ABSTRACT Title of thesis: NO CHANGE IN ENSO HYDROCLIMATE VARIABILITY AFTER THE INDUSTRIAL REVOLUTION AS RECORDED IN ๐›ฟ18O OF TECTONA GRANDIS L.F. FROM SOUTHEAST SULAWESI, INDONESIA Sandy Hardian Susanto Herho, Master of Science, 2023 Thesis directed by: Dr. Michael N. Evans Department of Geology, University of Maryland College Park El Ninฬƒo-Southern Oscillation (ENSO) is a quasi-periodic interannual oscillation of the ocean-atmosphere system in the tropical Pacific which greatly influences global climate variability. However, the long-term response to greenhouse gas forcing is still controversial. In this study, we measured the oxygen isotopic composition of ๐›ผ-cellulose samples at intra- annual resolution from independently crossdated teak cores (Tectona grandis L. f.) collected at Muna, Indonesia (5.3ยบS, 123ยบE, elev. 10m). The site and observation has been previously shown to provide an indirect measure of ENSO activity via local precipitation amount variations associated with ENSO. We created an ensembled composite of the interannual variability for the period 1680-2005 (316 years) using empirical high pass filtering and random sampling of intra-annual resolution measurements. In processing this time series composite, we used Singular Spectrum Analysis (SSA) to high pass filter the data for the interannual variability associated with ENSO. The annually-resolved composite time series of ๐›ฟ18O that we constructed has a higher resolution than other studies that have been conducted to reconstruct ENSO-hydroclimate activities in the western tropical Pacific region over this period. Using this ๐›ฟ18O composite, we compared the distribution of events in the period before and after the industrial revolution using the two-sample Kolmogorov-Smirnov (KS) test. We found no statistically significant change in the distribution of ๐›ฟ18O anomalies. The same statistical test was applied to the Ninฬƒo 3.4 reconstruction from the Last Millennium Reanalysis (LMR). The results of this study suggest that if there is indeed a forced response of ENSO, it is as yet indetectable. This may be because the forcing is not yet large enough or the forced response is small relative to the unforced variability. Additional factors that might explain this result in the ๐›ฟ18O composite include its observational and interpretational uncertainty, and in the LMR reconstruction, the scarcity of tropical observational constraints and systematic error in the representation of ENSO in climate simulations. NO CHANGE IN ENSO HYDROCLIMATE VARIABILITY AFTER THE INDUSTRIAL REVOLUTION AS RECORDED IN ๐›ฟ18O OF TECTONA GRANDIS L.F. FROM SOUTHEAST SULAWESI, INDONESIA by Sandy Hardian Susanto Herho Thesis submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial fulfillment of the requirements for the degree of Master of Science 2023 Advisory Committee: Dr. Michael N. Evans, Chair/Advisor Dr. Karen L. Prestegaard Dr. Raden D. Susanto ยฉ Copyright by Sandy Hardian Susanto Herho 2023 Acknowledgments I am grateful to Austin Carter, Austin Mickels, Cristy Ho, Jenna Wollney, and Mary Gbenro for laboratory support. I also benefitted from helpful discussions with Faiz Fajary, Karen Prestegaard, Michael N. Evans, and R. Dwi Susanto. I also thank Ann Wylie (through the Green Fellowship in Global Climate Change) and Neera Gupta, who partially financed my study at UMD. This study was funded by the NSF grant AGS1903626. ii Table of Contents Acknowledgements ii 1 Introduction 1 1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.2 Is ENSO changing? Evidence from climate models . . . . . . . . . . . . . 1 1.3 Is ENSO changing? Evidence from paleoclimatology . . . . . . . . . . . . 3 1.4 This study . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 1.4.1 Study location . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 1.4.2 Data modeling: ๐›ฟ18O of ๐›ผโˆ’cellulose . . . . . . . . . . . . . . . . . 6 1.4.3 Goals of this study . . . . . . . . . . . . . . . . . . . . . . . . . . 8 2 Materials and methods 9 2.1 Stable isotope analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 2.1.1 Imaging, intra-annual sampling, and ๐›ผโˆ’cellulose extraction . . . . . 9 2.1.2 ๐›ฟ18๐‘‚๐‘ data acquisition . . . . . . . . . . . . . . . . . . . . . . . . 10 2.2 Statistical analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 2.2.1 Time-series compositing . . . . . . . . . . . . . . . . . . . . . . . 11 2.2.2 Nonparametric statistical test . . . . . . . . . . . . . . . . . . . . . 13 3 Results 15 4 Discussion 22 4.1 Summary of results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 4.2 Comparison with prior studies . . . . . . . . . . . . . . . . . . . . . . . . 22 5 Conclusion 24 A Statistical analyses for other embedding dimensions of the high-pass filtered time series 26 Bibliography 33 iii Chapter 1: Introduction 1.1 Motivation The El Ninฬƒo/Southern Oscillation (ENSO) phenomenon produces the most influential impacts on the climate of the tropical Pacific, where it occurs, as well as having far- reaching global consequences [11, 22, 60]. One of the areas that most affected by ENSO is the western equatorial Pacific, since ENSO is a major contributor to the interannual and interseasonal atmospheric convergence and precipitation in this region [50]. While the interannual variability of ENSO under normal conditions (without external radiative forcings) has been extensively studied, it is currently uncertain whether the patterns of its occurrence are significantly affected by external radiative forcing that arise from human activities, such as greenhouses gas (GHG) emissions [16, 17]. 1.2 Is ENSO changing? Evidence from climate models Several studies using climate models indicate an increase in ENSO cold and warm phase event amplitude, suspected to be related to anthropogenic forcing since the industrial revolution [39, 44, 48]. This occurs because of a balance between the release of latent heat in precipitation and radiative cooling in the troposphere [2]. This fact results in an increase in ENSO-driven precipitation in the western equatorial Pacific as a consequence of an increase in GHG emissions. However, the forcing of cold phase events by GHG increases 1 is small relative to what has been estimated from natural forcings [44]. This ENSO-driven precipitation intensification process in the western equatorial Pacific can be explained by the ocean-dynamical thermostat theory [5, 10, 14, 44]. Increased warming at the surface will have more impact on the western equatorial Pacific than on the eastern equatorial Pacific, this is because warming in the eastern region is offset by the upwelling process. This increase in the SST gradient then exerts an additional force on the pressure gradient which causes the Walker circulation to become stronger, and ultimately increases the SST gradient even more. This mechanism is known as Bjerknes feedback [44]. This finding is still debatable because the consensus results from the simulations of the Coupled Model Intercomparison Project-Phase 3 and 5 (CMIP3 and CMIP5) that the Walker circulation weakens, which will reduce trade winds and make the equatorial thermocline lower, resulting in more โ€El Ninฬƒo-likeโ€ phenomena [62, 63, 69]. Another study conducted using the Community Earth System Model-Last Millennium Ensemble (CESM-LME) project by Stevenson et al. [59] revealed that the dominant ENSO variability in the last millennium was governed by internal variability. Although there appears to be an increase in ENSO variability in the 20th century, this signal is not robust. However, using the Last Millennium experiments of CMIP5 and Paleoclimate Modelling Intercomparison Project-Phase 3 (PMIP3), Lewis and LeGrande [42] and Hope et al. [37] instead found an increase in ENSO variability in the historical period (1906 - 2005) as a result of an increase in GHG concentrations and land-use changes after the industrial revolution. It can be concluded that there is still debate about how ENSO changes in response to anthropogenic forcings as projected by various climate models after the industrial revolution [17, 32, 41, 72]. 2 1.3 Is ENSO changing? Evidence from paleoclimatology Additional studies have analyzed paleoclimatic observations and reconstructions of ENSO indicators for changes in properties before and after the anthropogenic increase in radiative forcing of climate, and to establish the range of natural, or unforced variability[e. g. 15, 43, 45, 46, 53, 56]. For example, many study area selections include broad networks of observations [e. g. 43], but are based solely on statistical correlation [e. g. 45] rather than on the forward mechanistic modeling of proxies [e. g. 24]. Others [e. g. 20] are from centers of action for ENSO, but have relatively low reconstruction skill. Leveraging forward modeling of proxies with respect to their environmental controls [26], paleoclimatic data assimilation can be used to update climate model simulations with proxy observations [e. g. 33]. In the approach of Hakim et al. [33] and related studies [73, 74], x๐‘Ž = x๐‘ +K(yโˆ’H(x๐‘)), (1.1) , where x๐‘Ž is the posterior estimate of Ninฬƒo 3.4 reconstruction, x๐‘ is the prior estimation from climate simulations, K is a Kalman gain matrix, y is the paleoclimatic observations, and H is the forward operator used to map x๐‘ to the proxy domain. Note that in this formulation, variance change over time in these products may be convolved with changes associated with observational availability y over time [e. g. 66]. In the work of Zhu et al. [73, 74], CCSM4-LME simulations [40] were used as the prior, and only coral-based observations and tree-ring records previously shown to record ENSO activity [19, 21, 43] were used for observations. In this study, Zhu et al. [73] found no evidence of an ENSO-mediated response to radiative forcing associated with explosive volcanic events. Historical analysis of gridded SST data, based on dense observations relative to paleoclimatic datasets, together with ENSO modeling suggest no forced response for the period 1856-2003 [12]. Overall, 3 there does not appear to be consensus from these studies on the anthropogenically forced response to ENSO since the Industrial Revolution. Thus, for the paleoclimatic strategy, it would be beneficial to acquire paleoclimatic data, whose interpretation is underpinned by mechanistic understanding, and which is from key locations in which there is an ENSO- related environmental response, to improve assessments of the sensitivity of ENSO to post-industrial anthropogenic forcing. 1.4 This study 1.4.1 Study location We addressed this problem by using the annually resolved oxygen isotopic composi- tion of ๐›ผโˆ’cellulose (๐›ฟ18๐‘‚๐‘) from teak (Tectona grandis L. f.) increment tree cores from Muna, Southeast Sulawesi, Indonesia. These samples were previously and independently dendrochronologically crossdated by Dโ€™Arrigo et al. [20] spanning the period 1656-2005 with a sample depth for each year of at least six increment cores per year [49]. The oxygen isotopic composition of cellulose in tree rings in Southeast Asia has been successfully used for reconstructing precipitation and ENSO-driven hydroclimatic responses [51, 58, 68, 75]. Muna is located in in the western equatorial Pacific, known as the Maritime Continent (MC) [52](20โ—ฆS - 20โ—ฆN, 90โ—ฆE - 160โ—ฆE, Figure 1.1). This area is directly affected by ENSO from a hydroclimatological perspective. Many studies have been conducted that specifically discuss the close relationship between the variability of precipitation over the MC and the ENSO [29, 34, 47, 54, 70]. The negative precipitation anomaly over the MC is considered to have a strong relationship with the ENSO warm phase. The precipitation deficit over the MC is generally associated with a large-scale air mass subsidence anomaly, a weakening of the Walker circulation, and a negative anomaly of sea surface temperature (SST) in the oceans surrounding the MC [54]. Muna is located on the border of region A and region C 4 in the study of monthly precipitation zone delineation using the double correlation method (DCM) by Aldrian and Susanto [1]. Region A has one peak, as a result of the wet northwest monsoon from November to April, and one trough, which occurs as a consequence of the dry southeast monsoon in the period March to May, in its average annual precipitation pattern. Meanwhile, region C has peak rainfall from June to July and one trough from November to February. Based on the precipitation spectral analysis conducted by Aldrian and Susanto [1], it is known that ENSO has a strong interannual effect on the two regions which are considered to represent the Muna region. Figure 1.1: Map showing the study location (red dots) in a tropical lowland rain forest (10 m.a.s.l) in the southeastern part of Sulawesi (5.3โ—ฆS, 123โ—ฆE). 5 1.4.2 Data modeling: ๐›ฟ18O of ๐›ผโˆ’cellulose Nurhati et al. [49] used a forward model of ๐›ฟ18๐‘‚๐‘ and comparison of simulations with triplicate observations for the period 1971-2005 to show that ENSO-driven precipitation acts as the main control of ๐›ฟ18๐‘‚๐‘ in this location. They used the ๐›ฟ18๐‘‚๐‘ model of Barbour et al. [4], as modified by Evans [24] for tropical regions, ๐›ฟ18๐‘‚๐‘ = ๐‘“ ( ๐‘‡, ๐‘๐‘Ÿ, ๐‘…๐ป ) (1.2) , where โ€ข ๐‘‡ : monthly average air temperature (โ—ฆC), โ€ข ๐‘๐‘Ÿ: monthly precipitation (mm), โ€ข ๐‘…๐ป: relative humidity (%), โ€ข ๐‘“ : forward-deterministic model function as described in Evans [24]. Here, ๐‘“ parameterizes the isotopic composition of precipitation (๐›ฟ18๐‘‚๐‘) as a linear function of precipitation amount, and of atmospheric water vapor (๐›ฟ18๐‘‚๐‘ฃ) as a constant value of 8โ€ฐ lower than ๐›ฟ18๐‘‚๐‘. ๐›ฟ18๐‘‚๐‘ can be modeled deterministically into the following equations, ฮ”๐‘ = ฮ”๐‘™ ( 1โˆ’ ๐‘๐‘ฅ ๐‘๐‘’๐‘ฅ ) + Y๐‘ (1.3) ฮ”๐‘™ = { (1+ Yโˆ—) [1+ Y๐‘˜ + (ฮ”๐‘ค๐‘ฃ๐‘Ž โˆ’ Y^ ๐‘’๐‘  ๐‘’๐‘– ) โˆ’1] } ( 1โˆ’ ๐‘’๐‘ƒ๐‘’ ๐‘ƒ๐‘’ ) . (1.4) Here, ฮ” is defined as the oxygen isotopic composition relative to the oxygen isotopic composition of the source water absorbed by plants, which is assumed to be equal to 6 unmodified ๐›ฟ18๐‘‚๐‘. The ( 1โˆ’ ๐‘๐‘ฅ ๐‘๐‘’๐‘ฅ ) fraction of the leaf water isotopic fractionation relative to source water (ฮ”๐‘™) is an isotope modification caused by the evaporation from leaves and by a partial re-equilibration process between the leaves and unmodified stem water (equation 1.3). The leaf evaporative enrichment (ฮ”๐‘™) is calculated as a function of the in- situ modified oxygen isotopic composition, temperature-dependent equilibrium and kinetic fractionation factors, and water vapor isotopic fractionation driven by leaf-air differences in specific humidity, modified by the Peฬclet effect (equation 1.4). Within this model, ๐›ฟ18๐‘‚๐‘ is predicted from the oxygen isotopic composition of precipitation, calculated evaporative enrichment at the leaf, partial re-equilibration of leaf water by Peฬclet back-diffusion, partial re-equilibration of photosynthate with unmodified stem water, and fractionation at cellulose synthesis. Evans [24] adapted the model from Barbour et al. [4] by parameterizing leaf temperature from air temperature, ๐›ฟ18๐‘‚๐‘ as a linear function of precipitation rate, cloud temperature as a linear function of moist adiabatic lapse rate and condensation level, and leaf stomatal conductance as an inverse function of leaf-air saturation vapor pressure deficit, calculable from temperature and relative humidity [24]. Thus, this model can be run with the inputs in equation (1.2). The temporal correlation between triplicate-observed Muna ๐›ฟ18๐‘‚๐‘, and November- April average ๐›ฟ18๐‘‚๐‘ simulation, produced from Climate Research Unit (CRU) TS3.01 temperature, precipitation and humidity [35] data at Muna, is statistically significant (1971- 2005: ๐‘Ÿ = 0.36, ๐‘’๐‘‘๐‘“ = 35, ๐‘ = 0.03) [49]. As predicted by the ๐›ฟ18๐‘‚๐‘ anomalies produced for ENSO events in an isotopically enabled reanalysis of atmospheric climate [71], the observed Muna ๐›ฟ18๐‘‚๐‘ record is significantly correlated with the Ninฬƒo 3.4 index (SST anomalies (averaged over 5โ—ฆN - 5โ—ฆS, 170โ—ฆW - 120โ—ฆW; 1971-2005, Sep-Oct annual averages: ๐‘Ÿ = 0.52, ๐‘‘๐‘“ = 35, ๐‘ < 0.01), and the statistically significant pattern correlation of reanalysis ๐›ฟ18๐‘‚๐‘ [71] is positive over the MC and negative over the central equatorial Pacific [49], consistent 7 with climatological understanding of ENSO as described earlier. 1.4.3 Goals of this study Here we assume that the same mechanistically developed interpretation applies to inter- pretation of the interannual component of Muna composite ๐›ฟ18๐‘‚๐‘ developed for the period extending into the pre-industrial past. By comparative analysis of data for the periods before and after the exponential increase in CO2 emissions starting in the mid-19th century, we can assess whether there was an associated change in ENSO activity. Specifically, we can test the prevailing hypothesis [7โ€“9] that the probability of ENSO-hydroclimate extremes due to global warming has increased in the industrial period [7โ€“9]. To test the sensitivity of the results and interpretation to uncertainty in the observation and interpretation of the Muna ๐›ฟ18๐‘‚๐‘ record, we can perform the same analysis but using as input the independently devel- oped Ninฬƒo 3.4 index from the Last Millennium Reanalysis โ€turboโ€ (LMRt) paleoclimatic data assimilation [73, 74], for which 1941-2000 Ninฬƒo 3.4 validation ๐‘Ÿ2 = 0.62). 8 Chapter 2: Materials and methods 2.1 Stable isotope analysis In this study, we sampled teak (Tectona grandis L. f.) increment cores in Muna, Southeast Sulawesi, Indonesia (5.3โ—ฆS, 123โ—ฆE) (Figure 1.1) which was dendrochronologically cross- dated by Dโ€™Arrigo et al. [20]. We analyzed three crossdated samples, tg01gc (1712 - 1995), tg11a (1680 - 1848), and mun6.3 (1798 - 1987). Procedures for preparing samples for isotopic analysis included sample imaging, microtoming, extracting ๐›ผโˆ’cellulose, and weighing and wrapping samples in silver capsules. To acquire ๐›ฟ18๐‘‚๐‘ data, we thermally converted samples to analyte gas in an elemental analyzer (EA), introduced the analyte into a stable isotope ratio mass spectrometer (IRMS) via a continuous flow interface, collected carbon and oxygen data, and calibrated the data using a 2-point correction for mean and variance to international reference scales [25]. All laboratory work was performed at the Paleoclimate CoLaboratory of the University of Maryland, College Park. 2.1.1 Imaging, intra-annual sampling, and ๐›ผโˆ’cellulose extraction Sample cores were scanned at 1200 dpi to match the core names, orientations, and scales before being separated from their wooden bases. We used a combination of rotary microtome and manual sampling using a razor blade to subsample each crossdated growing season into at least six sequential samples per year. Samples were placed into 1.5 ml centrifuge tubes for ๐›ผโˆ’cellulose extraction following a modified Brendel method; a detailed description is 9 in Anchukaitis et al. [3]. This procedure was done to isolate ๐›ผโˆ’cellulose, for which we have a well-developed and tested mechanistic interpretative model (see subsection 1.4.2) [24, 27] from other wood components. This extraction process was carried out by adding nitric acid and acetic acid in a ratio of 1:10 to the sample tubes, then capped sample tubes were heated to 119 - 124โ—ฆC for 30 minutes. Next, samples were sequentially rinsed with ethanol, then water, then ethanol, then acetone, before being dried at 45โ—ฆC for 30 minutes. Extracted ๐›ผโˆ’cellulose was then stored in a vacuum dessicator for 24 hours or until isotopic analysis. Then each of these dried samples was wrapped and weighted to 200ยฑ20 `g or 100ยฑ10 `g (depending on the sample size) in a silver capsule to be analyzed for isotopic composition. 2.1.2 ๐›ฟ18๐‘‚๐‘ data acquisition For the analysis of ๐›ฟ18๐‘‚๐‘, both samples and standards were run so data could be normalized, allowing drift in the instrument to be accounted for and comparisons to be made with other datasets. Samples and standards were loaded into the autosampler in order, and names and weights were filled into the batch processing file, with care taken to ensure the information matched the order of the samples and standards in the autosampler. The autosampler was then purged with helium to remove any remaining traces of air. Samples were left under helium overnight to make sure they were dry before running the analysis. Once each run was started, C and O from the samples were converted into CO via thermal conversion at 1080โ—ฆC, and CO2 and H2O were removed with an acid/water trap before sample CO was admitted to the IRMS by continuous flow interface. Currents measured for masses 28 (12C16O), 29 (13C16O), and 30 (12C18O) were used to determine isotope ratios. Two independent working standards, SAC (๐›ฟ13๐ถ = โˆ’25.13ยฑ 0.19 โ€ฐ; ๐›ฟ18๐‘‚ = 31.0ยฑ 0.24 โ€ฐ) and AKC (๐›ฟ13๐ถ = โˆ’13.80ยฑ0.10 โ€ฐ; ๐›ฟ18๐‘‚ = 23.61ยฑ0.24 โ€ฐ) were used for correction 10 of systematic mean and variance biases introduced by drift and amplitude through each sample batch using scripted functions in MATLABยฎ [25]. Diagnostic plots were used to check the correction algorithm produced valid corrections. Uncertainty in the isotopic data was estimated as the standard deviation of replicate corrected working standard data across all working standards, and is estimated as 0.4 โ€ฐ averaged across 36 batches of samples and working standards. 2.2 Statistical analysis 2.2.1 Time-series compositing A compositing step is required to produce an annual unidimensional time series of ๐›ฟ18๐‘‚๐‘ based on the three cores of differing time intervals and intra-annual resolutions. The first step was random sampling with replacement of a sub-annual ๐›ฟ18๐‘‚๐‘ value from each available annual increment. This step produced 1000 realizations of each of the three time series. We then subtracted each series mean, producing anomaly series, and to high-pass filter for analysis of ENSO-band variations, we used singular spectrum analysis (SSA). SSA is a nonparametric time-series analysis method which can be implemented on series with missing data [28, 30, 36, 57, 61]. SSA can be divided into two main steps: the decomposition, which consists of embedding and singular value decomposition (SVD), and the reconstruction, which consists of grouping and diagonal averaging [30, 36]. The embedding was implemented for each anomaly realization, a unidimensional series Y๐‘ = {๐‘ฆ1, ยท ยท ยท , ๐‘ฆ๐‘ } to be mapped into a multi-dimensional series X1, ยท ยท ยท ,X๐พ , with vector x๐‘– = (๐‘ฆ๐‘–, ยท ยท ยท , ๐‘ฆ๐‘–+๐ฟโˆ’1)๐‘‡ โˆˆ R๐ฟ , where ๐พ = ๐‘ โˆ’ ๐ฟ + 1. The window length ๐ฟ has to satisfy the requirement: 2 โ‰ค ๐ฟ โ‰ค ๐‘/2. The result of this embedding calculation is a trajectory matrix X with size ๐ฟร—๐พ which is a Hankel matrix, where all anti-diagonal elements have the same 11 value, expressed in the following form, X = ( x๐‘– ๐‘— )๐ฟ,๐พ = ยฉยญยญยญยญยญยญยญยญยญยซ ๐‘ฆ1 ๐‘ฆ2 . . . ๐‘ฆ๐พ ๐‘ฆ2 ๐‘ฆ3 . . . ๐‘ฆ๐พ+1 ... ... . . . ... ๐‘ฆ๐ฟ ๐‘ฆ๐ฟ+1 . . . ๐‘ฆ๐‘ ยชยฎยฎยฎยฎยฎยฎยฎยฎยฎยฌ . (2.1) After the embedding calculation, the decomposition step was continued by applying SVD to the trajectory matrix X. We began SVD analysis by defining the covariance matrix S = XX๐‘‡ . Then using the following equation, det(Sโˆ’_I) (2.2) , we determined the eigenvectors U๐‘– and their corresponding eigenvalues _๐‘–. This procedure was followed by constructing a singular value matrix ๐šบ, ๐šบ = ยฉยญยญยญยญยญยญยซ โˆš _1 . . . 0 ... . . . ... 0 . . . โˆš _๐ฟ ยชยฎยฎยฎยฎยฎยฎยฌ . (2.3) The principal component matrix is the transpose of the V๐‘– matrix, which was calculated through the following procedure, V๐‘‡ ๐‘– = ( X๐‘‡U๐‘–โˆš _๐‘– )๐‘‡ . (2.4) The final result of the SVD is the decomposition of the trajectory matrix X into eigentriple 12 as follows, X = U๐‘–๐šบV๐‘‡ ๐‘– . (2.5) The matrix X๐‘– = X1 +X2 + ยท ยท ยท +X๐‘‘ from the eigentriple was then used to reconstruct the time series through a grouping process by grouping the set of indices ๐‘– = {1,2, ยท ยท ยท , ๐‘‘} into a subset of disjoint ๐ผ = {๐ผ1, ๐ผ2, ยท ยท ยท ๐ผ๐‘š}, with ๐‘š = ๐‘‘, so that it could be expanded into this following form, X๐ผ = X๐ผ1 +X๐ผ2 + ยท ยท ยทX๐ผ๐‘š . (2.6) The final step to reconstruct the time series based on the power of the spectrum is to perform diagonal averaging. This stage was performed by transforming each matrix in the equation 2.6 into a Hankel matrix to produce the initial time series form as follows, y๐‘ก = ๐‘šโˆ‘๏ธ ๐‘˜=1 ๏ฟฝฬƒ๏ฟฝ๐‘š๐‘ก , , where ๐‘ก = 1, ยท ยท ยท๐‘š. (2.7) Because we aimed to do high-pass filtering on each anomaly realization to get interannual variability, we substituted๐‘š = 7 years and removed ๏ฟฝฬƒ๏ฟฝ1. The ensemble of 3ร—1000 high-pass filtered ๐›ฟ18๐‘‚๐‘ anomaly timeseries, spanning the entire observational period 1712-1995, were then used to estimate the composite anomaly median and 95th percentile confidence intervals. We implemented all of the time-series compositing processes in the MATLABยฎ computing environment. 2.2.2 Nonparametric statistical test The Kolmogorov-Smirnov (KS) test is a nonparametric test of continuous, one-dimensional probability distributions that can be used to compare a sample with a reference probability distribution (one-sample KS test), or to compare two samples (two-sample KS test) [65, 67]. In this study, we used the two-sample KS test. A nonparametric statistical test was used 13 because it does not assume normality in the time series distribution. The KS test was used to test the null hypothesis that two independent samples come from populations with identical distributions. โ€ข ๐ป0: ๐น (๐‘ฅ) = ๐บ (๐‘ฅ), for all ๐‘ฅ. โ€ข ๐ป1: ๐น (๐‘ฅ) โ‰  ๐บ (๐‘ฅ), for at least one value of ๐‘ฅ. , where ๐น (๐‘ฅ) and ๐บ (๐‘ฅ) are the cumulative distribution functions (CDFs) of highpassed ๐›ฟ18O๐‘ anomalies (hereafter annual ๐›ฟ18๐‘‚๐‘ variability) for the population before and after the industrial revolution, respectively. Let ๐‘†๐‘›1(๐‘‹) and ๐‘†๐‘›2(๐‘‹) be the empirical cumulative distribution functions of the ๐›ฟ18O๐‘ anomalies before and after the baseline periods, respec- tively, given by ๐‘†๐‘›1 = ๐‘˜/๐‘›1 and ๐‘†๐‘›2 = ๐‘˜/๐‘›2, where ๐‘˜ is the same or less than ๐‘‹ , and ๐‘›1 and ๐‘›2 are the number of data points in each observation. The KS statistic is ๐ท = max |๐‘†๐‘›1(๐‘‹) โˆ’ ๐‘†๐‘›2(๐‘‹) |. (2.8) To test the sensitivity of the test to definition of the pre-industrial and industrial eras, we used two defensible definitions: (1) 1850 and (2) 1970, defined by the initial point at which anthropogenic greenhouse forcing is estimated to have begun, and the point at which the decadally averaged rate of increase in greenhouse radiative forcing suddenly increased, respectively [72]. We also tested the sensitivity of the results to sample size imbalance by repeating the latter test for independent 25-year time windows before and after 1970. We used the scipy library [64] in the Python computing environment for test calculations, and consider p๐‘๐‘Ÿ๐‘–๐‘ก = 0.05 for inference. We applied the same test and definitions to the independent December-February (DJF) Ninฬƒo-3.4 index from the LMRt paleoclimate data assimilation product.1 1Data available at: https://zenodo.org/record/5893781.ZAqnh3WYXIE. 14 https://zenodo.org/record/5893781##.ZAqnh3WYXIE Chapter 3: Results There are 1,356 sub-annual measurements of ๐›ฟ18๐‘‚๐‘ on tg01c, 788 on tg11a, and 1,048 on mun6.3. Intra-annual resolution ๐›ฟ18๐‘‚๐‘ measurements are shown as time series in Figure 3.1. From tg01c, the maximum ๐›ฟ18๐‘‚๐‘ is 30.7 โ€ฐ in 1928, and the minimum is 19.3 โ€ฐ in 1765 (Figure 3.1a). From tg11a, the maximum value of ๐›ฟ18๐‘‚๐‘ is 28.7 โ€ฐ in 1686, and the minimum value is 22.2 โ€ฐ in 1720 (Figure 3.1b). From mun6.3, the maximum ๐›ฟ18๐‘‚๐‘ is 32.2 โ€ฐ in 1981, and the minimum value is 23.0 โ€ฐ in 1934 (Figure 3.1c). The median and 2.5th - 97.5th percentile for each core are shown in Table 3.1 (๐›ฟ18๐‘‚๐‘ values). Series medians differ by almost 2 permil (Table 3.1), and the 95th percentile ranges of values are about 4 โ€ฐ, 4 โ€ฐ, and 5 โ€ฐ, respectively (Table 3.1, ๐›ฟ18๐‘‚๐‘ values). core name ๐›ฟ18O๐‘ ๐›ฟ18O๐‘ variability median (โ€ฐ) [2.5th, 97.5th] (โ€ฐ) median (โ€ฐ) [2.5th, 97.5th] (โ€ฐ) tg01c 25.53 [23.56, 27.52] 0.01 [-1.43, 1.4] tg11a 25.32 [23.41, 27.27] 0.03 [-1.54, 1.34] mun6.3 27.33 [24.95, 29.75] -0.01 [-1.94, 2.22] Table 3.1: Median and 2.5th - 97.5th percentile confindence limits for each core. Columns 2 - 3 are statistics for the intra-annual ๐›ฟ18O๐‘ series. Columns 4 - 5 are statistics for ๐›ฟ18O๐‘ variability by sample core prior to compositing across cores. The high pass ๐›ฟ18๐‘‚๐‘ composite is shown in Figure 3.2. The solid green line indicates the median values. Meanwhile, the green shaded area shows the 95% percentile interval obtained from 3ร—1000 randomized realizations. The highest median of the ๐›ฟ18๐‘‚๐‘ variability is 1.40 โ€ฐ in 1712, and the lowest is -1.90 โ€ฐ in 1706. The grand median of ๐›ฟ18๐‘‚๐‘ variability 15 (a) (b) (c) Figure 3.1: Intra-annual resolution ๐›ฟ18๐‘‚๐‘ measurements collected from Muna increment core samples: (a) tg01c, (b) tg11a, and (c) mun6.3. Vertical axis shows oxygen isotopic composition in permil units relative to the Vienna Standard Mean Ocean Water standard (VSMOW), with each scale range = 10 โ€ฐ centered on the median series value. The single point in black at 1700, 27.33 shows observational uncertainty for intra-annual measurement analytical uncertainty, as ยฑ2 standard deviations of the mean of corrected standard measurements. Sequential sample data were assigned sequential dates for Nov of the prior calendar year through April of the following calendar year by linear interpolation; horizontal axis shows time (years CE). 16 is 0.00 โ€ฐ, and the ensemble 95th percentile confidence interval is [-0.98 โ€ฐ, 0.88 โ€ฐ]. The ensemble ๐›ฟ18๐‘‚๐‘ variability is our best estimate of the site composite ๐›ฟ18๐‘‚๐‘ record at interannual timescales, sampling observational uncertainty and filtering for mechanistically supported timescales of variation. Figure 3.2: ๐›ฟ18๐‘‚๐‘ variability composite at annual resolution. Solid green line shows the median, while shaded region denotes the central 95% quantiles from the 3ร—1000 random realizations. Error bar at point (0,1670) shows ยฑ2 SE of the observational uncertainty for the annual ๐›ฟ18๐‘‚๐‘ variability composite, estimated as ๐‘ / โˆš ๐‘›, ๐‘› = 6 samples/year, and neglecting the reduction of variance associated with highpass filtering. The results of empirical CDF calculations of the ๐›ฟ18๐‘‚๐‘ variability from the pre- and post-industrial revolution periods are shown in Figure 3.3. Figure 3.3a shows the CDFs of the ๐›ฟ18๐‘‚๐‘ variability for the 1680 - 1849 (as the pre-industrial period) and the 1850 - 1995 (as the post-industrial period). Meanwhile, Figure 3.3b shows the CDFs of the ๐›ฟ18๐‘‚๐‘ variability for the 1695 - 1969 (as the pre-industrial period) and the 1970 - 1995 (as the post-industrial period). The results of the two-sample KS test showed that there 17 is no significant difference in the CDFs between the pre-and post-industrial periods, both when the industrial revolution began in 1850 (with statistical values: ๐ท = 0.072, ๐‘โˆ’value = 0.780), or when the industrial revolution began in 1970 (Table 3.2). This result was not sensitive to the choice of SSA embedding dimension for ๐‘š = 5,7,9,11 (Appendix A). Table 3.2: Statistical results of the two-sample KS test of ๐›ฟ18๐‘‚๐‘ anomaly for every 25 years compared to the period after the industrial revolution (1970 - 1995). period ๐ท๐‘š๐‘Ž๐‘ฅ ๐‘โˆ’value 1695 - 1719 0.177 0.733 1720 - 1744 0.142 0.910 1745 - 1769 0.217 0.495 1770 - 1794 0.268 0.233 1795 - 1819 0.226 0.424 1820 - 1844 0.188 0.652 1845 - 1869 0.148 0.882 1870 - 1894 0.137 0.929 1895 - 1919 0.134 0.942 1920 - 1944 0.266 0.241 1945 - 1969 0.257 0.297 As a comparison, we also applied the two-sample KS test to the LMRt DJF Ninฬƒo 3.4 reconstruction [73, 74] during the same period (1680 - 1995) (Figure 3.4a). Figure 3.4b shows the CDFs in 1680 - 1849 (as the pre-industrial period) and 1850-1995 (as the post-industrial period). Figure 3.4c shows the CDFs in the period 1680 - 1969 (as the pre-industrial period) and 1970-1995 (as the post-industrial period). Based on the a-priori ๐‘ = 0.05 critical level, we infer that there is no significant difference between the CDFs for the periods 1680 - 1849 and 1850 - 1995, with statistical values: ๐ท = 0.140 and ๐‘โˆ’value = 0.082 (Figure 3.4b). We make the same inference if the industrial revolution period is defined to start in 1970 (Table 3.3, Figure 3.4c). 18 Table 3.3: Statistical results of the two-sample KS test of Ninฬƒo 3.4 reconstruction for every 25 years compared to the period after the industrial revolution (1970 - 1995). period ๐ท๐‘š๐‘Ž๐‘ฅ ๐‘โˆ’value 1695 - 1719 0.345 0.060 1720 - 1744 0.346 0.059 1745 - 1769 0.228 0.413 1770 - 1794 0.268 0.233 1795 - 1819 0.305 0.128 1820 - 1844 0.268 0.233 1845 - 1869 0.305 0.128 1870 - 1894 0.300 0.147 1895 - 1919 0.285 0.216 1920 - 1944 0.331 0.098 1945 - 1969 0.337 0.079 19 (a) (b) Figure 3.3: Empirical cumulative probability density function (CDF) curves of the annual ๐›ฟ18๐‘‚๐‘ variability. The red line shows the period after the industrial revolution, while the blue lines show the periods before the industrial revolution. We used two different definitions of the anthropogenic era to show that the results are not sensitive to its definition of the beginning of the exponential increase in the anthropogenic forcings. In (a) the industrial revolution began in 1850, while in (b) the industrial revolution began in 1970. 20 (a) (b) (c) Figure 3.4: (a) DJF Ninฬƒo 3.4 reconstruction (1680 - 1995) using the LMR turbo (LMRt) framework [74] from the analysis of Zhu et al. [73], solid red line shows median values; shaded area denotes the 95% quantiles. (b - c) same as the Figure 3.3, but for the LMR data. 21 Chapter 4: Discussion 4.1 Summary of results We constructed an annually resolved site composite record from three temporally over- lapping, intra-annually resolved ๐›ฟ18๐‘‚๐‘ series. Based on prior work demonstrating a link to ENSO activity on interannual timescales, we infer no change in the inferred distribution of ENSO events through the 316 year (1680-1995) composite record. This result is sensitive to neither the parameters in data compositing and highpass filtering, nor the choice of startpoint of significant greenhouse gas radiative forcing as either 1850 or 1970. Our study shows no difference in ENSO variability as a result of changes in anthropogenic forcings regardless of the initial period of the exponential increase. 4.2 Comparison with prior studies Our results are independently replicated by reaching the same inference from analysis of LMRt Ninฬƒo-3.4 index reconstructions for the same intervals. However, we also find that in these results, some intervals give significances that are only slightly above the canonical statistically significant ๐‘โˆ’value (๐‘ < 0.05) (Table 3.3, Figure 3.4b, c). This result, which is most apparent for the 17th and 20th centuries (Table 3.3), may be due to artifacts in the LMR product. Historical observations [12]1 do not support the โ‰ˆ1๐‘œC warming in the 1https://iridl.ldeo.columbia.edu/SOURCES/.Indices/.nino/.EXTENDED/.NINO34/ figviewer.html?plottype=line. 22 https://iridl.ldeo.columbia.edu/SOURCES/.Indices/.nino/.EXTENDED/.NINO34/figviewer.html?plottype=line https://iridl.ldeo.columbia.edu/SOURCES/.Indices/.nino/.EXTENDED/.NINO34/figviewer.html?plottype=line reconstructed Ninฬƒo 3.4 index. One factor that might explain the 17th century change in distribution could be the tendency of data assimilation products to resolve less variance as observational coverage decreases. For the spurious 20th century warming and associated change in distribution of Ninฬƒo3.4 anomalies, we speculate that the forward operators used in the data assimilation may have transformed a calibration period trend into a reconstructed warming. Alternatively, this trend could arise from biased representation of ENSO, or the response of ENSO to radiative forcing, in the CCSM4-LME simulations. Further investigation is needed, both statistically (e.g. by using the False Discovery Rate (FDR) [18]) and dynamically (e.g. by correcting the cold tongue bias representation of the input model used in the LMRt framework [31]), to address this issue. Our findings do differ from the those of Liu et al. [45]. In their analysis of ๐›ฟ18๐‘‚๐‘ observations from Taiwan for 1190-2007 CE, they inferred an increase in ENSO variability at the end of the 20th century. However, Liu et al. [45] did not test statistical significance of variance changes in their study. 23 Chapter 5: Conclusion Our inference is consistent with most studies of paleo ENSO since the Pliocene, which suggests that the external radiatively forced response of ENSO may be small relative to the unforced internal variability [23]. This conclusion is provisional because there are considerable uncertainties in the observations, reconstructions and modeling of ENSO and its forced response. Nurhati et al. [49] showed that the interpretation of the Muna composite ๐›ฟ18๐‘‚๐‘ variability record explains less than 1/3 of the interannual variance associated with ENSO, and the Muna record is from only one location. Further confidence can be developed from paleoclimatic data assimilation, but products such as LMRt also have uncertainties associated with data modeling, changes in observational coverage over time, and biases in the prior climate states used. Finally, there is as yet no model consensus on the forced response of ENSO. At present, the evolution of ENSO amplitude, frequency, and variability of anthropogenic forcings is still poorly constrained, and there is no inter-model consensus on projected changes in the dynamics of ENSO diversity in response to the exponential increase in GHG since the industrial revolution [6, 13, 17]. A more in-depth physical investigation is needed on the relationship of anthropogenic forcings to changes in ENSO variability, for instance concerning ENSO diversity, represented by Central Pacific (CP) and Eastern Pacific (EP) events [38]. Historical observations suggest a high degree of inter-event variations from the mid- 19th century to the present, consistent with what an intermediate compexity ENSO forecast model might produce without anthropogenic forcings [12]. Paleoclimatic observations 24 and reconstructions extend this result further into the pre-anthropogenic past, but their analysis fails to disprove the hypothesis that the forced signal, if present, is difficult to detect within the large spread of the natural variability. Such efforts will be improved by the establishment of more high resolution observations in other ENSO-sensitive regions, such as equatorial South America [55], data modeling, and the further development of spatially resolved data assimilation products. Until either the climate simulations produce a clearer consensus, and actual radiative forcing and observed variability exceeds the historical and paleoclimatogically observed or reconstructed envelope, the effect of anthropogenic forcings on ENSO variability will remain unclear. 25 Appendix A: Statistical analyses for other embedding dimensions of the high-pass filtered time series To test that the computational results of our high-pass filtered time series are not sensitive to the arbitrary embedding dimension (๐‘š = 7), we performed high-pass SSA filtering with ๐‘š = 5, 9, and 11. The results are shown in Figure A.1. After extracting ๐›ฟ18๐‘‚๐‘ variabilities with different embedding dimensions, we applied the two-sample KS test as we did in subsection 2.2.2. Empirical CDFs for ๐‘š = 5, 9, and 11, respectively shown in Figures A.2, A.3, and A.4. KS test results comparing CDFs (for ๐‘š = 5, 9, and 11) in 1680 - 1850 with 1850 - 1995 are shown in Table A.1. The results of the KS tests for the post-1970 period for ๐‘š = 5, 9, and 11 are shown in Tables A.2, A.3, and A.4, respectively. The KS tests with these different embedding dimensions show no statistically significant differences (๐‘ < 0.05) between ๐›ฟ18๐‘‚๐‘ variabilities before and after the industrial revolution. Table A.1: Statistical results of the two-sample KS test of ๐›ฟ18๐‘‚๐‘ variabilities (๐‘š = 5,9,11) for 1680 - 1849 compared to the period after the industrial revolution (1850 - 1995). embedding dimension (๐‘š) ๐ท๐‘š๐‘Ž๐‘ฅ ๐‘โˆ’value 5 0.061 0.91 9 0.078 0.688 11 0.091 0.487 26 Table A.2: Statistical results of the two-sample KS test of ๐›ฟ18๐‘‚๐‘ variability (๐‘š = 5) for every 25 years compared to the period after the industrial revolution (1970 - 1995). period ๐ท๐‘š๐‘Ž๐‘ฅ ๐‘โˆ’value 1695 - 1719 0.138 0.923 1720 - 1744 0.178 0.721 1745 - 1769 0.218 0.483 1770 - 1794 0.191 0.627 1795 - 1819 0.226 0.424 1820 - 1844 0.186 0.664 1845 - 1869 0.111 0.987 1870 - 1894 0.138 0.923 1895 - 1919 0.092 0.999 1920 - 1944 0.226 0.424 1945 - 1969 0.302 0.140 Table A.3: Statistical results of the two-sample KS test of ๐›ฟ18๐‘‚๐‘ variability (๐‘š = 9) for every 25 years compared to the period after the industrial revolution (1970 - 1995). period ๐ท๐‘š๐‘Ž๐‘ฅ ๐‘โˆ’value 1695 - 1719 0.137 0.929 1720 - 1744 0.143 0.904 1745 - 1769 0.142 0.910 1770 - 1794 0.228 0.413 1795 - 1819 0.226 0.424 1820 - 1844 0.225 0.436 1845 - 1869 0.143 0.904 1870 - 1894 0.132 0.947 1895 - 1919 0.172 0.766 1920 - 1944 0.305 0.128 1945 - 1969 0.189 0.640 27 Table A.4: Statistical results of the two-sample KS test of ๐›ฟ18๐‘‚๐‘ variability (๐‘š = 11) for every 25 years compared to the period after the industrial revolution (1970 - 1995). period ๐ท๐‘š๐‘Ž๐‘ฅ ๐‘โˆ’value 1695 - 1719 0.112 0.985 1720 - 1744 0.143 0.904 1745 - 1769 0.103 0.994 1770 - 1794 0.225 0.436 1795 - 1819 0.223 0.448 1820 - 1844 0.192 0.616 1845 - 1869 0.143 0.904 1870 - 1894 0.209 0.554 1895 - 1919 0.135 0.936 1920 - 1944 0.343 0.063 1945 - 1969 0.228 0.413 28 (a) (b) (c) Figure A.1: Annual ๐›ฟ18๐‘‚๐‘ variability. Solid green line shows the median, while shaded region denotes the central 95% quantiles from the 3ร—1000 random realizations for (a) ๐‘š = 5, (b) ๐‘š = 9, (c) ๐‘š = 11. 29 (a) (b) Figure A.2: Empirical cumulative probability density function (CDF) curves of the annual ๐›ฟ18๐‘‚๐‘ variability (high-pass filtered ๐‘š = 5). The red line shows the period after the industrial revolution, while the blue lines show the periods before the industrial revolution. We used two different two different definitions of the anthropogenic era to show that the results are not sensitive to its definition of the beginning of the exponential increase in the anthropogenic forcings. In (a) the industrial revolution began in 1850, while in (b) the industrial revolution began in 1970. 30 (a) (b) Figure A.3: Empirical cumulative probability density function (CDF) curves of the annual ๐›ฟ18๐‘‚๐‘ variability (high-pass filtered ๐‘š = 9). The red line shows the period after the industrial revolution, while the blue lines show the periods before the industrial revolution. We used two different definitions of the anthropogenic era to show that the results are not sensitive to its definition of the beginning of the exponential increase in the anthropogenic forcings. In (a) the industrial revolution began in 1850, while in (b) the industrial revolution began in 1970. 31 (a) (b) Figure A.4: Empirical cumulative probability density function (CDF) curves of the annual ๐›ฟ18๐‘‚๐‘ variability (high-pass filtered ๐‘š = 11). The red line shows the period after the industrial revolution, while the blue lines show the periods before the industrial revolution. We used two different definitions of the anthropogenic era to show that the results are not sensitive to its definition of the beginning of the exponential increase in the anthropogenic forcings. In (a) the industrial revolution began in 1850, while in (b) the industrial revolution began in 1970. 32 Bibliography [1] E. Aldrian and R. D. Susanto. Identification of three dominant rainfall regions within Indonesia and their relationship to sea surface temperature. International Journal of Climatology, 23(12):1435โ€“1452, 2003. [2] M. R. Allen and W. J. Ingram. Constraints on future changes in climate and the hydrologic cycle. Nature, 419(6903):228โ€“232, Sep 2002. [3] K. J. Anchukaitis, M. N. Evans, T. Lange, D. R. Smith, S. W. Leavitt, and D. P. Schrag. Consequences of a rapid cellulose extraction technique for oxygen isotope and radiocarbon analyses. Analytical Chemistry, 80(6):2035โ€“2041, 2008. [4] M. Barbour, J. Roden, G. Farquhar, and J. Ehleringer. Expressing leaf water and cellulose oxygen isotope ratios as enrichment above source water reveals evidence of a Peฬclet effect. Oecologia, 138:426โ€“35, 03 2004. [5] E. Bauer, M. Claussen, V. Brovkin, and A. Huฬˆnerbein. Assessing climate forcings of the Earth system for the past millennium. Geophysical Research Letter, 30, 03 2003. [6] H. Bellenger, E. Guilyardi, J. Leloup, M. Lengaigne, and J. Vialard. ENSO represen- tation in climate models: From CMIP3 to CMIP5. Climate Dynamics, 42:1999โ€“2018, 2014. [7] W. Cai, S. Borlace, M. Lengaigne, P. Rensch, M. Collins, G. Vecchi, A. Timmermann, A. Santoso, M. McPhaden, L. Wu, M. England, G. Wang, E. Guilyardi, and F-F. Jin. Increasing Frequency of Extreme El Ninฬƒo Events due to Greenhouse Warming. Nature Climate Change, 4, 01 2014. [8] W. Cai, A. Santoso, M. Collins, B. Dewitte, C. Karamperidou, J-S. Kug, M. Lengaigne, M. J. McPhaden, M. F. Stuecker, A. S. Taschetto, et al. Changing El Ninฬƒoโ€“Southern Oscillation in a warming climate. Nature Reviews Earth & Environment, 2(9):628โ€“ 644, 2021. 33 [9] W. Cai, G. Wang, A. Santoso, M. McPhaden, L. Wu, F-F. Jin, A. Timmermann, M. Collins, G. Vecchi, M. Lengaigne, M. England, D. Dommenget, K. Takahashi, and E. Guilyardi. Increased frequency of extreme La Ninฬƒa events under greenhouse warming. Nature Climate Change, 5:132โ€“137, 01 2015. [10] M. Cane, A. Clement, A. Kaplan, Y. Kushnir, D. Pozdnyakov, R. Seager, S. Zebiak, and R. Murtugudde. Twentieth-Century Sea Surface Temperature Trends. Science, 275:957, 02 1997. [11] M. A. Cane. Forecasting El Ninฬƒo with a Geophysical Model. In M. H. Glantz, R. W. Katz, and N. Nicholls, editors, Teleconnections Linking Worldwide Climate Anomalies., chapter 11, pages 345โ€“369. Cambridge U. Press, Cambridge, England, 1991. [12] D. Chen, M. A. Cane, A. Kaplan, S. E. Zebiak, and D. Huang. Predictability of El Ninฬƒo over the past 148 years. Nature, 428(6984):733โ€“736, 2004. [13] L. Chen, T. Li, Y. Yu, and S. K. Behera. A possible explanation for the divergent projection of ENSO amplitude change under global warming. Climate Dynamics, 49:3799โ€“3811, 2017. [14] A. C. Clement, R. Seager, M. A. Cane, and S. E. Zebiak. An Ocean Dynamical Thermostat. Journal of Climate, 9(9):2190 โ€“ 2196, 1996. [15] Kim M Cobb, Christopher D Charles, Hai Cheng, and R Lawrence Edwards. El Ninฬƒo/Southern Oscillation and tropical Pacific climate during the last millennium. Nature, 424(6946):271โ€“276, 2003. [16] M. Collins. El Ninฬƒo or La Ninฬƒa-like climate change? Clim. Dyn., 24:89โ€“104, 2005. doi: 10.1007/s00382-004-0478-x. [17] M. Collins, S. I. An, W. Cai, A. Ganachaud, E. Guilyardi, F.-F. Jin, M. Jochum, M. Lengaigne, S. Power, A. Timmermann, G. Vecchi, and A. Wittenberg. The impact of global warming on the tropical Pacific Ocean and ENSO. Nature Geoscience, 3:391โ€“397, 2010. doi: 10.1038/NGEO868. [18] David Colquhoun. An investigation of the false discovery rate and the misinterpretation of p-values. Royal Society Open Science, 1(3):140216, 2014. [19] PAGES2k Consortium et al. A global multiproxy database for temperature reconstruc- tions of the common era. Scientific data, 4, 2017. [20] R. Dโ€™Arrigo, R. Wilson, J. Palmer, P. Krusic, A. Curtis, J. Sakulich, S. Bijaksana, S. Zulaikah, L. O. Ngkoimani, and A. Tudhope. The reconstructed Indonesian warm pool sea surface temperatures from tree rings and corals: Linkages to Asian monsoon drought and El Ninฬƒoโ€“Southern Oscillation. Paleoceanography, 21(3), 2006. 34 [21] S. G. Dee, K. M. Cobb, J. Emile-Geay, T. R. Ault, R. L. Edwards, H. Cheng, and C. D. Charles. No consistent ENSO response to volcanic forcing over the last millennium. Science, 367(6485):1477โ€“1481, 2020. [22] H. F. Diaz, M. P. Hoerling, and J. K. Eischeid. ENSO variability, teleconnections and climate change. International Journal of Climatology, 21(15):1845โ€“1862. [23] J. Emile-Geay, K. M. Cobb, J. E. Cole, M. Elliot, and F. Zhu. Past ENSO Variability, chapter 5, pages 87โ€“118. American Geophysical Union (AGU), 2020. [24] M. N. Evans. Toward forward modeling for paleoclimatic proxy signal calibration: A case study with oxygen isotopic composition of tropical woods. Geochemistry, Geophysics, Geosystems, 8(7), 2007. [25] M. N. Evans, K. J. Selmer, B. T. Breeden III, A. S. Lopatka, and R. E. Plummer. Correction algorithm for online continuous flow ๐›ฟ13C and ๐›ฟ18O carbonate and cellulose stable isotope analyses. Geochemistry, Geophysics, Geosystems, 17(9):3580โ€“3588, 2016. [26] M. N. Evans, S. E. Tolwinski-Ward, D. M. Thompson, and K. J. Anchukaitis. Appli- cations of proxy system modeling in high resolution paleoclimatology. Quaternary science reviews, 76:16โ€“28, 2013. [27] A. Gessler, J. P. Ferrio, R. Hommel, K. Treydte, R. A. Werner, and R. K. Monson. Sta- ble isotopes in tree rings: towards a mechanistic understanding of isotope fractionation and mixing processes from the leaves to the wood. Tree Physiology, 34(8):796โ€“818, 06 2014. [28] M. Ghil, M. R. Allen, M. D. Dettinger, K. Ide, D. Kondrashov, M. E. Mann, W. Robert- son, A, A. Saunders, Y. Tian, F. Varadi, et al. Advanced spectral methods for climatic time series. Reviews of Geophysics, 40(1):3โ€“1, 2002. [29] A. Giannini, A. Robertson, and J. Qian. A role for tropical tropospheric temperature adjustment to El Ninฬƒo-Southern Oscillation in the seasonality of monsoonal Indonesia precipitation predictability. Journal of Geophysical Research, 112, 08 2007. [30] N. Golyandina, A. Korobeynikov, and A. Zhigljavsky. Singular Spectrum Analysis with R. Springer, 2018. [31] E. Guilyardi, A. Capotondi, M. Lengaigne, S. Thual, and A. Wittenberg. ENSO Modeling: History, Progress, and Challenges, pages 199โ€“226. 11 2020. [32] Eric Guilyardi. El Ninฬƒoโ€“mean stateโ€“seasonal cycle interactions in a multi-model ensemble. Climate Dynamics, 26:329โ€“348, 2006. 35 [33] G. J. Hakim, J. Emile-Geay, E. J. Steig, D. Noone, D. M. Anderson, R. Tardif, N. Steiger, and W. A. Perkins. The last millennium climate reanalysis project: Frame- work and first results. Journal of Geophysical Research: Atmospheres, 121(12):6745โ€“ 6764, 2016. [34] J-I. Hamada, M. Yamanaka, J. Matsumoto, S. Fukao, P. Winarso, and T. Sribimawati. Spatial and Temporal Variations of the Rainy Season over Indonesia and their link to ENSO. Journal of the Meteorological Society of Japan, 80:285โ€“310, 04 2002. [35] I. Harris, P. D. Jones, T. J. Osborn, and D. H. Lister. Updated high-resolution grids of monthly climatic observations โ€“ the CRU TS3.10 Dataset. International Journal of Climatology, 34:623โ€“642, 2014. doi: 10.1002/joc.3711. [36] H. Hassani and R. Mahmoudvand. Singular spectrum analysis: Using R. Springer, 2018. [37] P. Hope, B. J. Henley, J. Gergis, J. Brown, and H. Ye. Time-varying spectral char- acteristics of ENSO over the Last Millennium. Climate Dynamics, 49:1705โ€“1727, 2017. [38] H-Y Kao and J-Yi Yu. Contrasting Eastern-Pacific and Central-Pacific Types of ENSO. Journal of Climate, 22(3):615 โ€“ 632, 2009. [39] B-H. Kim and K-J. Ha. Observed changes of global and western Pacific precipitation associated with global warming SST mode and Mega-ENSO SST mode. Climate Dynamics, 45, 02 2015. [40] L. Landrum, B. L. Otto-Bliesner, E. R. Wahl, A. Conley, P. J. Lawrence, N. Rosen- bloom, and H. Teng. Last Millennium Climate and Its Variability in CCSM4. Journal of Climate, 26(4):1085 โ€“ 1111, 2013. [41] M. Latif and N. S. Keenlyside. El Ninฬƒo/Southern Oscillation response to global warming. Proceedings of the National Academy of Sciences, 106(49):20578โ€“20583, 2009. [42] S. C. Lewis and A. N. LeGrande. Stability of ENSO and its tropical Pacific telecon- nections over the last millennium. Climate of the Past, 11(10):1347โ€“1360, 2015. [43] J. Li, S-P. Xie, E. R. Cook, M. S. Morales, D. A. Christie, N. C. Johnson, F. Chen, R. Dโ€™Arrigo, A. M. Fowler, X. Gou, et al. El Ninฬƒo modulations over the past seven centuries. Nature Climate Change, 3(9):822โ€“826, 2013. [44] J. Liu, B. Wang, M. A. Cane, S-Y. Yim, and J-Y. Lee. Divergent global precipitation changes induced by natural versus anthropogenic forcing. Nature, 493(7434):656โ€“659, Jan 2013. 36 [45] Y. Liu, K. M. Cobb, H. Song, Q. Li, C-Y. Li, T. Nakatsuka, Z. An, W. Zhou, Q. Cai, J. Li, et al. Recent enhancement of central Pacific El Ninฬƒo variability relative to last eight centuries. Nature Communications, 8(1):15386, 2017. [46] Z. Lu, Z. Liu, J. Zhu, and K. M. Cobb. A Review of Paleo El Ninฬƒo-Southern Oscillation. Atmosphere, 9(4):130, 2018. [47] J. L. McBride, M. R. Haylock, and N. Nicholls. Relationships between the Maritime Continent Heat Source and the El Ninฬƒoโ€“Southern Oscillation Phenomenon. Journal of Climate, 16(17):2905 โ€“ 2914, 2003. [48] W. J. Merryfield. Changes to ENSO under CO2 doubling in a Multimodel Ensemble. Journal of Climate, 19(16):4009 โ€“ 4027, 2006. [49] I. C. Nurhati, M. N. Evans, S. Y. Cahyarini, R. D. Dโ€™Arrigo, K. Y. Yoshimura, and S. H. S. Herho. ๐›ฟ18O of marine-influenced Tectona grandis l. f. from equatorial Indonesia: A local rainfall amount and remote ENSO indicator. in preparation for submission to Paleoceanography and Paleoclimatology. [50] J.P. Peixoto and A.H. Oort. Physics of Climate. American Institute of Physics, New York, 1991. [51] N. Pumijumnong, P. Payomrat, S. Buajan, A. Braฬˆuning, C. Muangsong, U. Chareon- wong, P. Songtrirat, K. Palakit, Y. Liu, and Q. Li. Teak Tree-Ring Cellulose ๐›ฟ13C, ๐›ฟ18O, and Tree-Ring Width from Northwestern Thailand Capture Different Aspects of Asian Monsoon Variability. Atmosphere, 12(6), 2021. [52] C. S. Ramage. ROLE OF A TROPICAL โ€œMARITIME CONTINENTโ€ IN THE ATMOSPHERIC CIRCULATION. Monthly Weather Review, 96(6):365 โ€“ 370, 1968. [53] B. Rein, A. Luฬˆckge, L. Reinhardt, F. Sirocko, A. Wolf, and W-C. Dullo. El Ninฬƒo variability off Peru during the last 20,000 years. Paleoceanography, 20(4), 2005. [54] A. Robertson, V. Moron, J. Qian, C-P. Chang, F. Tangang, E. Aldrian, T. Koh, and L. Juneng. The Maritime Continent monsoon, pages 85โ€“98. 04 2011. [55] M. Rodriguez-Caton, L. Andreu-Hayles, V. Daux, M. Vuille, A. M. Varuolo-Clarke, R. Oelkers, D. A. Christie, R. Dโ€™Arrigo, M. S. Morales, M. Palat Rao, A. M. Srur, F. Vimeux, and R. Villalba. Hydroclimate and ENSO Variability Recorded by Oxygen Isotopes From Tree Rings in the South American Altiplano. Geophysical Research Letters, 49(4):e2021GL095883, 2022. e2021GL095883 2021GL095883. [56] G. T. Rustic, A. Koutavas, T. M. Marchitto, and B. K. Linsley. Dynamical excitation of the tropical Pacific Ocean and ENSO variability by Little Ice Age cooling. Science, 350(6267):1537โ€“1541, 2015. 37 [57] D. H. Schoellhamer. Singular spectrum analysis for time series with missing data. Geophysical Research Letters, 28(16):3187โ€“3190, 2001. [58] K. Schollaen, I. Heinrich, B. Neuwirth, P. J. Krusic, R. D. Dโ€™Arrigo, O. Karyanto, and G. Helle. Multiple tree-ring chronologies (ring width, ๐›ฟ13C and ๐›ฟ18O) reveal dry and rainy season signals of rainfall in Indonesia. Quaternary Science Reviews, 73:170โ€“181, 2013. [59] S. Stevenson, A. Capotondi, J. Fasullo, and B. Otto-Bliesner. Forced changes to twentieth century ENSO diversity in a last millennium context. Climate Dynamics, 52:7359โ€“7374, 2019. [60] A. S. Taschetto, C. C. Ummenhofer, M. F. Stuecker, D. Dommenget, Karumuri A., R. R. Rodrigues, and S-W. Yeh. ENSO Atmospheric Teleconnections, chapter 14, pages 309โ€“335. American Geophysical Union (AGU), 2020. [61] R. Vautard and M. Ghil. Singular spectrum analysis in nonlinear dynamics, with ap- plications to paleoclimatic time series. Physica D: Nonlinear Phenomena, 35(3):395โ€“ 424, 1989. [62] G. A. Vecchi and B. J. Soden. Global Warming and the Weakening of the Tropical Circulation. Journal of Climate, 20(17):4316 โ€“ 4340, 2007. [63] G. A. Vecchi, B. J. Soden, A. T. Wittenberg, I. M. Held, A. Leetmaa, and M. J. Harrison. Weakening of tropical Pacific atmospheric circulation due to anthropogenic forcing. Nature, 441(7089):73โ€“76, 2006. [64] P. Virtanen, R. Gommers, T. E. Oliphant, M. Haberland, T. Reddy, D. Cournapeau, E. Burovski, P. Peterson, W. Weckesser, J. Bright, S. J. van der Walt, M. Brett, J. Wil- son, K. J. Millman, N. Mayorov, A. R. J. Nelson, E. Jones, R. Kern, E. Larson, C. J. Carey, Iฬ‡. Polat, Y. Feng, E. W. Moore, J. VanderPlas, D. Laxalde, J. Perktold, R. Cim- rman, I. Henriksen, E. A. Quintero, C. R. Harris, A. M. Archibald, A. H. Ribeiro, F. Pedregosa, P. van Mulbregt, and SciPy 1.0 Contributors. SciPy 1.0: Fundamental Algorithms for Scientific Computing in Python. Nature Methods, 17:261โ€“272, 2020. [65] R. E. Walpole, R. H. Myers, S. L. Myers, and K. Ye. Probability and Statistics for Engineers and Scientists, volume 5. Macmillan New York, 1993. [66] J. Wang, J. Emile-Geay, D. Guillot, J. E. Smerdon, and B. Rajaratnam. Evaluat- ing climate field reconstruction techniques using improved emulations of real-world conditions. Climate of the Past, 10(1):1โ€“19, 2014. [67] D. S. Wilks. Statistical Methods in the Atmospheric Sciences, volume 100. Academic Press, 2011. 38 [68] C. Xu, M. Sano, and T. Nakatsuka. Tree ring cellulose ๐›ฟ18O of Fokienia hodginsii in northern Laos: A promising proxy to reconstruct ENSO? Journal of Geophysical Research: Atmospheres, 116(D24), 2011. [69] S-W. Yeh, J-S. Kug, B. Dewitte, M-H. Kwon, B. P. Kirtman, and F-F. Jin. El Ninฬƒo in a changing climate. Nature, 461(7263):511โ€“514, 2009. [70] S. Yoden, S. Otsuka, N. J. Trilaksono, and T. W. Hadi. Recent Progress in Research on the Maritime Continent Monsoon, chapter 6, pages 63โ€“77. 2017. [71] K. Yoshimura, M. Kanamitsu, D. Noone, and T. Oki. Historical isotope simulation using Reanalysis atmospheric data. Journal of Geophysical Research: Atmospheres, 113(D19), 2008. [72] Z. Zhongming, L. Linong, Y. Xiaona, L. Wei, et al. AR6 Climate Change 2021: The Physical Science Basis. 2021. [73] F. Zhu, J. Emile-Geay, K. J. Anchukaitis, G. J. Hakim, A. T. Wittenberg, M. S. Morales, M. Toohey, and J. King. A re-appraisal of the ENSO response to volcanism with paleoclimate data assimilation. Nature Communications, 13(1):747, 2022. [74] F. Zhu, J. Emile-Geay, G. J. Hakim, R. Tardif, and A. Perkins. LMR Turbo (LMRt): a lightweight implementation of the LMR framework, 2021. [75] M. Zhu, L. Stott, B. Buckley, K. Yoshimura, and K. Ra. Indo-Pacific Warm Pool convection and ENSO since 1867 derived from Cambodian pine tree cellulose oxygen isotopes. Journal of Geophysical Research: Atmospheres, 117(D11), 2012. 39 Acknowledgements Introduction Motivation Is ENSO changing? Evidence from climate models Is ENSO changing? Evidence from paleoclimatology This study Study location Data modeling: ^18O of -cellulose Goals of this study Materials and methods Stable isotope analysis Imaging, intra-annual sampling, and -cellulose extraction 18Oc data acquisition Statistical analysis Time-series compositing Nonparametric statistical test Results Discussion Summary of results Comparison with prior studies Conclusion Statistical analyses for other embedding dimensions of the high-pass filtered time series Bibliography