim Dyn 25:401–413. doi:10.1007/s00382-005-0049-9 Weng C, Bush MB, Curtis JH, Kolata AL, Dillehay TD, Bin- ford MW (2006) Deglaciation and Holocene climate change in the western Peruvian Andes. Quat Res 66:87– 96. doi:10.1016/j.yqres.2006.01.004 J Paleolimnol 123 Biogeochemical processes controlling oxygen and carbon iso- topes of diatom silica in lacustrine rhythmites Armand Hernández1,2*, Roberto Bao3, Santiago Giralt1, Philip A. Barker4, Melanie J. Leng5, Hilary J. Sloane5, Alberto Sáez2 1Institute of Earth Sciences ‘Jaume Almera’-CSIC, C/Lluís Solé i Sabarís s/n, 08028 Barcelona, Spain 2Faculty of Geology, University of Barcelona, C/ Martí Franquès s/n, 08028 Barcelona, Spain 3Faculty of Sciences, University of A Coruña, Campus da Zapateira s/n, 15701 A Coruña, Spain 4Lancaster Environment Centre, Lancaster University, Lancaster LA1 4YQ, UK 5NERC Isotope Geosciences Laboratory, British Geological Survey, Nottingham NG12 5GG, UK Abstract Biogeochemical cycles and sedimentary records in lakes are related to climate controls on hydrology and catchment processes. Changes in the isotopic composition of the diatom frustules (δ18O diatom and δ13C diatom ) in lacustrine sediments can be used to reconstruct palaeoclimatic and palaeoenvironmental changes. The Lago Chungará diatomaceous laminated record is made up of white and green multiannual rhythmites. White laminae were formed during short-term super- blooms, and are composed almost exclusively of large-size Cyclostephanos andinus which bloom during mixing events when recycled nutrients from the bottom waters are brought to the surface and/or when nutrients are introduced from the catchment during periods of strong runoff. Conversely, the green laminae, are thought to have been deposited over several years and are composed of a mixture of diatoms (mainly smaller valves of Cyclostephanos andinus and Discostella stelligera) and organic matter. These green laminae reflect the lake’s hydrological recovery from the conditions favouring the diatom super-blooms (white laminae) towards baseline conditions. Analyses of both δ18O diatom and δ13C diatom in these rhythmites are interpreted in terms of shifts in the precipitation/evaporation ratio and changes in the lake water dissolved carbon concentration, respectively. δ18O diatom composition shows that white laminae formation occurred mainly during low lake level stages, whereas green laminae formation generally occurred during high lake level stages. The isotope and chronostratigraphical data together suggest that white laminae deposition is caused by extraordinary environmental events. El Niño-Southern Oscillation and solar activity are the most likely main climate forcing mechanisms that could trigger such events, favouring hydrological changes at interannual-to-decadal scale. This study demonstrates the potential for laminated lake sediments to document extreme events. Keywords: Oxygen isotopes, carbon isotopes, diatoms, Lago Chungará, ENSO *Corresponding author: Institute of Earth Sciences ‘Jaume Almera’ (CSIC). C/Lluis Solé i Sabarís s/n. E-08028 Barcelona (Spain). Phone: +34.934.095.410 Fax: +34.934.110.012 E-mail: ahernandez@ija.csic.es 1.Introduction Rhythmites are finely laminated sequences (millime- tre- to submillimetre thick) made up of regular alterna- tions of two or three contrasting sediment types called couplets or triplets (Talbot and Allen, 1996). Rhythmite formation is generally associated with seasonally het- erogeneous sediment supply and a lack of physical or biological reworking processes (Grimm et al. 1996). Thus, laminated sediments indicate high-frequency environ- mental change through time. A number of studies have described laminated lacustrine sediments, but they have mainly dealt with annual-rhythmites (varves) with dif- ferent clastic grain-size and/or biogenic content depos- ited over different seasons (e.g. Bird et al. 2009). At mid- to high latitudes the processes that lead to rhythmite for- mation are often well constrained (e.g. Chang et al. 2003), whereas the biogeochemical processes and climate events which prompt laminated sediments in tropical lacustrine sediments are often less understood. In these cases, tropical rainfall regimes associated with intense storms and wind may be responsible for extraordinary external nutrient loading or upwelling of nutrient rich- waters which trigger phytoplankton blooms (Talbot and Allen, 1996). These tropical climate regimes follow a seasonal behaviour (e.g. monsoons), but they can also be highly influenced by climatic multiannual phenom- ena (e.g. ENSO). Changes in the oxygen isotopic composition of the diatom frustules (δ18O diatom ) in lacustrine sediments are used to infer hydrological variations. For closed lakes in the tropics, these variations are mostly related to the pre- cipitation-evaporation ratio (P/E), which is generally directly linked to lake level change (Leng and Barker, 2008). The isotope-inferred reconstructions can thus be used to unreveal the climate history of the region (e.g. Barker et al. 2007) although this may be mitigated by biological and sedimentary processes. Besides d18O diatom , the isotopic signature of carbon occluded within the dia- tom silica (δ13C diatom ), can give other relevant palaeoenvironmental information, including insights on the lakes’ carbon cycle. There are few studies of carbon isotopes from organic inclusions within diatom frustules, and of those published, most have dealt with marine sedi- mentary records (e.g. Crosta and Shemesh, 2002). Stud- ies on δ13C diatom in lake sediments are now emerging and providing valuable insights into the complex carbon cy- cle of lakes (Hurrell et al., submitted). The aim of this paper is to understand high frequency biological, chemical and sedimentary processes which cause the laminae formation in the sedimentary record of Lago Chungará, a high altitude tropical lake located in the Central Andes. δ18O diatom and δ13C diatom data from indi- vidual lamina are presented for a period between 11,990 and 11,530 cal yr BP. High frequency environmental perturbations brought about by interannual-decadal cli- matic events are rarely recorded in lake sediments, and therefore, the laminated sediments here are a good record of their intensity and their effect on lacustrine hydrologi- cal and carbon cycles. 2. Lago Chungará setting 2.1. Climate, geology and limnology A variable precipitation pattern dominates the Lago Chungará region, where the annual rainfall ranges from 100 to 750 mm yr-1 (mean 411 mm yr-1), and more than the 70% of it falls during the austral summer (Decem- ber–February). At this time, a strong low pressure re- gion, known as the South American Summer Monsoon (SASM), is formed over Central South America driving convection and pulling moisture from the equatorial At- lantic to the Andean Altiplano (Zhou and Lau, 1998; Vuille and Werner, 2005). The SASM is a major component of the climate system over tropical and subtropical South America during the austral summer and is remotely forced by tropical Pacific SSTs (Vuille and Werner 2005). At interannual timescales, El Niño-Southern Oscillation (ENSO) is the most important forcing causing climatic fluctuations over the tropical Americas owing it controls changes in the Pacific Tropical Sea Surface Temperatures (SSTs) (Dettinger et al. 2001; Vuille et al. 2003). Moreo- ver, decadal variations in solar activity are currently modulating the sign and strength of the zonal winds above the Altiplano (Theissen et al. 2008). There is an interac- tion between the solar activity and ENSO at the decadal and greater scales and it is likely that solar forcing is A Hernández et al. / Palaeogeography, Palaeoclimatology, Palaeoecology (submitted)176 transmitted to the Lago Chungará record via ENSO modu- lation of the South American monsoon (Hernández et al. in press). Lago Chungará (18º15’S, 69º10’W, 4520 m a.s.l.) is a cold-polymictic and oligo- to meso-eutrophic lake located in the Andean Altiplano (Central Andes). The lake sits on the Cenozoic Lauca Basin sedimentary deposits sur- rounded by volcanoes. The Chungará infill mostly com- prises organic diatomaceous sediments with abundant tephra from the Parinacota Volcano which was active during most of the Late Glacial and Holocene (Sáez et al. 2007) (Fig. 1A). The lake occupies 21.5 km2 and has a maximum water depth of 40 m (Fig. 1B). It is moderately alkaline (pH between 8.99 and 9.30), well mixed (7.6 ppm O 2 at 34 m deep), salinity is around 1.2 g·l-1, conduc- tivity values range between 1500 and 3000 mS cm-1 and waters are of the Na+-Mg2+-HCO 3 --SO 4 2- type (Sáez et al., 2007). The phytoplankton community is made up of a few major species; diatoms dominant the cold season, whereas Chlorophyceae dominate during the austral sum- mer (Dorador et al. 2003). Macrophyte communities form SUBUNIT 1a. Green laminated diatomaceous ooze SUBUNIT 1b. Brown laminated diatomaceous ooze SUBUNIT 2b. Mafic minerals-rich diatomaceous ooze and tephras Peat, silt and diatomaceous oozes, rich in charophytes and mollusc Gravel, sand and sitl, including deformed massflows Pre-lacustrine Quaternary fluvial-alluvial deposits Miocene volcanic rocks 40 30 20 10 0 m Central plain E platformRise YX SUBUNIT 2a. Diatomaceous ooze, carbonate and tephra layers 2 km Lago Chungará Parinacota volcano 32m 27m 17m Chungará Parinacota volcano river2 km 7m E platform Rise Central plain 69 10’ W o 18 15’ S o X Y A NW-SE B C studied core Fig. 1. A. Panoramic view of Lago Chungará. B. Bathymetric map of Lago Chungará showing the main morphological units of the lake floor cited in the text, and position of the studied core. Black line indicates the cross section (C) throughout the lake. C. Cross section of sediment infilling of Lago Chungará. Position of the studied core is shown; note that the position of the core is projected in its equivalent position at the lake central plain. Arrows indicate major hydrological inputs and sedimentary contributions to the lake. Simplified from Sáez et al. (2007). dense patches and microbial colonies in the littoral zone contribute to primary productivity. The local vegetation in the catchment is characterised by low cover values (<30%), being dominated by grasses, shrubs, soligenous peatlands, and Quenoa dwarf forests (Baied and Wheeler, 1993; Moreno et al. 2007). The lake is considered hydrologically closed as there is no surface outlet and the residence time of the lake water is approximately 15 years (Herrera et al. 2006). The main inlet to the lake is the Chungará River (300- 460 l s-1), whereas evaporation causes the main water loss (3.107 m3·yr-1), and represents about 80% of the total outflow. The d18O and dD composition of the lake water in 2002 and 2004 (–1.4‰ SMOW and –43.4‰ SMOW, respectively) diverge significantly from the Glo- bal Meteoric Water Line (GMWL), the Regional Meteoric Water Line (RMWL, where d18O presents a mean value of - –14.3‰ and dD shows a mean value of -–95‰) and isotope composition of the inflowing water (–12.6‰ SMOW and –108.5‰ SMOW, respectively) (Herrera et al. 2006). The lake water is enriched compared to the A Hernández et al. / Palaeogeography, Palaeoclimatology, Palaeoecology (submitted) 177 inflowing water (δ18O by +11.2‰ and δD by +65‰) due to evaporation. 2.2. Rhythmite type sedimentary model Stratigraphy and facies association for the upper- most part of the sedimentary sequence infill of Lago Chungará was established by fifteen Kullenberg cores and seismic imagery (Valero-Garcés et al. 2000; Sáez et al. 2007). Laminated sediments present in the lower sedi- mentary unit 1 defined in Sáez et al. (2007) were divided in the subunits 1a and 1b according to its green or brown dominating colour and were correlated over the lake off- shore zone (central basin) (Fig. 1C). A petrographical study established a preliminary depositional rhythmite type for those sediments where rhythmites are composed by variable-thickness couplets of light-white and dark-green laminae (Hernandez et al. 2008). According to the chronological model, based on 17 14C AMS and one 238U/230Th dates (Giralt et al. 2008), each couplet was deposited during time intervals rang- ing from 4 to 24 yr (Hernandez et al. 2008). The oxygen isotope composition of diatoms from selected samples taken from subunits 1a and 1b have been previously published (Hernández et al. 2008, in press). 3. Methods A 43 cm-thick section of the finely laminated green- ish sediments from subunit 1a (831 cm to 788 cm core depth) was selected as this had well resolved laminae and abundant diatom frustules, and sampled for δ18O diatom , δ13C diatom and %C on diatom-bound organic matter (%C diatom ) analyses. These sediments were continuously sampled for thin sections in order to carry out a detailed petrographical study. Thin sections of 120 mm x 35 mm (30 mm in thickness), with an overlap of 1 cm at each end, were obtained after freeze-drying and balsam-hard- ening. Detailed petrographical descriptions and lamina thickness measurements were performed with a Zeiss Axioplan 2 Imaging petrographic microscope. A number of samples were also selected for observation with a Jeol JSM-840 electron microscope in order to comple- ment the petrographical study. Moreover, a grey-colour curve was calculated using the ImageJ software package (Rasband 1997–2009). The results are presented in a 21 running mean smoothed curve. A total of 102 samples were obtained and 100 were successfully analysed for δ18O diatom . Additionally, 11 of these samples, due to the difficulty to get enough amount of sample and obtain reliable results, were also analysed for δ13C diatom and %C diatom . Two previous studies de- scribed δ18O diatom data from 22 (Hernández et al. 2008) and 40 (Hernández et al. In press) dark-green sample levels to establish the baseline environmental evolution of Lago Chungará. δ18O diatom , δ13C diatom and %C diatom samples were treated following the method proposed by Morley et al. (2004) with some modifications (Hernández et al., 2008; Hurrell et al; submitted). For δ18O diatom analyses the classical step- wise fluorination method was applied to strip hydrous components from diatom silica before a full reaction with BrF 5 (Leng and Barker, 2006; Leng and Sloane, 2008). The oxygen liberated was then converted to CO 2 and normalised through the laboratory standard (BFC) and the NBS-28 quartz standard, referenced to VSMOW. A random selection of more than 30 samples were ana- lysed in duplicate or even in triplicate giving a reproduc- ibility between 0.0‰ and 0.3‰ with a mean value of 0.15‰. Three samples with a reproducibility >0.3‰ were rejected. Isotope variations of consecutive sam- ples are between 0‰ and 6.5‰, with a mean value of 1.0‰. Samples with differences <0.15‰ have not been used because they were considered essentially the same. As a result, until 81 inter-samples relationships between samples have been included in the analysis. The δ13C diatom content on diatom-bound organic mat- ter analyses were performed by combustion in an el- emental analyser (Costech ECS4010) interfaced with a VG dual inlet isotope ratio mass spectrometer. The δ13C diatom values were calculated to the VPDB scale using a within-run laboratory standards calibrated against NBS 18 and 19, and additionally cross checked with NBS 22. %C analyses were performed by combustion separately in the elemental analyser calibrated against an Acetani- lide standard. Replicate ä13C diatom and %C analysis of well- mixed samples indicated a precision of + <0.1. All the analyses were carried out at the NERC Isotope Geosciences Laboratory, British Geological Survey (UK). A Hernández et al. / Palaeogeography, Palaeoclimatology, Palaeoecology (submitted)178 4. Results 4.1 Laminae biogenic composition The present study extends the petrographical exami- nation of the diatomaceous laminated sediments of Lago Chungará for the Late Glacial to early Holocene transi- tion (11,990 - 11,530 cal yr BP) described in Hernández et al. (2008). A hundred laminae have been differenti- ated and grouped under the white, light-green and dark- green laminae categories according to their diatom com- position, organic matter content and colour. Addition- ally, nine laminae were undifferentiated due to their mixed features between the three groups (Fig. 2). White laminae are formed almost exclusively by dia- tom frustules of the large (diameter > 50 μm) 31.1 35340 40 80 120 160 36 37 38 4039 41 O (SMOW) 18 diatomGrey scale 11,990 11,530 11,600 11,700 11,800 11,900 790 788 800 810 820 830 D ep th (c m ) A ge (c al . y rs B P ) C yc le s C or e 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 1 2 ? ? ? ? ? ? ? 3 4* 5 6* 7 8 9 10 11 12 13 14 15 16 17 18 19 20* 21 22 23 24 25 26* 27 28* 29 30 31 32 33* 34 35* 36 37 38 39 40 41 42 43 44* 45 46* 47 48 49 A B C D Fig. 2. A. Digital XRF ITRAX core scanner image from the selected and sampled interval indicating the age and its correspondent core depth. B. The 49 defined cycles composed by couplet/triplets from 102 sampled laminae. C. The smoothed grey-colour curve D. δ18O diatom values associated to each lamina. Note the diatom super-blooms are indicated by thicker white laminae and the higher values of the curve.. euplanktonic diatom Cyclostephanos andinus (Fig. 3G). Dark-green laminae, which contain a higher organic matter content, probably derived from diatoms and other algal groups, are made up of a mixture of different dia- tom species. This mixture is mainly composed of smaller (diameter < 50 μm) Cyclostephanos andinus valves, with diatoms of the Discostella stelligera species complex as co-dominant taxa. Subdominant diatom taxa comprise a number of tychoplanktonic (mainly Staurosira construens aff. venter and Fragilaria spp.) and benthic life forms (including Cocconeis placentula, Gomphonema minutum, Nitzschia tropica and Opephora sp. aff. mutabilis) (Fig. 3C). The light-green laminae are made up of components from the white laminae progressively grading upwards to the typical constituents of the dark- green laminae. Diatoms of the light-green laminae are A Hernández et al. / Palaeogeography, Palaeoclimatology, Palaeoecology (submitted) 179 usually embedded in an organic matrix creating a prefer- ential orientation of the valves (Fig. 3B and E). Thus, a lower white lamina, an intermediate light-green lamina and an upper dark-green lamina form a typical sedimen- tary triplet. These light-green laminae may be variable in thickness or even absent. The transition between well- defined laminae within the triplets (from here on called intra-cycle relationships) is gradual, whereas the transi- tion between different triplets is abrupt (from here on called inter-cycle relationships) (Fig. 3 B, D, F and H). 4.2. Laminae isotope composition In spite of the very high sampling resolution (mean=4 yr, sd=1.5, n=101) δ18O diatom values display a large vari- ability, ranging between +40.1‰ and +31.1‰ with a mean value of +37.5‰ for the whole record (sd=1.1, n=97) (Fig. 2). The studied interval shows three δ18O diatom major enrichment trends which coincide with similar trends in the grey-colour curve (Fig. 4). The %C diatom val- ues range from 0.63% in the uppermost sample (rhythmite 48) to 0.32% in the lowermost sample (rhythmite 8) (mean=0.42%, sd=0.10, n=11) whereas δ13C diatom values oscillate between –26.1‰ and –29.5‰ (mean=–28.1‰, sd=0.95, n=11). The white laminae gen- erally display lower %C diatom and δ13C diatom values than the dark laminae from the same rhythmite, in addition there is an increase in C/Si ratios and δ13C diatom values through- out the 5 studied intra-cycle relationships (Table 1). δ18O diatom inter-cycle relationships have been studied in 49 cases. From these, 12 cases could not be taken into account due to the absence of δ18O diatom data or because the difference between the two consecutive isotopic val- ues was below the mean analytical error. Valid δ18O diatom inter-cycle relationships (n=37) are characterised by higher oxygen isotope values. The most common inter- cycle relationship is the dark-green to white laminae (n = 25), and it shows isotope enrichment (i.e. values in- crease) in 60% of the cases. Likewise, the difference between dark-green laminae to undifferentiated laminae shows similar levels of increasing δ18O diatom , whereas relationships between undifferentiated and white lami- nae show both increases and decreases in δ18O diatom (Ta- ble 2A). There are 51 valid (out of 62) relationships between D E F G H C C D E F G H B B A 1 mm2 mm Fig. 3. A. Digital XRF ITRAX core scanner image of laminated sediments of core 11 corresponding to the sampled interval of Subunit 1a. Note that the lamination is composed by millimetre thick white lamina and green lamina forming rhythmites. B. Photomosaic from a thin-section showing an ideal triplet rhythmite sequence made up of (from base to top): (H) Abrupt contact between dark-green and white laminae; (G) A white lamina formed by skeletons of the large diatom Cyclostephanos andinus (> 50 μm); (F) Gradual contact between white and light-green laminae; (E) A light-green lenticular and discontinous lamina which is made up of a mixture of white and dark-green lamina; (D) Gradual contact between light- and dark- green laminae; (C) A dark-green lamina made up of diatoms embedded in an organic matter matrix. C. SEM image of dark-green lamina mainly made up by Cyclostephanos andinus (black arrows) and diatoms of the Discostella stelligera species complex (white arrows). Note the smaller Cyclostephanos andinus size (diameter < 50μm) D. SEM image showing the decreasing upwards size of the diatoms throughout an intra-cycle contact between a light-green lamina and a dark-green lamina. Arrows indicate the different size of the diatoms. E. SEM image of a light-green lamina. The lamina is made up of complete valves and fragments of Cyclostephanos andinus valves, both showing a preferential orientation. F. SEM image showing an intracycle contact between the white and light-green laminae. Note the preferential orientation of de diatoms placed at the top of the image (light-green lamina). G. SEM image of a white lamina. The lamina is exclusively composed by large Cyclostephanos andinus (diameter > 50μm). The excellent preservation of the diatom frustules can be observed in the image (white arrow). There are no signs of dissolution. H. SEM image showing an intercycle contact between a dark- green and a white lamina. The arrows indicate the exact position of the contact which can be perfectly followed. Note the different size of the diatoms. A Hernández et al. / Palaeogeography, Palaeoclimatology, Palaeoecology (submitted)180 laminae that take place within a rhythmite (intra-cycle relationships). These intra-cycle relationships are domi- nated by isotope depletion (values decrease). The most usual case shows changes from white to dark-green lami- nae (n=23) where isotope decreases occur in 67% of the cases (Table 2B). 34 Gr ey sc ale Ag e ( ca l y r B P) 25 5 B 03 1.1 11,53011,600 11,700 11,800 11,90011,990 C o re se ctio n A 41 C  18Odia tom Fig. 4. A. Digital XRF ITRAX core scanner image from the selected interval. B Grey-colour surface plot elaborated from the digital image. Decadal-scale main grey-colour trends to whiter values are indicated by means of red arrows. C. d18O diatom record. Decadal-scale main d18O diatom trends to higher values are indicated by means of blue arrows. Note the good agreement between both proxies. 5. Discussion 5.1. Biological and sedimentary processes forming rhythmites White laminae features (relatively thick, good dia- tom preservation and monospecific diatom composi- tion) suggest accumulation during short-term massive diatom blooms, perhaps of only days to weeks in dura- tion. According to the chronological model rhythmites are not a product of annual variations in sediment sup- ply, but due to some kind of multiannual processes (Hernández et al., 2008). Causes of super-blooms can be different to regular seasonal blooms which occur as part of the normal phytoplankton succession (Reynolds, 2006). We suggest that our diatom super-blooms may have been triggered by abnormally high nutrient con- centrations coupled with hydrological conditions prompt- ing diatom population growth. Low lake level stages and/ or strong wind episodes would favour upwelling of nu- trient-rich hypolimnion waters (Talbot and Allen, 1978). Strong mixing would also select diatoms over other types of phytoplankton due to their relative buoyancy. Alter- natively, the increase in nutrient external loading due to exceptional catchment erosion during wet events could also have had the same effect (Bradbury et al. 2002). ENSO cyclicity signals recorded at this time in the Lago Chungará record (Hernández et al., in press) provide Sample Cycle Colour Depth (cm) Age (cal yr BP)  18 Odiatom (SMOW)  13 Cdiatom (PDB) %Cdiatom 5 48 Dark-green 789.1 11,543 36.27 -28,99 0,63 6 48 White 789.5 11,547 38.01 -28,51 0,47 13 43 Dark-green 793.4 11,588 38.07 -26,05 0,57 14 43 Light-green 794.1 11,595 37.29 -28,30 0,38 15 43 White 794.5 11,599 38.65 -28,46 0,39 29 37 Dark-green 799.3 11,650 37.67 -27,87 0,40 30 37 White 799.6 11,653 38.16 -29,01 0,33 77 13 Dark-green 820.5 11,872 38.44 -27,21 0,44 78 13 White 820.9 11,876 38.91 -29,53 0,32 87 8 Dark-green 824 11,909 38.62 -27,69 0,38 88 8 White 824.3 11,912 37.03 -28,89 0,32 Table 1 List of samples where both δ18O diatom and δ13C diatom analyses were carried out, including main sample features. A Hernández et al. / Palaeogeography, Palaeoclimatology, Palaeoecology (submitted) 181 Intercycle relationship types Enrichments (%) Depletions (%) n Dark-green to white laminae 60 40 25 Dark-green to undifferentiated laminae 67 33 6 Undifferentiated to white laminae 50 50 6 Total 37 Intracycle relationship types Enrichments (%) Depletions (%) n White laminae to light-green laminae 33 67 9 Light-green to light-green laminae 0 100 1 Light-green to dark-green laminae 56 44 9 White laminae to dark-green laminae 35 65 23 White laminae to dark-green laminae (non- consecutive laminae, base to top of the rhythmite) 33 67 9 Total 51 Table 2 A. Intercycle isotope relationships between the defined rhythmites. B. Intracycle isotope relationships between the defined rhythmites. Relationship types are established according to the colour of the laminae that are in contact. support to the existence of those two contrasting dry (El Niño) and wet (La Niña) conditions (Vuille et al., 2000; Valero-Garcés et al., 2003). Dark-green laminae (made up of a mixture of diatom valves belonging to several planktonic and benthic taxa, all embedded in an organic matter matrix) represent the baseline lake conditions where the complete phytoplankton successions over several years are pre- served. These laminae therefore record the ‘normal’ in- tra- and inter-annual changes in the water column mix- ing regime characterised by the shifting species compo- sition throughout general annual phythoplankton cycles. Preservation occurs as skeletons belonging to several diatom taxa, or simply as the organic matter mainly be- longing to other algal groups (likely Chlorophycean, Cyanobacteria, etc.) that embed the valves in the dark- green laminae. Regular seasonal diatom blooms, are likely manifested in the dark-green laminae by the abun- dance of the small Cyclostephanos andinus (< 50 μm), a large centric diatom whose buoyancy depends on the existence of a turbulent regime. Seasonal Cyclostephanos andinus (< 50 μm) blooms reflected in the dark-green laminae would therefore be triggered by the same proc- ess during the super-blooms of the larger Cyclostephanos andinus (> 50 μm) that make up the white laminae (i.e. water stratification breakdown). The dark-green lami- nae are sometimes preceded by light-green laminae. This observation indicates that recovery of the baseline con- ditions from the super-blooms can be more or less gradual (forming couplets or triplets, respectively). Flocculation of diatoms by extracellular polymeric substances is a common feature in the marine realm (Thornton, 2002). This phenomenon occurs towards the end of a diatom bloom, due to the onset of nutrient limi- tation. Diatom aggregation and subsequent rapid sedi- mentation of species having any kind of resting cell stages would favour future recruitment once nutrient resources were again available (Smetacek, 1985). Biosiliceous laminae in marine sediments have been interpreted as the product of changes in the mass sedi- mentation of diatoms by means of the formation of ag- gregates (Grimm et al., 1996, 1997). At Lago Chungará a similar phenomenon could have taken place in the for- mation of the light-green laminae once the super-blooms of the large (> 50 μm) Cyclostephanos andinus come to an end. Aggregation of cells enclosed in a gelatinous A Hernández et al. / Palaeogeography, Palaeoclimatology, Palaeoecology (submitted)182 matrix could therefore have taken place, being rapidly deposited in the form of the transitional light-green lami- nae. Although the life cycle details of Cyclostephanos are far from fully known, the closely related genera Stephanodiscus, to which Cyclostephanos once belonged (Round et al., 1990), is known to produce resting cells (Sicko-Goad et al., 1989), whose aggregation and rapid sedimentation represents a transition to a resting phase (Smetacek, 1985; Alldredge et al., 1995). It is therefore likely that the mechanism of formation of triplets is mediated by processes of self-sedimentation triggered by Cyclostephanos andinus (Grimm et al., 1997). 5.2. δ18O diatom and δ13C diatom interpretation Variation in δ18O diatom can result from a variety of proc- esses, such as oxygen isotope composition of the lake water (δ18O lakewater ), temperature, vital effects and post depositional diagenesis (Leng and Barker, 2006). In hydrologically closed lakes under arid climate conditions evaporative concentration processes have a much larger effect on δ18O lakewater than any other process (Gasse and Fontes, 1992; Leng and Marshall, 2004; Hernández et al, in press). In these circumstances, the δ18O diatom record can be used as an indicator of changes in the P/E related to climatic change (Leng and Barker, 2006). At present, Lago Chungará can be considered a closed lake due to its water residence time (ca. 15 years), and the fact that δ18O lakewater is enriched by 14‰ relative to δ18O of the inputs (precipitation, springs and river) (Herrera et al. 2006). This was probably also the case in the Late Glacial-early Holocene described here because d18O diatom values are similar (around +37.5‰) to other diatom-isotope sequences in tropical sites (e.g. Lakes from Mount Kenya (Kenya), Barker et al. 2001; Lake Malawi (Malawi, Mozambique, Tanzania), Barker et al. 2007; Lake Tilo (Ethiopia); Lamb et al. 2005). Thereby the variations in the δ18O diatom from Lago Chungará sediments must be mainly derived from changes in the δ18O lakewater resulting from shifts in the balance between P/E, rather than other factors. The organic matter enclosed within diatom frustules contains polysaccharides, proteins and long-chain polyamines (Kröger and Poulsen, 2008). These sub- stances host carbon which is protected from post- depositional diagenetic alteration (Des Combes et al. 2008). As these carbon compounds will be synthesised from the surrounding waters, isotope analysis of the carbon contained in the diatom frustules can be used as a proxy for reconstructing the lake’s carbon cycle. Pre- viously published studies suggest primary productivity and CO 2(aq) concentration as the main factors which de- termine δ13C diatom in marine environments (Schneider- Mor et al. 2005), although lake δ13C diatom is likely control- led by more complex environmental conditions (Hurrell et al. submitted). δ13C diatom variations due to the species effect, cell size, growth rate or/and metabolic pathway are neglected here since in our case the δ13C diatom analysis was always carried out on similar sized-cells (38-62 μm) and on the same diatom species (Cyclostephanos andinus). In lakes, it is usually assumed that the carbon pool in the water becomes enriched in 13C during the periods of enhanced productivity (Leng et al. 2005; Singer and Shemesh, 1995) since phytoplankton preferentially use the lighter isotope. However, the maximum productiv- ity events found here, associated with the white laminae (short-term diatom super-blooms), show the lowest δ13C diatom values. Therefore, although the diatom blooms will have preferentially incorporated 12C, this cannot have been sufficient to positively shift the isotope value of the dissolved carbon. Instead, the supply of carbon avail- able to the diatoms must have been sufficient not to lead to limiting conditions. The carbon isotope values from bulk sediment (δ13C bulk ) in the Lago Chungará laminated unit range from –21‰ to –19‰ (J.J. Pueyo, unpublished data), yielding a difference of more than 5‰ when compared to the measured δ13C diatom values. Nevertheless, the C/N ratio from bulk sediments of the laminated unit have values ranging between 7 and 11 (J,J. Pueyo, unpublished data), indicating that the δ13C bulk signal would have a mainly algal origin (Meyers and Terranes, 2001). For this rea- son, it seems that the δ13C diatom , rather than being mainly affected by changes in the source of organic matter, is mostly conditioned by changes in dissolved carbon con- centration. Mineralisation of terrestrial or previously deposited carbon through microbial decomposition could create a pool of isotopically lighter carbon avail- able to the diatoms. Release of this through respiration A Hernández et al. / Palaeogeography, Palaeoclimatology, Palaeoecology (submitted) 183 (CO 2(aq) and possible also CH 4 ) would be partly control- led by lake dynamics under the control of external forc- ing factors. Lake water dynamics are mainly governed by two contrasting situations (Hernández et al. 2008): (1) a water column subjected to episodes of very strong mixing, which is represented by the white laminae, and (2) a more stable condition, including periods of lake water stratification, and represented by the dark-green laminae. During the stratified periods, concentrations of oxygen and other electron acceptors typically decrease in the hypolimnion, while CO 2(aq) , CH 4 , and nutrients ac- cumulate (Bedard and Knowles 1991). These dissolved nutrients, as well as the accumulated CO 2(aq) and CH 4 , are released into the entire lake during mixis (Houser et al., 2003), when the diatom blooms occur, being the carbon incorporated into the diatom frustules. 5.3. δ18O diatom inter-cycle relationships (white laminae for- mation) The characterisation of δ18O diatom values through the inter-cycle relationships gives clues to understand the underlying processes involved in the formation of the white laminae. The massive diatom blooms that pro- duce the white laminae have to be triggered by an excep- tional injection of nutrients into the water column which may or may not be associated with a water mass change. The start of the rhythmite is usually accompanied by δ18O diatom enrichment (Table 2A), indicating a decrease in the P/E ratio, a drop in the lake water level and a remobilization of nutrients from the hypolimnion as the more likely scenario during the white laminae forma- tion (Fig. 5A and B, transition 1). Episodes of diatom super-blooms occur throughout the whole studied section, but their formation is a time scale-dependent process. At decadal-centennial scales white laminae are more marked (higher values in the grey colour curve) and thicker (around 6 mm) with higher isotope oxygen values (up to 39.2‰) than during other laminae deposition periods (Hernandez et al. in press). Deposition of these white laminae stretches are related to low-stand conditions, as shown in the uppermost part of the three shallowing upwards trends observed in the d18O diatom record (Fig. 4). However, at interannual scales the inter-cycle isotope relationships reveal that both changes to drier or wetter conditions may trigger the formation of the white laminae, but falls in lake level were more likely responsible for the development of the massive diatom blooms (Table 2A). 5.4. δ18O diatom and δ13C diatom intra-cycle relationships (green laminae formation) The intra-cycle relationship between d18O diatom and δ13C diatom provides a means of better understanding of the environmental processes involved in the origin of the green laminae. The δ18O diatom intra-cycle relationships show that transitions from white to dark-green laminae are mainly governed by δ18O diatom depletions by up to - 2.7‰ (65%; n = 23), although there are also a signifi- cant percentage of enrichments (Table 2B). As in the case of the white laminae formation, green laminae can be formed under both lake-water level drops and rises, but their formation is clearly favoured by increasing P/E ratios with subsequent lake-water level rises (Fig. 5A and C, transitions 2 and 3). Green laminae record the baseline conditions in the water column mixing regime including water table strati- fication periods. The intra-cycle relationships which show δ18O diatom depletions indicate that the lake tended to progressively recover the previous environmental con- ditions by means of a gradual increase in water avail- ability (Fig. 5A and C, transition 2 and 3). Conversely, the intra-cycle relationships which show δ18O diatom enrichments would indicate the recovery to a lower lake level after a super-bloom caused by a massive allochthonous nutrient input associated to enhanced rain- fall. This is suggested by the prevalence of δ18O diatom de- pletions that precede those super-blooms (90%; n= 10) (Table Additional Material). This model suggests that the green laminae occurred most of the time as a result of the recovery phase favoured by lake water rises. Finally, when the lake is already in the recovery phase (transi- tional and baseline conditions) it may evolve, indis- tinctly, towards rise or fall lake water level stands, as indicated by the light- to dark-green isotope transitions (enrichments= 56%; n= 9) (Fig. 5A and C, transition 4). Comparison between δ13C diatom and δ18O diatom in the intra-cycle relationships show that the former can be associated to either δ18O diatom enrichments or depletions. A Hernández et al. / Palaeogeography, Palaeoclimatology, Palaeoecology (submitted)184 However, δ13C diatom values show that all intra-cycle rela- tionships yield carbon isotope enrichment during the formation of the organic-rich green laminae (Table 1). This occurs because during the white laminae forma- tion, the strong mixing necessary for the formation of the massive diatom blooms break the lake water strati- fication. Transport of CO 2(aq) and CH 4 from an hypolimnion enriched in these compounds is then al- lowed, depleting the δ13C of the total carbon pool. 5.5. Climatic forcing of the laminae formation The biogeochemical reconstruction presented above suggests that laminae are formed by the occurrence of diatom super-blooms. These are directly affected by nu- trient availability which, in turn, is mainly controlled by 30 diatom size nutrients 50 90 m - + Cyclostephanos andinus Cyclotella stelligera complex benthic or tychoplanktonic diatom Sedimentary facies (W=white laminae, L=light-green laminae, D=dark-green laminae) Rhythmite and sedimentary facies transitions organic-rich mud Anoxic water Baseline conditions Mixing D Extraordinary bloom conditions W Nutrient resuspension A B C Transitional conditions Weak Mixing L Anoxic water Baseline conditions Mixing D Extraordinary bloom conditions Strong Mixing W Water level drop Increasing O 18 diatom Water level rise Decreasing Increasing   18 Odiatom diatom 13 C Water level rise Decreasing Increasing C   18 Odiatom diatom 13 Water level drop Increasing Increasing   18 13 O C diatom diatom D D D L W W 43 2 1 1 4 3 3 2 Nutrient resuspension Strong Mixing Fig. 5. A. Rhythmite log succession showing facies and transitions (indicated by letters and numbers, respectively). B. The most frequent intercycle relationship scenarios. Transition case 1: From dark-green to white laminae, the white laminae formation (diatom super- blooms) is more often favoured by drops of the lake water level (increases in δ18O diatom values) and therefore related to recycled nutrients from the hypolimnion. C. The most common intracycle relationship scenarios. Transition case 2: From white to dark-green laminae, the dark-green lamina formation is usually favoured by rises of the lake level water (decreases in δ18O diatom values). Transition case 3: From white to light-green laminae, the light-green laminae formation is usually favoured by rises of the lake water level (lower δ18O diatom values). Transition case 4: From light-green to dark green laminae, the dark-green laminae formation is almost indistinctly favoured by drops or rises of the lake water level, with a slight predominance of the former as the δ18O diatom show. the lake level fluctuations and mixing as stable isotopes (δ13C diatom and δ18O diatom ) demonstrate. Hence, the laminae formation in Lago Chungará seems to be mainly induced by environmental forcing such as short-term climate variability. ENSO and solar activity, as well as interactions be- tween both phenomena, have been key factors prompt- ing changes in the atmospheric conditions over the Altiplano region during the Late Glacial-early Holocene at decadal and longer term time scales (Hernandez et al. in press). ENSO and solar activity, as responsible for the more accentuated sub-millennial wet or dry conditions over the Andean Altiplano (Theissen et al. 2008), are very likely the main environmental factors in the fre- quency and production of the white laminae. In the Cen- tral Andes, there is a weak trend towards wet conditions A Hernández et al. / Palaeogeography, Palaeoclimatology, Palaeoecology (submitted) 185 during La Niña phase and to dry conditions during El Niño phase (Valero-Garcés et al., 2003). According to our depositional model, white laminae formation could be triggered by both phases. However, El Niño events seem most likely to be responsible, since the white lami- nae formation is usually favoured by changes from wet- to-dry conditions, which is seen in the isotope enrichments. It is important to point out that white lami- nae are more intense (higher intensity grey colour val- ues) and better developed (thicker) during periods show- ing higher isotope values which point to their forma- tion during drier conditions (Hernández et al. in press). Thus, the diatom super-blooms likely occurred during El Niño-like periods. The presence of exceptionally in- tense and thick white laminae could, on the other hand, be indicative of the overlapping of both ENSO and solar activity phenomena during such periods. Isotope data show that high intensity of ENSO and solar activities can be recorded beyond the white laminae deposition. The short duration (days) of extreme blooms in relation to lake water residence time gives an isotope signature that will remain for longer periods (years) being re- corded in the dark laminae (Hernández et al. in press). 6. Conclusions Lago Chungará rhythmites record multiannual dia- tom super-blooms lasting from days to weeks (white laminae) and the lake hydrology recovery towards the baseline conditions throughout several years (dark- green laminae). Self-sedimentation phenomena taking place immediately after the diatom super-blooms can- not be discarded as a sign of the end of the super-bloom (light-green laminae). The diatom super-blooms are fa- voured episodes of extreme turbulent conditions affect- ing the whole water column, and/or by strong runoff during wet episodes. In the first case upwelling from nutrient-rich hypolimnion waters allowed an extraordi- nary nutrient availability, whereas in the second case allochthonous nutrient enrichment would be implicated. In Lago Chungará, the δ18O diatom record can be used as an indicator of changes in the precipitation to P/E re- lated to climatic changes, whereas the δ13C diatom variabil- ity would be mainly influenced by changes in CO 2(aq) con- centration. δ18O diatom values show that both white and green laminae formation may occur in either dry or wet conditions, but the diatom super-blooms were more intense (thicker white laminae) during decadal-centen- nial lowstands. δ18O diatom composition shows that the white laminae formation was mainly favoured by low lake levels, whereas the green laminae formation was especially prompted by lake level rises. ENSO and solar activity are the most likely main cli- mate forcing mechanisms triggering the white laminae formation. Both El Niño and La Niña phases could be responsible for this, but geochemical data indicate that dry conditions associated to El Niño could have the pri- mary role since the white laminae formation was usu- ally favoured by changes from wet-to-dry conditions in the Altiplano region. Moreover, the periods where the white laminae present major thickness and whiter col- ours might be indicative of phases with overlapping El Niño and solar activity. On the contrary, green-laminae were deposited during the baseline climate phases, when the normal plankton succession throughout several years and associated regular diatom blooms occur. High resolution isotope analysis of the oxygen and carbon isotopes in diatom silica in this uniquely lami- nated sequence has displayed links between limnology, catchment runoff variations, hydrology and climate forc- ing at different time scales. Strong El Niño phases have triggered nutrient and carbon release from the hypolimnion and sediments that has led to diatom su- per-blooms. Such phenomena may be found in many lakes but few preserve evidence in their sedimentary architecture. Further work on other parts of this record and in similarly laminated sites may reveal the full im- pact of these multi-annual events on lake ecosystems and biogeochemical cycles. Acknowledgments The Spanish Ministry of Science and Innovation funded the research at Lago Chungará through the projects ANDESTER (BTE2001-3225), Complementary Action (BTE2001-5257-E), LAVOLTER (CGL2004-00683/BTE), GEOBILA (CGL2007-60932/BTE) and CONSOLIDER- Ingenio 2010 GRACCIE (CSD2007-00067). A. Hernández have benefited from a FPI grant from The Spanish Min- istry of Science and Innovation. The Limological Re- A Hernández et al. / Palaeogeography, Palaeoclimatology, Palaeoecology (submitted)186 A Hernández et al. / Palaeogeography, Palaeoclimatology, Palaeoecology (submitted) search Center (USA) provided the technology and exper- tise to retrieve the cores. We are grateful to CONAF (Chile) for the facilities provided in Parque Nacional Lauca. The NIGL (UK) funded the isotope analyses, Chris P. Kendrick is thanked for conducting the carbon isotope measure- ments. We also wish to thank Alice Chang and Juan J. Pueyo for valuable discussions on the self-sedimenta- tion process and on the manuscript, respectively. References Alldredge AL, Gotschalk C, Passow U, Riebesell U. 1995. Mass ag- gregation of diatom blooms: Insights from a mesocosm study. Deep Sea Research Part II: Topical Studies in Oceanography 42: 9-27. Baied CA, Wheeler JC. 1993. Evolution of high Andean puna ecosys- tems: environment, climate, and culture change over the last 12 000 years in the central Andes. Mountain Research & Devel- opment 13: 145-156. Barker PA, Leng MJ, Gasse F, Huang Y. 2007. Century-to-millennial scale climatic variability in Lake Malawi revealed by isotope records. Earth and Planetary Science Letters 261: 93–103. doi:10.1016/j.epsl.2007.06.010 Bedard C, Knowles R. 1991. Hypolimnetic O_{2} consumption, denitrification, and methanogenesis in a thermally stratified lake. Canadian Journal of Fisheries and Aquatic Sciences 48: 1048–1054. Bird BW, Abbott MB, Kutchko B, Finney BP. 2009. A 2000-year Varve-Based Climate Record from the Central Brooks Range, Alaska. Journal of Paleolimnology 41: 25-41. Bradbury P, Cumming B, Laird K. 2002. A 1500-year record of cli- matic and environmental change in Elk Lake, Minnesota III: measures of past primary productivity. Journal of Paleolimnology 27: 321–40. Chang AS, Patterson RT, McNeely R. 2003. Seasonal sediment and diatom record from late Holocene laminated sediments, Effingham Inlet, British Columbia, Canada. Palaios 18: 477–494. Crosta X, Shemesh A. 2002. Reconciling down core anticorrelation of diatom carbon and nitrogen isotopic ratios from the South- ern Ocean. Paleoceanography 17: 1010. doi:10 1029/ 2000PA000565. Des Combes HJ, Esper O, De la Rocha CL, Abelmann A, Gersonde R, Yam R, Shemesh A. 2008. Diatom δ13C, δ15N, and C/N since the Last Glacial Maximum in the Southern Ocean: Potential impact of species composition. Paleoceanography 23: PA4209. doi:10.1029/2008PA0001589. Dettinger MD, Battisti DS, Garreaud RD, McCabe GJ, Bitz CM. 2001. Interhemispheric effects of interannual and decadal ENSO-like climate variations on the Americas. In Interhemispheric climate linkages: Present and Past climates in the Americas and their Societal Effects. Markgraf V. (ed.). Academic Press: 1-16. Dorador C, Pardo R, Vila I. 2003. Variaciones temporales de parámetros físicos, químicos y biológicos de un lago de altura: el caso del Lago Chungará. Revista Chilena de Historia Natural 76: 15–22. Gasse F, Fontes JC. 1992. Climatic changes in northwest Africa dur- ing the last deglaciation (16–7 ka BP). NATO ASI Series 12: Kluwer, Dordrecht; 295–325. Giralt S, Burjachs F, Roca JR, Julià R. 1999. Late glacial to early Holocene environmental adjustment in the Mediterranean semi-arid zone of the Salines playa-lake (Alicante, Spain). Jour- nal of Paleolimnology 21: 449–460. Grimm KA, Lange CB, Gill AS. 1996. Biological forcing of hemipelagic sedimentary laminae: evidence from ODP site 893, Santa Barbara Basin, California. Journal of Sedimentary Research 66: 613–624. Grimm KA, Lange CB, Gill AS. 1997. Self-sedimentation of phytoplankton blooms in the geologic record. Sedimentary Geology 110: 151–161. Hernández A, Bao R, Giralt S, Leng MJ, Barker PA, Sáez A, Pueyo JJ, Moreno A, Valero-Garcés BL, Sloane HJ. 2008. The palaeohydrological evolution of Lago Chungará (Andean Altiplano, northern Chile) during the Lateglacial and early Holocene using oxygen isotopes in diatom silica. Journal of Quaternary Science 23: 351–363.doi: 10.1002/jqs.1173 Hernández A, Giralt S, Bao R, Leng MJ, Barker PA. In press. ENSO and solar activity signals from oxygen isotopes in diatom silica during late glacial-Holocene transition in Central Andes (18ºS). Journal of Paleolimnogeology. Herrera C, Pueyo JJ, Sáez A, Valero-Garcés BL. 2006. Relación de aguas superficiales y subterráneas en el área del lago Chungará y lagunas de Cotacotani, norte de Chile: un estudio isotópico. Revista Geológica de Chile 33: 299–32. Houser JN, Bade DL, Cole JJ, Pace ML. 2003. The dual influences of dissolved organic carbon on hypolimnetic metabolism: or- ganic substrate and photosynthetic reduction. Biogeochemis- try 64: 247–269. Hurrell ER. 2009. Climate change and biogeochemical cycles on East African mountains by stable isotopes of diatom frustules. PhD Thesis. Lancaster Environment Center, Lancaster University: Lancaster. Hurrell ER, Barker PA, Leng MJ, Vane CH, Wynn P, Kendrick CP, Verschuren D, Street-Perrott FA. Developing a methodology for carbon isotope analysis of lacustrine diatoms. Journal of Palaeolimnology: Submitted. Jones V, Leng MJ, Solovieva N, Sloane H, Tarasov P. 2004. Holocene climate on the Kola Peninsula; evidence from the oxygen iso- tope record of diatom silica. Quaternary Science Reviews 23: 833–839. DOI:10.1016/j.quascirev.2003.06.014 Kröger N, Poulsen N. 2008. Diatoms: from Cell Wall Biogenesis to Nanotechnology. Annual Review of Genetics 42: 83–107. Leng MJ, Barker PA. 2006. A review of the oxygen isotope compo- sition of lacustrine diatom silica for palaeoclimate reconstruc- tion. Earth-Science Reviews 75: 5–27. doi: 10.1016/ j.earscirev.2005.10.001 Leng MJ, Marshall JD. 2004. Palaeoclimate interpretation of stable isotope data from lake sediment archives. Quaternary Science Reviews 23: 811– 831. doi: 10.1016/j.quascirev.2003.06.012 Leng MJ, Sloane HJ. 2008. Combined oxygen and silicon isotope analysis of biogenic silica. Journal of Quaternary Science 23: 313–319. Leng MJ, Lamb AL, Heaton THE, Marshall JD, Wolfe BB, Jones MD, Holmes JA, Arrowsmith C. 2005a. Isotopes in lake sediments. In Isotopes in Palaeoenvironmental Research, Leng MJ (ed.). Springer: Dordrecht, Netherlands; 147–184. Meyers PA, Teranes JL. 2001. Sediment organic matter. In Tracking Environmental Change Using Lake Sediments, Physical and 187 Geochemical Techniques, vol. 2. Last WM, Smol JP. (Eds.). Kluwer Academic Publishers: Dordrecht, The Netherlands; 239–270. Moreno A, Giralt S, Valero-Garcés BL, Sáez A, Bao R, Prego R, Pueyo JJ, González-Sampériz P, Taberner C. 2007. A 13 kyr high- resolution record from the tropical Andes: the Chungará Lake sequence (18º S, northern Chilean Altiplano). Quaternary In- ternational 161: 4–21. doi: 10.1016/j.quaint.2006.10.020 Morley DW, Leng MJ, Mackay AW, Sloane HJ, Rioual P, Battarbee RW. 2004. Cleaning of lake sediment samples for diatom oxygen iso- tope analysis. Journal of Paleolimnology 31: 391–401. Rasband WS. 1997-2009. ImageJ. U. S. National Institutes of Health: Bethesda, Maryland, USA; http://rsb.info.nih.gov/ij/ Reynolds CS. 2006. The Ecology of Phytoplankton. Cambridge Uni- versity Press: Cambridge, UK. Rosqvist GC, Jonsson C, Yam R, Karlen W, Shemesh A. 2004. Diatom oxygen isotopes in pro-glacial lake sediments from northern Sweden: a 5000 year record of atmospheric circulation. Qua- ternary Science Reviews 23: 851– 859. Round FE, Crawford RM, Mann DG. 1990. The Diatoms, Biology & Morphology of the Genera. Cambridge University Press: Cam- bridge; 747. Rundel PW, Palma B. 2000. Preserving the unique puna ecosystems of the Andean Altiplano: A descriptive account of Lauca Na- tional Park, Chile. Mountain Research and Development 20: 262-271. Sáez A, Valero-Garcés BL, Moreno A, Bao R, Pueyo JJ, González- Sampériz P, Giralt S, Taberner C, Herrera C, Gibert RO. 2007. Volcanic controls on lacustrine sedimentation: the late Quater- nary depositional evolution of lake Chungará (northern Chile). Sedimentology 54: 1191–1222. doi: 10.1111/j.1365- 3091.2007.00878.x Schneider-Mor A, Yam R, Bianchi C, Kunz-Pirrung M, Gersonde R, Shemesh A. 2005. Diatom stable isotopes, sea ice presence and sea surface temperature records of the past 640 ka in the At- lantic sector of the Southern Ocean. Geophysical Research Let- ters 32: L10704. Sicko-Goad L, Stoermer EF, Kociolek JP. 1989. Diatom resting cell rejuvenation and formation: time course, species records and distribution. Journal of Plankton Research 11: 375–389. Singer AJ, Shemesh A. 1995. Climatically linked carbon-isotope variation during the past 430,000 years in Southern-Ocean sediments. Paleoceanography 10: 171–177. Smetacek VS. 1985. Role of sinking in diatom life-history cycles: ecological, evolutionary and geological significance. Marine Biology 84: 239–251. Talbot MR, Allen PA. 1996. Lakes. In Sedimentary Environments, Reading HG (Ed.). Blackwell: Oxford; 83–124. Theissen KM, Dunbar RB, Rowe HD, Mucciarone DA. 2008. Multidecadal- to century-scale arid episodes on the Northern Altiplano during the middle Holocene. Palaeogeography, Palaeoclimatology, Palaeoecology 257: 361–376. Thornton DCO. 2002. Diatom aggregation in the sea: mechanisms and ecological implications. European Journal of Phycology 37: 149–161. Valero-Garcés BL, Delgado-Huertas A, Navas A, Edwards L, Schwalb A, Ratto N. 2003. Patterns of regional hydrological variability in central-southern Altiplano (18º-26ºS) lakes during the last 500 years. Palaeogeography, Palaeoclimatology, Palaeoecology 194: 319– 338. Valero-Garces BL, Grosjean M, Schwalb A, Schreir H, Kelts K, Messerli B. 2000. Late Quaternary lacustrine deposition in the Chilean Altiplano (18º–28ºS). In Lake Basins through Space and Time, Gierlowski-Kordesch E, Kelts K (eds). American Association of Petroleum Geologists Studies in Geology 46: 625–636. Vuille M, Werner M. 2005. Stable isotopes in precipitation record- ing South American summer monsoon and ENSO variability: Observations and model results. Climate Dynamics 25: 401– 413. Vuille M, Bradley RS, Werner M, Keimig F. 2003. 20th century climate change in the tropical Andes: observations nd model results. Cli- matic Change 59: 75-99. Zhou J, Lau KM. 1998. Does a Monsoon Climate Exist over South America?. Journal of Climate 11: 1020–1040. A Hernández et al. / Palaeogeography, Palaeoclimatology, Palaeoecology (submitted)188