Abstract
Oxygen isotopes (δ18O) are the most commonly utilized speleothem proxy and have provided many foundational records of paleoclimate. Thus, understanding processes affecting speleothem δ18O is crucial. Yet, prior calcite precipitation (PCP), a process driven by local hydrology, is a widely ignored control of speleothem δ18O. Here we investigate the effects of PCP on a stalagmite δ18O record from central Vietnam, spanning 45 – 4 ka. We employ a geochemical model that utilizes speleothem Mg/Ca and cave monitoring data to correct the δ18O record for PCP effects. The resulting record exhibits improved agreement with regional speleothem δ18O records and climate model simulations, suggesting that the corrected record more accurately reflects precipitation δ18O (δ18Op). Without considering PCP, our interpretations of the δ18O record would have been misleading. To avoid misinterpretations of speleothem δ18O, our results emphasize the necessity of considering PCP as a significant driver of speleothem δ18O.
Subject terms: Palaeoclimate, Geochemistry, Climate and Earth system modelling
This study finds that in-cave processes affect speleothem oxygen isotope records. Correcting for these processes improves agreement with other regional records and climate models, providing a more accurate reflection of past hydroclimate change.
Introduction
The effects of karst and in-cave processes on speleothem oxygen isotopes (δ18O) have been largely overlooked when using speleothems for the reconstruction of past climate variability. When precipitated under equilibrium conditions, the primary controls on the δ18O of speleothem calcite are cave temperature and drip water δ18O, which typically reflects the δ18O of precipitation (δ18Op) falling above the cave1. Karst hydrology can complicate these signals, but these complexities can often be constrained through hydrological modeling and cave monitoring, and typically only introduce drip water δ18O variability of < 1 ‰2,3.
In addition to karst hydrology, in-cave processes affect the fractionation of δ18O between drip water and calcite4–8, resulting in the disequilibrium precipitation of most speleothem calcites9. Among these processes is prior calcite precipitation (PCP), which refers to all calcite precipitation from infiltrating water prior to speleothem formation. A recent modeling study demonstrated that PCP can theoretically offset speleothem δ18O by several ‰4. This finding raises concerns that PCP could obscure temperature- and δ18Op-driven speleothem δ18O variability, making records susceptible to misinterpretation. Given the multitude of published speleothem δ18O records worldwide10, there is a critical need to further study this phenomenon.
PCP encompasses all calcite precipitation from infiltrating water prior to the drip water reaching the stalagmite. This includes calcite precipitation in the epikarst, on the cave ceiling, and on an overhanging stalactite or soda straw (Fig. 1)11,12. Crucially, PCP changes the geochemical signature of drip water. When CO2 degasses from drip water, lighter isotopes (12C and 16O) preferentially degas, enriching the remaining drip water HCO3− pool in heavier isotopes (13C and 18O). Cave-analog laboratory experiments show that the calcite that precipitates from the HCO3− pool reflects this enriched isotopic signature (higher δ13C and δ18O values)5,13. Similarly, Mg2+ and other trace elements (TEs) with a partition coefficient < 1 are preferentially excluded during calcite precipitation14, with Mg/Ca and other TE/Ca ratios increasing alongside δ18O and δ13C in the remaining solution (Fig. 1). Dry conditions enhance PCP because more air in the karst and slower infiltration rates increase the duration of calcite precipitation, leading to stronger enrichment of heavier isotopes and trace element concentrations in the drip water (higher δ18O, δ13C, and Mg/Ca), and hence the speleothem calcite that precipitates from that water. Thus, PCP-sensitive geochemical proxies often reflect local hydroclimate change.
Fig. 1. Schematic of prior calcite precipitation (PCP) in a cave system.
The left panel shows the flow path of drip water and calcite precipitation along the infiltration pathway. Green arrows denote CO2 fluxes, and blue dashed lines denote water flow. The right panel shows detailed changes in isotopes (δ18O and δ13C) and trace elements (Mg/Ca) during PCP in (a) the epikarst and (b) the cave ceiling.
While PCP is a well-known control of speleothem δ13C and Mg/Ca records12,14, it is generally not considered when interpreting speleothem δ18O. Traditionally, H2O exchange - the isotopic re-equilibration between HCO3− and the ambient H2O reservoir - is thought to eliminate the degassing signal (referred to hereafter as the PCP signal) in drip water HCO3− δ18O prior to speleothem formation. With enough time, H2O exchange, also known as buffering, reestablishes the initial pre-PCP δ18O value of the drip water HCO3− pool, thereby erasing the PCP signature15. However, laboratory experiments have demonstrated that H2O exchange takes much longer than previously estimated16. Using the revised estimate of the exchange time, recent modeling studies suggest that speleothem δ18O does preserve a PCP signal when H2O isotopic re-equilibration in the drip water is incomplete4,8. Importantly, ref. 4 demonstrates that incomplete H2O exchange is most likely to be preserved by speleothems when PCP occurs on the cave ceiling and/or stalactite. This is because the residence time of drip water in the epikarst far exceeds the time necessary for re-equilibration, whereas the residence time of drip water on the cave ceiling or stalactite may be sufficiently short to maintain isotopic disequilibrium during calcite precipitation.
Although geochemical models successfully simulate the effects of PCP on calcite δ18O, these effects have only recently been observed in speleothem δ18O records7,17. Using a correction method to remove the PCP signal from a speleothem δ18O record of the Holocene from central Vietnam, ref. 17 found that PCP obscured the δ18Op signal in portions of their δ18O record. They reinterpreted their record using the corrected δ18O, which resolved disagreements in their data and led to a deeper understanding of the local and regional components of the monsoon system in Southeast Asia during the Holocene, a relationship still unresolved over glacial time periods.
Here, we build upon this work by examining the impact of PCP on another previously published δ18O record from central Vietnam18. Using a geochemical model to remove the PCP signal from the speleothem δ18O record (see “Methods”), we demonstrate that PCP overprints the expected δ18Op signal in the record. After removing the PCP signal, the corrected δ18O record shows substantially improved agreement with regional δ18O speleothem records and climate model output, providing a more accurate estimation of past δ18Op variability in central Vietnam from 45 – 4 ka.
Results/discussion
Controls on δ18Op in central Vietnam
Central Vietnam receives rainfall from two isotopically distinct monsoon systems, the southwest monsoon (more negative δ18Op, 30% of annual rainfall from June-August) and the northeast monsoon (more positive δ18Op, 47% of annual rainfall from September-November) (Fig. 2)18. Importantly, large-scale atmospheric processes, rather than local rainfall amounts, control the δ18Op of both monsoon systems. Upstream rainout of moisture sourced from the Bay of Bengal and Indian Ocean drives southwest monsoon δ18Op19,20. While northeast monsoon δ18Op is not correlated with local rainfall amount21, the large-scale mechanisms driving its variability are less understood. Annual δ18Op, the signal recorded by the majority of speleothem records in central Vietnam, reflects a combination of rainfall from these two monsoon systems (seasonality), but not local rainfall amount21. On modern timescales, the proportion of southwest vs northeast rainfall amount, which is driven by the timing of the SW-NE wind direction shift during the southward migration of the ITCZ, controls interannual δ18Op variability21. However, less is known about central Vietnam δ18Op variability on longer timescales.
Fig. 2. Modern and pre-industrial iTRACE climatology.
a Map of total annual precipitation in Mainland Southeast Asia derived from Asian Precipitation-Highly Resolved Observational Data Integration Towards Evaluation (APHRODITE)59. Arrows show wind direction of the SW monsoon and NE monsoon with average δ18Op values of JJA and SON precipitation. The orange square denotes the study site location. The average seasonal cycle of (b) rainfall from APHRODITE (dark blue) and iTRACE Pre-industrial simulations (light blue) and (c) δ18Op measured from rainfall collected at Phong Nha-Ke Bang Park Headquarters (orange) and from iTRACE Pre-industrial simulations (red). APHRODITE and iTRACE time series taken from grid cell encompassing the study site.
PCP-corrected δ18O
In this study, we revisit a δ18O record (~ 50 yr resolution) from a 3.7 m long previously broken stalagmite (HH-1) collected from Hoa Huong cave in Phong Nha-Ke Bang National Park in central Vietnam (Supplementary Figs. S1, S2; 17.5°N, 106.2°E). HH-1 grew continuously from 45 – 4 ka, and the Mg/Ca and δ13C records were previously used to investigate local hydroclimate change in central Vietnam18. A strong correlation between Mg/Ca and δ13C indicated an important role of local water balance (precipitation minus evapotranspiration) in driving speleothem geochemistry through the PCP mechanism. Notably, the authors also found that HH-1 δ18O strongly correlated with both δ13C (r = 0.79, p < 0.01) and Mg/Ca (r = 0.87, p < 0.01) records (Fig. 3), suggesting that a common mechanism drove variability in all three proxies. This result was unexpected because δ18Op variability in central Vietnam does not reflect local rainfall amount, but rather regional processes such as upstream rainout and seasonality21. From this observation, the authors hypothesized that PCP, the primary control on the HH-1 δ13C and Mg/Ca records, likely influenced the HH-1 δ18O record. We test this hypothesis by using the HH-1 Mg/Ca record, cave monitoring data, and a geochemical model to remove the PCP signal from the HH-1 δ18O record. The corrected δ18O record (δ18Ocorr) should reflect δ18O changes driven by regional hydroclimate change (δ18Op) rather than local water balance (PCP), giving a history of central Vietnam δ18Op variability from 45 – 4 ka.
Fig. 3. Stalagmite (HH-1) proxies from 45 – 4 ka.
a HH-1 δ18O (orange) and corrected δ18O (δ18Ocorr, black). The black line is the correction using the best choice parameter configuration (see Supplementary Table S1) and αcalcite-water from ref. 53 (see “Methods”). The gray shading denotes the standard deviation of the 1000 Monte Carlo simulations from the sensitivity testing. b HH-1 Mg/Ca (red) and δ13C (pink). The black dashed line shows the Mg/Ca threshold used to determine Mg/Cai. Beige shading denotes the Younger Dryas and Heinrich Stadials 1–4. The Mg/Ca, δ13C, and δ18O curves are from ref. 18.
Portions of the HH-1 δ18O and δ18Ocorr records are notably different (Fig. 2a and Supplementary Fig. S2), with corrections reaching > 3–4 ‰ in places. Due to uncertainties introduced by choices we make in the correction method, including values of the model parameters, choice of temperature-dependent fractionation factor (αcalcite-water), and assumptions about the cave environment (see “Methods”), we focus our analyses on large changes (> 1 ‰) in the HH-1 δ18Ocorr record. The original δ18O record ranges from − 3 to − 10 ‰, whereas the δ18Ocorr record has a smaller range of − 6 to − 10 ‰. Because the correction removes the PCP signal, the largest changes occur during periods of high PCP (high δ18O, δ13C, and Mg/Ca values), while the smallest changes occur during periods of low PCP (low δ18O, δ13C, and Mg/Ca values). Generally, PCP is highest during the sea-level lowstand of the last glacial period and lowest during the sea-level highstands of the Holocene and late MIS 3. Low sea level exposes land adjacent to the study site (Gulf and Tonkin and South China Shelf), which reduces autumn moisture delivery to central Vietnam18. Ultimately, this results in smaller corrections during 45 – 35 ka and 13 – 4 ka, and larger corrections during 35 – 30 ka and 25 – 14 ka. The largest exceptions to this sea-level mechanism are two low PCP excursions from 30 – 25 ka and at 16 ka.
Notably, several features that the original δ18O record shares with the HH-1 Mg/Ca and δ13C records are no longer apparent. This includes the increasing trend starting at 35 ka and the abrupt shift to lower values at 14 ka. In the Mg/Ca and δ13C records, both of these features are interpreted to reflect changes in local rainfall amount driven by sea level change18. The correction also removes the anomalous negative excursions from 30 – 25 ka and at 16 ka. Features preserved in the δ18Ocorr record include positive excursions during the Younger Dryas and Heinrich Stadials 1, 3, and 4. Overall, the δ18Ocorr record no longer shows a relationship with sea level change at our site, but the δ18Ocorr record remains sensitive to changes in Atlantic Meridional Overturning Circulation (Younger Dryas and Heinrich Stadials). These multi-proxy results now reveal the important differences between local hydrology (Mg/Ca and δ13C) and large-scale hydroclimate (δ18Ocorr) that were not evident in the original uncorrected data18.
To test whether HH-1 δ18Ocorr tracks the changes in δ18Op, we compare the HH-1 δ18Ocorr record to regional speleothem δ18O records and isotope-enabled climate model output. Since there are no long speleothem (> 8000 years) δ18O records from Mainland Southeast Asia, we compared HH-1 δ18Ocorr to a composite δ18O stalagmite record from China, which reflects large-scale Asian summer monsoon strength22. Even though these records are somewhat distant from each other, the controls on summer monsoon δ18Op are similar (upstream rainout of moisture sourced from the Bay of Bengal and the Indian Ocean)19–21, which suggests speleothem δ18O records may covary across the greater region. The China composite shows far better agreement with the HH-1 δ18Ocorr record than the original HH-1 δ18O record (Fig. 4 and Supplementary Table S2). They have similar values and are strongly in phase from 45 – 11 ka. Both show positive excursions during Heinrich Stadials and the Younger Dryas, and negative excursions during some Dansgaard Oeschger Events. Since δ18Op is the primary driver of the China composite δ18O record, its similarities to the HH-1 δ18Ocorr indicate that HH-1 δ18Ocorr also reflects δ18Op variability. Interestingly, these records are antiphased during the early- and mid-Holocene. Since the HH-1 record ends at ~ 4 ka, we are unable to determine whether this antiphasing continues during the late Holocene.
Fig. 4. HH-1 compared to Chinese stalagmites.
a Time series of HH-1 δ18O (orange), HH-1 δ18Ocorr (black), and China composite δ18O (blue)22. Scatter plots of China composite δ18O vs. b HH-1 δ18O and c HH-1 δ18Ocorr. The Pearson correlation coefficient (r) and Spearman’s rank correlation coefficient (ρ) are displayed in the bottom right corner of panels (b) and (c). For the correlation coefficient values, the China composite record was interpolated to the time steps of the HH-1 record, and Holocene values (gray circles) were excluded. The p-values for all coefficients are < 0.01. See Supplementary Table S2 for r values using δ18Ocorr derived with different αcalcite-water.
Given the lack of Late Pleistocene terrestrial δ18O records from Vietnam, we compare the HH-1 δ18O record to simulated central Vietnam δ18Op from iTRACE23, an isotope-enabled transient climate model spanning the deglaciation, 20 – 11 ka (see “Methods”). Like the comparisons to regional δ18O records, the JJA (southwest monsoon), SON (northeast monsoon), and annual iTRACE δ18O time series from the grid cell encompassing the study site better correlate with the HH-1 δ18Ocorr record than the original δ18O record (Fig. 5 and Supplementary Table S2). When converted to δ18O of calcite (δ18Oc), the simulated iTRACE δ18Oc time series are more negative than the HH-1 δ18Ocorr time series (Fig. 5a). This difference is likely driven by the offset between iTRACE δ18Op and δ18Op measured from rainfall collected ~ 15 km from the study site (see “Methods”), where iTRACE δ18Op is ~ 1–2 ‰ more negative than observed from July-January (Fig. 2c). Notably, δ18Ocorr computed with a cave-derived αcalcite-water (see “Methods”), shows a smaller offset (Supplementary Fig. S3), suggesting the choice of αcalcite-water may be important when determining the absolute value of the correction. While the correlation with JJA is highest (Fig. 5e and Supplementary Table S2), the HH-1 δ18Ocorr time series better matches the δ18O magnitude of change in the annual and SON time series (Fig. 5f, g). Nevertheless, the strong agreement between these records suggests that iTRACE correctly simulates δ18Op variability in central Vietnam and provides further evidence that the HH-1 δ18Ocorr record reflects changes in δ18Op.
Fig. 5. HH-1 compared to iTRACE simulations.
a Time series of HH-1 δ18O (orange), HH-1 δ18Ocorr (black), iTRACE JJA δ18Oc (purple), iTRACE SON δ18Oc (light blue), and iTRACE annual δ18Oc (red). iTRACE δ18Oc values (VPDB) calculated from mean weighted iTRACE δ18Op (VSMOW) and temperature output using αcalcite-water from ref. 53. For visual clarity, iTRACE curves are smoothed with a 100-year running mean. Scatter plots of iTRACE δ18Op vs (b–d) HH-1 δ18O (orange) and (e–g) HH-1 δ18Ocorr (black). iTRACE output was smoothed with a 50-year running mean. The Pearson correlation coefficient (r) and Spearman’s rank correlation coefficient (ρ) are displayed in the bottom right corner of each panel. The p-values for all coefficients are < 0.01. See Supplementary Table S2 for r values using δ18Ocorr derived with different αcalcite-water.
Central Vietnam δ18Op variability over the last 45,000 years
We interpret HH-1 δ18Ocorr as a record of weighted annual mean δ18Op in central Vietnam. Understanding this record is challenging because both southwest monsoon (JJA) and northeast monsoon (SON) rainfall, specifically the relative contribution of each monsoon system, drive modern annual δ18Op variability in central Vietnam21. This means that HH-1 δ18Ocorr variability could reflect changes in JJA δ18Op, SON δ18Op, simultaneous changes in both seasons’ δ18Op, and/or changes in precipitation seasonality (e.g., the relative proportion of moisture delivered in autumn vs summer). For this reason, we consider both the southwest and northeast monsoons when interpreting the HH-1 δ18Ocorr record.
We first examine δ18Op variability during the deglaciation using iTRACE simulations. The similarity between the simulated JJA and SON δ18Op time series and their strong correlation with the HH-1 δ18Ocorr record makes disentangling the JJA and SON signals difficult (Fig. 5). A previous analysis of iTRACE simulations found that changes in seasonality (the amount of summer vs. autumn rainfall) minimally contributes to annual δ18Op change across South China and Southeast Asia24. Instead, they show that changes in the isotopic composition of JJA rainfall control annual δ18Op variability during the deglaciation, with SON δ18Op playing a minor role. However, iTRACE simulations underestimate the SON rainfall amount in central Vietnam (Fig. 2b), suggesting that the model may not capture the full influence of SON rainfall on the annual δ18Op signal. Ultimately, the HH-1 δ18Ocorr record most closely approximates the values of the annual δ18Op simulation (Fig. 5, Supplementary Fig. S3 and Supplementary Table S2), suggesting that our stalagmite likely records the combined δ18Op signal of both monsoon seasons.
When we consider the forcings individually, iTRACE δ18Op simulations reveal that meltwater forcing drives the majority of δ18Op variability during the deglaciation (MWF curve in Supplementary Fig. S4). This includes positive excursions during the HS1 and YD meltwater events, features also present in HH-1. In addition to HS1 and the YD, the HH-1 δ18Ocorr record shows positive excursions during Heinrich Stadials 3 and 4, indicating central Vietnam δ18Op responds somewhat consistently to meltwater events. The absence of HS2 in the HH-1 record suggests that HS2 is potentially weaker compared to other instances, which aligns with observations made in other paleoclimate records25,26. The signature of meltwater events on the Asian summer monsoon is well documented in speleothem δ18O records22,27. During meltwater events, the southward shift of the ITCZ and westerly jet weakens moisture transport28, which reduces upstream rainout and increases δ18Op values across the region19. Conversely, the meltwater-driven mechanism on δ18Op during SON is currently unknown.
Deciphering the drivers of Holocene δ18Op variability from the HH-1 δ18Ocorr record is challenging. During the Holocene, speleothem δ18O records of Asian summer monsoon intensity reach their minimum δ18O values in the early Holocene (Fig. 4), lagging Northern Hemisphere Summer Insolation by ~ 3 kyrs22,29. A recent synthesis of Holocene (8 – 0 ka) Southeast Asian speleothem δ18O records found the same relationship between the Southeast Asian summer monsoon (southwest monsoon) and summer insolation (PC1 from ref. 17). As for the Southeast Asian northeast monsoon (PC2 from ref. 17), the same study shows that speleothem δ18O (including a record from central Vietnam) tracks autumn insolation, which peaks in the mid-Holocene. Interestingly, the HH-1 δ18Ocorr record resembles neither the southwest nor northeast monsoon but rather has an increasing trend from 11 – 7 ka and then becomes relatively stable from 7 – 4 ka (Supplementary Fig. S5). Notably, the PCP correction is minimal during the Holocene (Fig. 3a), so the uncertainty introduced from the correction method does not explain these differences. It is possible that the correction method underestimates PCP during the Holocene, which would cause δ18Ocorr to be too high. While we cannot rule this out, the comparatively low and stable values of the Mg/Ca record indicate that PCP was low during the Holocene, making a large PCP-driven shift in speleothem δ18O unlikely. In fact, a recent modeling study found that rate-dependent fractionation when the PCP duration is small may decrease calcite δ18O by up to 0.5 ‰30. If this is indeed the case during the Holocene, then PCP would reduce the difference between HH-1 and regional records rather than increase it. Thus, this effect also cannot explain our observed differences. The δ18Ocorr record suggests that insolation does not directly drive changes in the HH-1 record δ18O during the Holocene. Instead, changes in seasonality (JJA versus SON precipitation amount), and changes in the isotopic composition of SON rainfall could both explain HH-1 δ18O variability during the Holocene (see Supplementary Text).
In summary, during the Late Pleistocene, meltwater-forced changes in δ18Op drive central Vietnam δ18Op variability. With the HH-1 δ18Ocorr record deviating from other regional δ18O records during the Holocene, our understanding of Holocene δ18Op variability in central Vietnam is less clear. Moving forward, more δ18O records from this region, particularly of the Holocene, and more in-depth investigations into autumn monsoon δ18Op dynamics are essential.
Implications
Our study emphasizes the potential effects of PCP on a speleothem δ18O record, which provides insights into our understanding of hydroclimate variability in central Vietnam δ18Op from 45 – 4 ka. Without correcting for the effects of PCP, we would have severely misinterpreted the HH-1 δ18O record. Furthermore, this would have led us to misdiagnose model-proxy disagreement with iTRACE, despite the model performing well. This has important implications, as speleothem δ18O records are increasingly being used to test paleoclimate model performance31–33. While some uncertainties in speleothem δ18O records are inevitable, it remains essential to address and minimize them to continue improving the utility of speleothem δ18O records.
Although PCP only affects speleothem δ18O under certain conditions (when PCP occurs on the cave ceiling or stalactite), future studies should screen for PCP-influenced δ18O when generating new speleothem records. This could be done through the generation of multiproxy stable isotope and trace elements records from the same speleothem. How widespread the effects of PCP on speleothem δ18O are is currently unknown; however, PCP is a well-documented driver of many existing δ13C and trace element records. Generating and interpreting speleothem δ13C and trace element records and other proxies sensitive to PCP (e.g., Ca isotopes), alongside δ18O records, could identify whether PCP occurs in drip waters feeding the speleothem sample. As shown in this study, strong similarities between speleothem δ18O and these proxies are an indicator of PCP-influenced δ18O. Replication with speleothem δ18O records from the same or nearby caves is another effective screening method since individual drip pathways would not likely be subject to identical PCP histories. If multiple δ18O records show similar variability, it is unlikely that PCP substantially affects speleothem δ18O.
Furthermore, we show that by utilizing Mg/Ca data and knowledge of the cave environment, it is possible to remove the PCP signal from a speleothem δ18O record. This method builds upon well-established geochemical relationships and the strong similarity of corrected data with existing records, suggesting it can be reliably used to constrain δ18Op. However, as detailed in the Methods, our model depends on cave monitoring measurements and makes several critical assumptions that may limit its applicability at certain cave sites. While not hugely impactful on the final correction, measurements of cave temperature, epikarst pCO2, cave pCO2, drip water δ18O, and the drip interval from an overlying or nearby dripsite are important for constraining the values of the corresponding model parameters. Nevertheless, when cave monitoring data is unavailable or difficult to measure (e.g., epikarst pCO2), we show that values sourced from the literature and climate model simulations may be suitable substitutes. Although not done in this study, measuring additional proxies, such as fluid inclusion δ18O34,35, fluid inclusion microthermometry36, and TEX8637, would provide estimates of how cave temperature and drip water δ18O change through time. Of all the parameters in the model, the correction is most sensitive to the initial Mg/Ca value of dripwater (Mg/Cai). Since we approximated Mg/Cai from a low Mg/Ca threshold in the speleothem Mg/Ca time series, Mg/Cai estimated from speleothem records without a clear low-end member may add substantial error to the correction. The choice of αcalcite-water also impacts the correction. In future applications of this method, we recommend performing a sensitivity analysis that includes changing the model parameters and αcalcite-water.
The model assumes that PCP is the only driver of speleothem Mg/Ca variability and that all PCP occurs on the cave ceiling. For these reasons, our method may not be suitable for cave sites where there are multiple controls on speleothem Mg/Ca or where the majority of PCP occurs in the epikarst. Ultimately, further model development and testing in well-monitored caves is needed before widespread application to other PCP-affected speleothem records. Nonetheless, our findings confirm that PCP on cave ceilings can substantially obscure the δ18Op signal, potentially leading to the misinterpretation of the widely-used speleothem δ18O proxy.
Methods
Rainfall δ18O measurements
A total of 83 rainwater samples were collected from January 2019 - August 2022 at the Phong Nha-Ke Bang National Park headquarters (17.6°N, 106.3°E) using collection methods outlined in ref. 26. We analyzed all samples for δ 18O (VSMOW): 50 at Chapman University (Orange, California) using cavity ring-down spectroscopy (Picarro L2130-i), and 33 at Northumbria University (Newcastle, UK) using off-axis integrated cavity output spectroscopy (Los Gatos Research LWIA-24EP) (Table S3). The long-term standard deviation of an independent quality control standard for each instrument is 0.11 ‰ δ 18O (Picarro) and 0.20 ‰ δ 18O (Los Gatos).
iTRACE
iTRACE is a water isotope-enabled transient simulation of the last deglaciation (20 to 11 ka) performed in iCESM1.323,38. iCESM1.3 is comprised of the Community Atmosphere Model version 1.3 (CAM5.3), the Community Land Model version 4 (CLM4), Parallel Ocean Program version 2 (POP2), and Los Alamos Sea Ice Model version 4 (CICE4). The land and atmosphere resolution is 1.9° x 2.5° (latitude and longitude), with 30 vertical levels in the atmosphere. iTRACE uses four forcing factors: ICE: continental ice sheets that are changed every 1000 years following the ICE-6G model39 and KMT ocean bathymetry (changed at 14 and 12 ka), ORB: orbital forcing, GHG: greenhouse gas concentrations from ice core reconstructions40–42, and MWF: meltwater fluxes following TRACE-21ka43. These forcings are applied additively to create four parallel simulations (ICE, ICE + ORB, ICE + ORB + GHG, and ICE + ORB + GHG + MWF). The effect of a single forcing can be approximated by subtracting simulations from each other (e.g., ORB = (ICE + ORB) – ICE). Refer to ref. 23 for additional details on the experiment setup and the forcings.
Preindustrial iCESM 1.3 simulations largely capture seasonal changes in rainfall and δ18Op in central Vietnam18. However, iCESM underestimates autumn rainfall amount by 2–5 mm/day and δ18Op by 1–2 ‰ for July-January (Fig. 2b, c). Strong agreement between simulated iTRACE precipitation and the HH-1 Mg/Ca and δ13C records indicates that iTRACE correctly models centennial-millennial-scale hydroclimate change in central Vietnam across the deglaciation18.
δ18O PCP correction
Correction procedure and model description
We corrected the HH-1 δ18O record for PCP using the HH-1 Mg/Ca time series and following methods from ref. 17. We used a Rayleigh distillation model that simulates δ18O, the evolution of HCO3−, and calcite precipitation (represented by the evolution of calcium concentration in the drip water) through time8,44–46. This model has three sinks (H2O, CO2 degassing, and CaCO3 precipitation), and can account for oxygen isotope exchange between drip water HCO3− and the water reservoir (H2O exchange). The evolution of calcium concentration in the drip water is approximated by exponential decay47.
Next, we outline the general procedure of the PCP correction method.
1. Estimate PCP duration. We first estimated PCP duration, defined as the length of calcite precipitation (in seconds) in the drip water prior to it reaching the speleothem, for each time point in the Mg/Ca time series with a corresponding δ18O measurement (n = 768, Supplementary Fig. S6a). To do this, we modeled the evolution of calcite precipitation over time, which requires estimates of initial Ca2+ concentration (Cai) and equilibrium Ca2+ concentration (Caeq) of the drip water, which are determined by the pCO2 in the soil or karst for Cai and the pCO2 in the cave for Caeq. For Caeq, we used measurements of the cave air pCO2 from Hoa Huong cave (see Model Parameters section). For Cai, we chose the minimum value necessary to simulate the range of measured Mg/Ca values from HH-1. We started the simulation at an estimated initial Mg/Ca value (Mg/Cai, the Mg/Ca of calcite precipitated when the duration of calcite precipitation from the drip water is 0 s, i.e., PCP = 0) and simulated the Mg/Ca evolution over time using a Rayleigh distillation model, with a partition coefficient calculated following ref. 48. We progressively increased Cai until the model simulated the Mg/Ca change necessary to explain the maximum Mg/Ca value of the HH-1 record (Supplementary Fig. S7). Thus, the modeled Mg/Ca evolution covers the whole range of Mg/Ca values measured from the HH-1 record and enables us to deduce the PCP duration for each Mg/Ca time point. We give further details on the Mg/Cai, Cai, and Caeq values, as well as the partition coefficient, in the Model Parameters section. In the Rayleigh model, longer PCP durations lead to increased Mg concentrations in the drip water and the subsequently precipitated calcite, resulting in larger deviations of the corresponding speleothem Mg/Ca from the Mg/Cai.
2. Calculate the PCP-induced change in δ18O. We then estimated the PCP-induced change in δ18O of drip water HCO3− for each timepoint using the Mg/Ca-based estimate of PCP duration. To do this, we modeled the oxygen isotope evolution of the HCO3− in the drip water and then determined the amplitude of δ18O change for each sample based on the estimated PCP duration. Similar to the increase in Mg/Ca over the course of calcite precipitation, longer PCP durations will correspond to larger increases in HCO3− δ18O in the Rayleigh model.
3. Calculate δ18Ocorr. The PCP-induced change in δ18O for each sample, as calculated in the previous step, was then subtracted from the corresponding measured speleothem δ18O value to produce the final δ18Ocorr time series (Fig. 3a and Supplementary Fig. S2). Ultimately, δ18Ocorr should approximate the initial δ18O value of calcite precipitated in equilibrium with the drip water (when PCP = 0) which is equivalent (in equilibrium with) to the drip water δ18O. In the case of stalagmite HH-1, we expect drip water δ18O to reflect δ18Op.
We make two important assumptions in this approach. First, we assume PCP is the only control on drip water Mg/Ca variability. Second, we assume all PCP occurs on the cave ceiling and/or overhanging stalactite. There may be other factors that substantially influence speleothem Mg/Ca (e.g., karst hydrology) and PCP may occur in both the epikarst and on the cave ceiling. We further explore these concepts and how they apply to the correction of the HH-1 δ18O record in the Model uncertainty and sensitivity section.
Model parameters
The model we used for the PCP correction requires a number of parameters as inputs (Mg/Cai, cave temperature, Cai, Caeq, drip interval, and drip water δ18O; Supplementary Table S1). When possible, we used direct measurements of the cave environment collected from March 2020 – March 2023 and further described in ref. 18. Otherwise, we used iTRACE model output and values from the literature. To assess the uncertainty of our correction, we also performed sensitivity testing by varying some of the model parameters within realistic ranges. Here, we describe the choices for each parameter.
We used a constant value of 2.23 mmol/mol for Mg/Cai (black dashed line in Fig. 3b). We derived this value by averaging two portions of the HH-1 Mg/Ca record (45 – 37 ka, and 13 – 4 ka). For these periods, we assume that little to no PCP occurred based on the comparatively low and relatively stable Mg/Ca values. For sensitivity testing, we varied Mg/Cai between 1.96 – 2.5 mmol/mol, based on the standard deviation of the record for the two time periods.
Since the temperature was not constant during the last 45,000 years, we approximated a temperature time series using a combination of cave monitoring data and iTRACE climate model output (Supplementary Fig. S8). We assumed a constant temperature of 21 °C during the Holocene (11 – 4 ka in the HH-1 record), which is an intermediate value between the modern annual temperature of Hoa Huong cave collected via data logger from March 2020 – August 2022 (~ 20 °C18) and Holocene (11.1 – 11 ka) iTRACE surface temperature values (~ 21.5 °C). From 20 – 11 ka, we varied temperature based on a 50-year running mean of simulated iTRACE surface temperature from the grid cell encompassing the study site. We used a constant temperature of 18.5 °C for the time period of 45 – 20 ka, based on the annual iTRACE simulated surface temperature from 20 ka. For sensitivity testing, we varied the temperature +/− 2 °C, the current seasonal range in Hoa Huong cave18.
The model also requires inputs of drip water Cai and Caeq. For our model to explain the full range of HH-1 Mg/Ca values, we had to use a minimum Cai value of 3.4 mol/m3, which is equivalent to a pCO2 of 49,000 ppm at 21 °C (we assume the same temperature for all the following pCO2 values). While this value exceeds the majority of soil pCO2 measurements, pCO2 levels in the epikarst can surpass those observed at the surface and in the soil. Notably, studies have recorded epikarst pCO2 values as high as 60,000 – 100,000 ppm49,50, and studies utilizing geochemical models to approximate epikarst pCO2 found similarly high levels51,52 We conducted sensitivity analyses using a range of Cai values between 3.4 and 4 mol/m3 (equivalent to 49,000 – 79,500 ppm) to test how higher epikarst pCO2 values would affect our results. For Caeq, we used a value of 0.8 mol/m, which corresponds to a pCO2 value of 635 ppm, an average of two modern pCO2 values measured in Hoa Huong Cave, 820 ppm (March 2020) and 510 ppm (March 2023). For the sensitivity testing, we used a range of 0.6–1.4 mmol/m3, which corresponds to pCO2 values of 270 – 3400 ppm. Our low estimate of 270 ppm reflects atmospheric conditions during the last glacial maximum41 and also assumes that the cave is fully ventilated. We based our highest estimate of 3400 ppm on the highest measured value of pCO2 recorded in August 2022 in Hoa Huong cave.
Drip interval affects speleothem Mg/Ca, and thus, also the PCP duration we estimate with our approach. Since there is no active drip at the collection site, we were unable to measure the drip interval and assume a constant drip interval of 1 drip/second for the method. Even though drip interval may have changed through time, previous work has found changes in drip interval had a smaller effect on the isotopic composition of calcite when compared to the effects of PCP8.
The model uses an infiltrating water parameter (drip water δ18O) as the initial HCO3− δ18O value. We used iTRACE simulations of weighted annual mean δ18Op to estimate the infiltrating water δ18O parameter (drip water δ18O). We assumed drip water values approximate the weighted annual mean of δ18Op because we have limited observations of drip water δ18O at the study site. From 20 – 11 ka, we varied temperature based on a 50-year running mean of simulated precipitation weighted annual mean δ18Op from the grid cell encompassing the study site. We used a constant δ18Op value of − 9.4 ‰ for the time period of 45 – 20 ka and − 9.5 ‰ for the time period 11 − 4 ka, based on the iTRACE weighted annual values from 20 ka and 11 ka, respectively. We did not perform sensitivity testing because when PCP duration is sufficiently short (Supplementary Fig. S6a), the effects of the drip water δ18O value are minimal8.
We used laboratory-based measurements to calculate the partition coefficient of Mg (D(Mg))48. Since D(Mg) is temperature-dependent, we calculated D(Mg) at each time step using the approximated cave temperature record (Fig S8). The model uses temperature-dependent equilibrium fractionation factors (see ref. 8).
Model uncertainty and sensitivity testing
For the sensitivity analyses, we performed 1000 Monte Carlo simulations and determined the standard deviation for each parameter. To assess the uncertainty of the full model, we randomly varied all parameters within the chosen ranges (Supplementary Table S1 and Supplementary Fig. S9a, gray shading on δ18Ocorr in Figs. 3–5). To assess the uncertainty of some parameters, we randomly varied a single parameter (Mg/Cai, cave temperature, Cai, and Caeq) while keeping the remaining variables constant (Supplementary Fig. S9b–e). For the full model run, the average 1 sigma uncertainty was +/− 0.21 ‰, a relatively small value when compared to the magnitude of variation in the HH-1 δ18Ocorr record. Our simulation also revealed that the Mg/Cai parameter causes nearly all of the variability in the simulations (Supplementary Fig. S9b). The variability of the remaining variables (temperature, Cai, and Caeq), had negligible effects on the results. This suggests an accurate estimate of Mg/Cai is crucial for future use of this correction method.
Given that we do not fully understand how the cave environment changes over time, we assume constant values for several model parameters: Mg/Cai, Cai, and Caeq. This assumption is likely incorrect, particularly between glacial and interglacial climate conditions. However, the insensitivity of the PCP correction to changes in Cai and Caeq suggests using constant values for these parameters is reasonable (Supplementary Fig. S9d, e). Conversely, the correction is very sensitive to changes in Mg/Cai. Our sensitivity testing accounts for some of this uncertainty, but this is currently a limitation that warrants further investigation. To avoid introducing additional uncertainty to the correction by choosing arbitrary changes in Mg/Cai, we use our best estimate of Mg/Cai for the entire record.
The sensitivity of the correction to the Mg/Cai parameter highlights the importance of the model assumption that PCP is the only driver of drip water Mg/Ca variability. As previously stated, the strong similarities between the HH-1 Mg/Ca and δ13C records indicate that PCP is the primary control on drip water Mg/Ca18 (Fig. 3). While we cannot entirely eliminate the influence of other controls on drip water Mg/Ca and whether these controls change through time, the relative stability of the low/no PCP periods in the HH-1 Mg/Ca record during 45 – 37 ka and 13 – 4 ka suggests that factors other than PCP have a negligible effect on Mg/Ca variability.
We test the difference of the PCP corrected δ18O record when using the fractionation factor (αcalcite-water) derived from laboratory experiments53 against three cave-derived fractionation factors9,54,55. The fractionation factor from ref. 53 results in the largest corrections, and the ref. 9 fractionation factor the smallest (Supplementary Fig. S2), and the maximum difference between the resulting δ18Ocorr time series is ~ 1 ‰ (Supplementary Fig. S10). Since the different corrections largely do not change the outcome of our interpretations (Supplementary Table S2), we choose to focus our discussion primarily on the PCP correction obtained using the laboratory-derived fractionation factor53.
We also found that using a version of the model that includes H2O exchange has a small impact on the correction (Supplementary Fig. S6b). This is because PCP duration for most data points is substantially shorter than the time required for complete H2O exchange (Supplementary Fig. S6a). For example, only 6 of the 768-time points exceed a PCP duration of 1000 s (orange dashed line in Supplementary Fig. S6a), where H2O exchange only reduces PCP-induced δ18O change by ~ 0.35 ‰ (orange dashed line in Supplementary Fig. S6c). In fact, for the majority of the record, the duration of PCP is sufficiently short to render the effects of H2O exchange negligible.
Importantly, our approach only accounts for PCP occurring immediately prior to speleothem precipitation (i.e., PCP on the cave ceiling or stalactite) and excludes PCP in the epikarst. Since PCP in the epikarst should not affect δ18O4, we expect the PCP impact on the speleothem δ18O record to originate from calcite precipitation on the cave ceiling or overlying stalactite. Conversely, PCP in the epikarst does affect drip water Mg/Ca as there is no exchange process that counterbalances PCP for Mg/Ca. In cases where PCP occurs in the epikarst, we would overestimate PCP-induced increase in speleothem δ18O and bias δ18O values of the corrected speleothem δ18O record toward low values. Our correction approach cannot currently disentangle epikarst PCP from in-cave PCP. While there is no stalactite or active drip above HH-1, we do see evidence of a previously connected soda straw on top of HH-1 and some secondary calcite deposits on the cave ceiling, indicating that some form of calcite precipitation likely occurred above the growing stalagmite (Supplementary Fig. S1b). To test the assumption that all PCP occurs on the cave ceiling or in stalactites, we compare our corrected δ18O record against other regional records less likely to be affected by PCP and climate model output.
Supplementary information
Acknowledgements
We thank the staff and the management board of Phong Nha-Ke Bang National Park, including Mr. D. Nguyen, Mr. H.Q. Nguyen, and Mr. T.V. Vo, for assistance with this project. We also thank D. Limbert and H. Limbert of Oxalis Adventure Tours for providing a map of Hoa Huong Cave and C. He of the University of Miami for providing access to the preindustrial simulations of iTRACE. This work was supported by US NSF P2C2 awards # 2103129/1603056 (K.R.J.), # 2103051/1602947 (M.L.G.), # 2102976 (D.M.), and the Ridge to Reef NSF Research Traineeship award # DGE- 1735040 (E.W.P.).
Author contributions
Fieldwork: E.W.P., A.W., K.R.J., M.L.G., T.N.B., M.X.T., T.H.D., Q.D., and V.E. Water isotope analyses: V.E., G.R.G. δ18O correction model: V.S. Climate model analyses: E.W.P. Visualization: E.W.P., A.W., and V.S. Supervision: K.R.J., M.L.G., and D.M. Writing—original draft: E.W.P., V.S., and A.W. Writing—review & editing: E.W.P., V.S., A.W., M.L.G., D.M., T.N.B., M.X.T., T.H.D., Q.D., G.R.G., V.E., and K.R.J.
Peer review
Peer review information
Nature Communications thanks Valdir Novello and the other anonymous reviewers for their contribution to the peer review of this work. A peer review file is available.
Data availability
All corrected δ18O time series and rainwater δ18O data from Phong Nha-Ke Bang National Park Headquarters are archived at Zenodo (10.5281/zenodo.13768949)56. The corrected δ18O time series and the original HH-1 δ18O, δ13C, and Mg/Ca time series can be found at the National Oceanic and Atmospheric Administration National Centers for Environmental Information Paleoclimatology archive (10.25921/643e-eb71 and 10.5281/zenodo.13768949)57,58.
Code availability
The code used for the PCP removal method, including instructions and example data, is accessible via Zenodo (10.5281/zenodo.13768862)56.
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Contributor Information
E. W. Patterson, Email: pattersone11@wpunj.edu
V. Skiba, Email: vanessa.skiba@awi.de
Supplementary information
The online version contains supplementary material available at 10.1038/s41467-024-53422-y.
References
- 1.Lachniet, M. S. Climatic and environmental controls on speleothem oxygen-isotope values. Quat. Sci. Rev.28, 412–432 (2009). [Google Scholar]
- 2.Baker, A. & Bradley, C. Modern stalagmite δ18O: Instrumental calibration and forward modelling. Glob. Planet. Change71, 201–206 (2010). [Google Scholar]
- 3.Treble, P. C. et al. Ubiquitous karst hydrological control on speleothem oxygen isotope variability in a global study. Commun. Earth Environ.3, 1–10 (2022). [Google Scholar]
- 4.Deininger, M. et al. Are oxygen isotope fractionation factors between calcite and water derived from speleothems systematically biased due to prior calcite precipitation (PCP)? Geochim. Cosmochim. Acta305, 212–227 (2021). [Google Scholar]
- 5.Hansen, M., Scholz, D., Schöne, B. R. & Spötl, C. Simulating speleothem growth in the laboratory: Determination of the stable isotope fractionation (δ13C and δ18O) between H2O, DIC and CaCO3. Chem. Geol.509, 20–44 (2019). [Google Scholar]
- 6.Mickler, P. J., Stern, L. A. & Banner, J. L. Large kinetic isotope effects in modern speleothems. GSA Bull.118, 65–81 (2006). [Google Scholar]
- 7.Skiba, V. et al. Millennial-scale climate variability in the Northern Hemisphere influenced glacier dynamics in the Alps around 250,000 years ago. Commun. Earth Environ.4, 1–10 (2023).37325084 [Google Scholar]
- 8.Skiba, V. & Fohlmeister, J. Contemporaneously growing speleothems and their value to decipher in-cave processes – A modelling approach. Geochim. Cosmochim. Acta348, 381–396 (2023). [Google Scholar]
- 9.Daëron, M. et al. Most Earth-surface calcites precipitate out of isotopic equilibrium. Nat. Commun.10, 429 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Kaushal, N. et al. SISALv3: A global speleothem stable isotope and trace element database. Earth Syst. Sci. Data16, 1933–1963 (2023).
- 11.Fairchild, I. J. et al. Modification and preservation of environmental signals in speleothems. Earth Sci. Rev.75, 105–153 (2006). [Google Scholar]
- 12.Johnson, K. R., Hu, C., Belshaw, N. S. & Henderson, G. M. Seasonal trace-element and stable-isotope variations in a Chinese speleothem: The potential for high-resolution paleomonsoon reconstruction. Earth Planet. Sci. Lett.244, 394–407 (2006). [Google Scholar]
- 13.Polag, D. et al. Stable isotope fractionation in speleothems: Laboratory experiments. Chem. Geol.279, 31–39 (2010). [Google Scholar]
- 14.Fairchild, I. J. & Treble, P. C. Trace elements in speleothems as recorders of environmental change. Quat. Sci. Rev.28, 449–468 (2009). [Google Scholar]
- 15.Hendy, C. H. The isotopic geochemistry of speleothems—I. The calculation of the effects of different modes of formation on the isotopic composition of speleothems and their applicability as palaeoclimatic indicators. Geochim. Cosmochim. Acta35, 801–824 (1971). [Google Scholar]
- 16.Beck, W. C., Grossman, E. L. & Morse, J. W. Experimental studies of oxygen isotope fractionation in the carbonic acid system at 15°, 25°, and 40 °C. Geochim. Cosmochim. Acta69, 3493–3503 (2005). [Google Scholar]
- 17.Wolf, A. et al. Deciphering local and regional hydroclimate resolves contradicting evidence on the Asian monsoon evolution. Nat. Commun.14, 5697 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Patterson, E. W. et al. Glacial changes in sea level modulated millennial-scale variability of Southeast Asian autumn monsoon rainfall. Proc. Natl. Acad. Sci. USA120, e2219489120 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Pausata, F. S. R., Battisti, D. S., Nisancioglu, K. H. & Bitz, C. M. Chinese stalagmite δ18O controlled by changes in the Indian monsoon during a simulated Heinrich event. Nat. Geosci.4, 474 (2011). [Google Scholar]
- 20.Yang, H., Johnson, K. R., Griffiths, M. L. & Yoshimura, K. Interannual controls on oxygen isotope variability in Asian monsoon precipitation and implications for paleoclimate reconstructions. J. Geophys. Res. D Atmos.121, 8410–8428 (2016). [Google Scholar]
- 21.Wolf, A., Roberts, W. H. G., Ersek, V., Johnson, K. R. & Griffiths, M. L. Rainwater isotopes in central Vietnam controlled by two oceanic moisture sources and rainout effects. Sci. Rep.10, 1–14 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Cheng, H. et al. The Asian monsoon over the past 640,000 years and ice age terminations. Nature534, 640–646 (2016). [DOI] [PubMed] [Google Scholar]
- 23.He, C. et al. Hydroclimate footprint of pan-Asian monsoon water isotope during the last deglaciation. Sci. Adv.7, 10.1126/sciadv.abe2611 (2021). [DOI] [PMC free article] [PubMed]
- 24.He, C. et al. Deglacial variability of South China hydroclimate heavily contributed by autumn rainfall. Nat. Commun.12, 1–9 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Lynch-Stieglitz, J. et al. Muted change in Atlantic overturning circulation over some glacial-aged Heinrich events. Nat. Geosci.7, 144–150 (2014). [Google Scholar]
- 26.Wright, K. T. et al. Dynamic and thermodynamic influences on precipitation in Northeast Mexico on orbital to millennial timescales. Nat. Commun.14, 2279 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Zhang, H. et al. East Asian hydroclimate modulated by the position of the westerlies during Termination I. Science362, 580–583 (2018). [DOI] [PubMed] [Google Scholar]
- 28.Zhang, R. & Delworth, T. L. Simulated tropical response to a substantial weakening of the Atlantic thermohaline circulation. J. Clim.18, 1853–1860 (2005). [Google Scholar]
- 29.Wang, Y. et al. Millennial- and orbital-scale changes in the East Asian monsoon over the past 224,000 years. Nature451, 1090–1093 (2008). [DOI] [PubMed] [Google Scholar]
- 30.Sade, Z., Hegyi, S., Hansen, M., Scholz, D. & Halevy, I. The effects of drip rate and geometry on the isotopic composition of speleothems: Evaluation with an advection-diffusion-reaction model. Geochim. Cosmochim. Acta317, 409–432 (2022). [Google Scholar]
- 31.Osman, M. B. et al. Globally resolved surface temperatures since the Last Glacial Maximum. Nature599, 239–244 (2021). [DOI] [PubMed] [Google Scholar]
- 32.Tierney, J. E. et al. Glacial cooling and climate sensitivity revisited. Nature584, 569–573 (2020). [DOI] [PubMed] [Google Scholar]
- 33.Ramos, R. D. et al. Constraining clouds and convective parameterizations in a climate model using paleoclimate data. J. Adv. Model. Earth Syst. 14, e2021MS002893 (2022).
- 34.Affolter, S. et al. Central Europe temperature constrained by speleothem fluid inclusion water isotopes over the past 14,000 years. Sci. Adv. 5, 10.1126/sciadv.aav3809 (2019). [DOI] [PMC free article] [PubMed]
- 35.Fleitmann, D., Burns, S. J., Neff, U., Mangini, A. & Matter, A. Changing moisture sources over the last 330,000 years in Northern Oman from fluid-inclusion evidence in speleothems. Quat. Res.60, 223–232 (2003). [Google Scholar]
- 36.Løland, M. H. et al. Evolution of tropical land temperature across the last glacial termination. Nat. Commun.13, 5158 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Baker, A. et al. Glycerol dialkyl glycerol tetraethers (GDGT) distributions from soil to cave: Refining the speleothem paleothermometer. Org. Geochem.136, 103890 (2019). [Google Scholar]
- 38.Brady, E. et al. The connected isotopic water cycle in the community earth system model version 1. J. Adv. Model. Earth Syst.11, 2547–2566 (2019). [Google Scholar]
- 39.Peltier, W. R., Argus, D. F. & Drummond, R. Space geodesy constrains ice age terminal deglaciation: The global ICE-6G_C (VM5a) model. J. Geophys. Res. Solid Earth120, 450–487 (2015). [Google Scholar]
- 40.Lüthi, D. et al. High-resolution carbon dioxide concentration record 650,000–800,000 years before present. Nature453, 379–382 (2008). [DOI] [PubMed] [Google Scholar]
- 41.Petit, J. R. et al. Climate and atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica. Nature399, 429–436 (1999). [Google Scholar]
- 42.Schilt, A. et al. Glacial–interglacial and millennial-scale variations in the atmospheric nitrous oxide concentration during the last 800,000 years. Quat. Sci. Rev.29, 182–192 (2010). [Google Scholar]
- 43.Liu, Z. et al. Transient simulation of last deglaciation with a new mechanism for bølling-allerød warming. Science325, 310–314 (2009). [DOI] [PubMed] [Google Scholar]
- 44.Deininger, M., Fohlmeister, J., Scholz, D. & Mangini, A. Isotope disequilibrium effects: The influence of evaporation and ventilation effects on the carbon and oxygen isotope composition of speleothems–A model approach. Geochim. Cosmochim. Acta96, 57–79 (2012). [Google Scholar]
- 45.Deininger, M. & Scholz, D. ISOLUTION 1.0: an ISOtope evoLUTION model describing the stable oxygen (δ18O) and carbon (δ13C) isotope values of speleothems. Int. J. Speleol.48, 21–32 (2019). [Google Scholar]
- 46.Scholz, D., Mühlinghaus, C. & Mangini, A. Modelling δ13C and δ18O in the solution layer on stalagmite surfaces. Geochim. Cosmochim. Acta73, 2592–2602 (2009). [Google Scholar]
- 47.Dreybrodt, W. Deposition of calcite from thin films of natural calcareous solutions and the growth of speleothems. Chem. Geol.29, 89–105 (1980). [Google Scholar]
- 48.Day, C. C. & Henderson, G. M. Controls on trace-element partitioning in cave-analogue calcite. Geochim. Cosmochim. Acta120, 612–627 (2013). [Google Scholar]
- 49.Benavente, J. et al. Air carbon dioxide contents in the vadose zone of a Mediterranean Karst. Vadose Zone J.9, 126 (2010). [Google Scholar]
- 50.Peyraube, N., Lastennet, R., Denis, A. & Malaurent, P. Estimation of epikarst air PCO2 using measurements of water δ13CTDIC, cave air PCO2 and δ13CCO2. Geochim. Cosmochim. Acta118, 1–17 (2013). [Google Scholar]
- 51.Baker, A., Flemons, I., Andersen, M. S., Coleborn, K. & Treble, P. C. What determines the calcium concentration of speleothem-forming drip waters? Glob. Planet. Change143, 152–161 (2016). [Google Scholar]
- 52.Treble, P. C. et al. Impacts of cave air ventilation and in-cave prior calcite precipitation on Golgotha Cave dripwater chemistry, southwest Australia. Quat. Sci. Rev.127, 61–72 (2015). [Google Scholar]
- 53.O’Neil, J. R., Clayton, R. N. & Mayeda, T. K. Oxygen isotope fractionation in divalent metal carbonates. J. Chem. Phys.51, 5547–5558 (1969). [Google Scholar]
- 54.Tremaine, D. M., Froelich, P. N. & Wang, Y. Speleothem calcite farmed in situ: Modern calibration of δ18O and δ13C paleoclimate proxies in a continuously-monitored natural cave system. Geochim. Cosmochim. Acta75, 4929–4950 (2011). [Google Scholar]
- 55.Johnston, V., Borsato, A., Spötl, C., Frisia, S. & Miorandi, R. Stable isotopes in caves over altitudinal gradients: fractionation behaviour and inferences for speleothem sensitivity to climate change. Clim. Past9, 99–118 (2013). [Google Scholar]
- 56.Patterson, E. W. et al. Data and code: Local hydroclimate alters interpretation of speleothem δ18O records. Zenodo. 10.5281/zenodo.13768949 (2024). [DOI] [PMC free article] [PubMed]
- 57.Patterson, E. W. et al. Hoa Huong Cave, Vietnam PCP-Corrected d18O Data from 4-45 ka. National Oceanic and Atmospheric Administration National Centers for Environmental Information Paleoclimatology archive. 10.25921/643e-eb71(2024).
- 58.Patterson, E. W. et al. Hoa Huong Cave, central Vietnam 4- 45 ka d18O, d13C, and Mg/Ca stalagmite data. National Oceanic and Atmospheric Administration National Centers for Environmental Information Paleoclimatology archive. 10.25921/axna-4s49 (2023).
- 59.Yatagai, A. et al. APHRODITE: Constructing a long-term daily gridded precipitation dataset for Asia based on a dense network of rain gauges. Bull. Am. Meteorol. Soc.93, 1401–1415 (2012). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
All corrected δ18O time series and rainwater δ18O data from Phong Nha-Ke Bang National Park Headquarters are archived at Zenodo (10.5281/zenodo.13768949)56. The corrected δ18O time series and the original HH-1 δ18O, δ13C, and Mg/Ca time series can be found at the National Oceanic and Atmospheric Administration National Centers for Environmental Information Paleoclimatology archive (10.25921/643e-eb71 and 10.5281/zenodo.13768949)57,58.
The code used for the PCP removal method, including instructions and example data, is accessible via Zenodo (10.5281/zenodo.13768862)56.