Introduction

Caves and speleothems are ubiquitous features in (carbonate-bearing) mountains worldwide, which often harbour glaciers at high elevations and/or during cold climate periods. Mountain glaciers are rapidly vanishing1 and have been classified as a regional impact tipping element2, underscoring the need to improve our understanding of their dynamics on short and long timescales.

Speleothems from the northern rim of the European Alps offer a valuable archive to investigate Northern Hemisphere glacial-interglacial climate variability and provide crucial insights into abrupt climate changes on millennial timescales3,4,5,6,7,8,9,10,11,12,13,14,15,16. The Northern Alps are strongly influenced by the North Atlantic10,17, a region which is thought to dictate global-scale abrupt climate events during glacial times18,19,20,21,22,23. Speleothems from Alpine caves have been shown to mimic millennial-scale variability and abruptness of climate transitions known from Greenland ice cores4,7,9,12,13,16. The large amplitude δ18O signals captured in speleothems from the Northern Alps (in contrast to speleothems from low-elevation sites outside the Alps24,25) have been attributed largely to changes in regional air temperature7,13, but the detailed processes controlling these isotopic changes remain to be fully understood.

Here, we aim to resolve these processes controlling the millennial-scale isotopic changes in the high-elevation Northern Alps on interstadial-stadial (IS-S) timescales. For this, we produced a high resolution, precisely U-Th-dated speleothem multi-proxy record (δ18O, δ13C, Mg/Ca and Sr/Ca) for the time span ~300 to 206 ka, i.e. covering the end of Marine Isotope Stage (MIS) 9, MIS 8, and the majority of MIS 7, from the northern rim of the Alps in central Switzerland. Speleothem records from this cave site (Melchsee-Frutt) have been found to record interstadial-stadial (IS-S) variability16 and we identify periods of pronounced IS-S variability during times of intermediate relative sea level. We use speleothem δ13C data to identify periods when the cave site was likely overlain by ice in order to provide constraints on the evolution of Alpine glaciers during glacial periods where essentially no other data are available. We propose that the speleothem C isotopic composition and the Mg/Ca ratio record changes in the amount of glacier surface meltwater vs. the amount of basal meltwater occurring on Northern Hemisphere IS-S timescales. Finally, we show that IS-S amplitudes obtained from our speleothem record, after accounting for cave-internal effects, are consistent with those derived from an isotope-enabled fully-coupled ocean-atmosphere model23 and likely reflect higher δ18O values of North Atlantic seawater during interstadials compared to stadials.

Results and discussion

Interstadial-stadial variability in the Northern Alps and other European records during MIS 9‒7

We present δ18O, δ13C and Mg/Ca composite records based on three stalagmites retrieved from two high-elevation caves in the Melchsee-Frutt region of central Switzerland (46°47′N, 8°16′E; ~2000 m a.s.l.; composites produced using the iscam age modelling algorithm26, see Supplementary Information and Figs. S1─S7 for details). We detect transitions to interstadial conditions in these records using an automatic algorithm (S-IS transitions, Fig. S8 in Supplementary Information). This algorithm is able to identify transitions based on their higher abruptness relative to other changes in the respective record. We compare our record and detected transitions to other European speleothem and sediment records (Fig. 1). The identified abrupt transitions exclusively occurred during periods of intermediate sea level (below about ‒30 m and above about ‒100 m relative to modern sea level, Fig. S9) as it was common for IS-S variability also during other glacial cycles. Periods of occurrence and absence of abrupt S-IS transitions in the presented speleothem δ18O record are corresponding to the ones from other European speleothem and sediment records (Fig. 1d–f). Additionally, these periods are also similar for more distant records including a synthetic Greenland ice-core record and a Chinese speleothem composite record (Fig. S10). Common periods of occurrence of S-IS transitions in these records are ~220–230, 240–255, 260–265 and 280–305 ka.

Fig. 1: Comparison of the new Alpine speleothem record with other records from Europe showing interglacial-glacial and millennial-scale (IS-S) variability.
figure 1

a Melchsee-Frutt δ18O record (grey and black are original and PCP-corrected records, see Methods and Supplementary Information for details on the PCP correction). b Melchsee-Frutt δ13C record (orange) and host rock δ13C value indicated by the orange dashed horizontal line together with a global relative sea level (RSL) reconstruction (96, black dashed line). c Melchsee-Frutt Mg/Ca record (purple). d Speleothem δ18O from Spannagel cave, Central Alps6,97, and Piani Eterni, Italy27. e Speleothem δ13C from Ejulve cave, Spain28 and Crovassa Azzurra, Italy29. f Lake Van sediment core reflectance30,31. g. Iberian Margin alkenone-\({U}_{37}^{{k}^{{\prime} }}\) reconstructed sea-surface temperature98. Stars mark the detected abrupt interstadial transitions (colours correspond to time series colours, see Supplementary Information for details on transition detection method). Vertical black dashed lines mark MIS substages as indicated by their number on top of the figure (at 204.1, 212, 216.1 and 234.397 for MIS 7, 243, 254, 269, 280, 291 and 309 ka99, for MIS 8 and 9).

It is important to note that abrupt climate transitions similar to glacial S-IS transitions occurred also during MIS 7d in the Northern Alps (Melchsee-Frutt speleothems, Fig. 1a–d), in southern Europe (Italy and Spain, recorded by speleothems from Piani Eterni27, Ejulve cave28 and Crovassa Azzurra29, Fig. 1d, e) and southeastern Europe (Lake Van30,31, Turkey, Fig. 1f). MIS 7d was a cold substage during an interglacial (MIS 7) and demonstrates that relatively short periods of global cooling are sufficient to enable trigger mechanisms for S-IS transitions provided that the conditions are favourable, i.e. relative sea level is at about −60 m for MIS 7d (Fig. 1).

Alpine speleothem δ18O and δ13C data records palaeoglacier dynamics

Melchsee-Frutt speleothem δ13C values fall into two populations separated by the mean host rock δ13C value of ~2.7 ‰ (Fig. 2). Carbon isotope values close to this value point to conditions where soil respiration was close to zero and hence soil organic matter was largely absent above the cave (cf.33). Speleothem formation requires that host rock is dissolved by acidified water, in order to enable carbonate precipitation later in the cave. Since we infer from speleothem δ13C that speleothems in this cave also grew during times of soil absence, a process other than soil respiration must have led to the acidification of the water. Previous studies suggested that oxidation of pyrite resulting in sulfuric acid can lead to host rock dissolution required for speleothem formation32,33,34,35,36,37. The study area was repeatedly glaciated during the last glacial period38,39,40,41,42,43,44 and was probably also covered by a glacier during earlier glacials. Thus, it is likely that during times when signs of biological activity were absent (speleothem δ13C higher than mean host rock δ13C), a glacier was present above the cave. Conversely, the group showing δ13C values lower than the host rock value indicates some input of soil-biogenic \({{p}}{{{\rm{CO}}}}_{2}\) (Fig. 2), suggestive of a climate when thin and patchy Alpine soils and sparse vegetation were present in the catchment, similar to today. Periods with soil coverage above the cave (δ13C ≤ 2.7 ‰) are characterised by δ18O values ranging from –9 to –8‰, while calcite precipitated during periods when no soil was present above the cave shows more depleted δ18O values (as low as –13.5‰; Fig. 2).

Fig. 2: Scatter plot of δ18O and δ13C values from Melchsee-Frutt stalagmites.
figure 2

Two groups can be separated by a δ13C threshold at ~2.7 ‰ (host rock value). The DBSCAN algorithm included in the Python package scikit-learn100 was used to cluster the data into groups (parameters: eps = 0.75, minimum number of points in a cluster = 330). Points are automatically assigned by the clustering algorithm where contiguous regions with high density of data points are classified as clusters and points which do not fall into such a region and do not form a contiguous region with high density of data points are not classified. Magenta stars indicate the data characterised by δ13C values lower than the threshold while the majority of data is characterised by δ13C values above the threshold (blue stars). Black stars mark extreme values which have not been assigned to one of the groups (74 samples).

We find very high δ13C values (up to +12‰, Fig. 2), comparable to extreme values previously found in other Alpine speleothems (e.g., up to +20‰ in Spötl & Mangini36, up to +7‰ in Spötl et al. 33). These anomalously high values cannot be explained solely by the lack of biogenic, low-δ13C soil \({{{\rm{CO}}}}_{2}\). Instead, such high δ13C values can be attributed to pyrite oxidation and resulting sulfuric acid dissolution followed by prior calcite precipitation (PCP)33,45 and/or rate-dependent fractionation33. PCP comprises calcite precipitated before the drip water reaches the top of the stalagmite.

Linear regression slopes between speleothem Mg/Ca and Sr/Ca ratios can be used to detect the influence of PCP on the proxy data46,47 and we indeed find such evidence (Figs. S11, S12 and S13 in the Supplementary Information). Recent isotope modelling studies have found that PCP does not only affect the δ13C value but can also result in changes in δ18O depending on the conditions for isotopic exchange processes48,49,50,51. To test this, we conducted isotope simulations using the model by Skiba and Fohlmeister49 where we enabled isotopic exchange of \({{p}}{{{\rm{CO}}}}_{2}\) and \({{{\rm{H}}}}_{2}{{{{\rm{O}}}}}\) with the bicarbonate in the drip water. We assumed a cave temperature of 3 °C (Tab. S1) and check if, after some time for PCP, exchange processes are able to bring the δ18O values of the bicarbonate back to initial δ18O. We find that exchange processes are too slow (when the PCP occurred within the cave48) to re-equilibrate the bicarbonate isotope composition to pre-PCP values when temperatures and cave air \({{p}}{{{\rm{CO}}}}_{2}\) are low as is expected for cold, high-altitude caves (Fig. S14).

We compared the PCP-sensitive Mg/Ca proxy to the δ13C and δ18O records to investigate whether there is a positive relationship between them which would be expected if PCP had a major impact on the isotope records. We applied a low- (>4000 yrs) and high-pass (<4000 yrs) filter on the composite records to distinguish between correlations on IS-S timescales (high-pass filtered data) and longer timescales (low-pass filtered data). Pearson correlation coefficients of low-pass filtered records are negative for δ13C vs. δ18O and δ18O vs. Mg/Ca but positive for δ13C vs. Mg/Ca (Table 1). For the high-pass filtered time series, all correlation coefficients are positive which indicates that PCP is influencing all proxies. Thus, we find that PCP is a dominant process affecting the δ13C, δ18O and Mg/Ca proxies in the Melchsee-Frutt speleothems on IS-S timescales. However, on longer timescales δ13C and Mg/Ca are as well dominated by PCP while δ18O is likely predominantly reflecting broader, regional-scale δ18Oppt variability which results in a negative relationship with the PCP-sensitive proxies (δ13C and Mg/Ca). Accordingly, we find a high and positive correlation between δ18O and RSL (r = 0.82, p < 0.0001) on longer timescales (longer than 4000 yrs).

Table. 1 Pearson correlation coefficients between high-pass and low-pass filtered δ13C, δ18O and Mg/Ca.

Enhanced PCP has commonly been attributed to water-limited conditions in the vadose karst system associated with a dry climate52,53,54,55,56. However, climate model simulations suggest that interstadials were characterised by higher precipitation than stadials in the Northern Hemisphere57 during both winter and summer, following the Clausius-Clapeyron-relationship (Fig. S15 b and c). Higher precipitation amounts lead to shorter times for PCP due to a higher availability of water in the karst system which could potentially result in decreased C and O isotope values during interstadials, contrary to what our record shows. Therefore, we suggest that the PCP variability in our Alpine speleothems reflects the extent of host rock dissolution where enhanced dissolution and, thus, higher initial \({{Ca}}^{2+}\) concentration in the drip water will lead to increased subsequent calcite precipitation, i.e. higher amounts of PCP (Fig. S14). It is important to note that this is independent of the amount of infiltrating water, which would solely influence the duration of PCP.

In this setting, the extent of host rock dissolution depends on the amount of pyrite oxidation. The oxidation of pyrite releases sulfate and requires the presence of water and dissolved oxygen58:

$${{FeS}}_{2}+{H}_{2}O+\frac{7}{2}{O}_{{2}_{{aq}}}\to {{Fe}}^{2+}+{2{SO}}_{4}^{2-}+{2H}^{+}$$
(1)

Studies in the Swiss Alps have found that water originating from snow melt has dissolved oxygen contents of up to 60-80 % of the atmospheric value. On the contrary, water derived from glacier ice melt is usually depleted in oxygen reaching only 35-40 % with respect to the atmospheric value. This can be explained by the closer contact of snow with atmospheric \({{{\rm{O}}}}_{2}\)59. Furthermore, subglacial meltwater is likely depleted in oxygen because the subglacial environment is known to host aerobic microbial communities59,60,61. Since the extent of pyrite oxidation (and, thus, drip water \({{Ca}}^{2+}\) concentration and the amount of PCP, Fig. S14) depends on the availability of dissolved oxygen in the water62,63, we suggest that during interstadials higher amounts of water with higher \({{{\rm{O}}}}_{2}\) content infiltrated into the karst in the caves' catchment, i.e. a higher amount of precipitation and snow melt (Fig. S15) and less meltwater from glacier basal ice. Conversely, a higher proportion of \({{{\rm{O}}}}_{2}\)-depleted glacier meltwater entered the karst during stadial times. We reflect on alternative mechanisms and why they are unlikely to explain enhanced PCP observed during interstadials in the Method section.

Two processes could have contributed to a larger proportion of \({{{\rm{O}}}}_{2}\)-rich water during interstadials compared to stadials (Fig. 3): (1) higher snow accumulation in winter and more summer surface meltwater due to higher summer temperatures together with higher rainfall in summer (Fig. S15), (2) larger glacier thickness during stadials, thus, lowering the melting point due to increased pressure and resulting in a higher basal meltwater contribution.

Fig. 3: Hypothesised variability of Alpine glaciers and associated input of meltwater into the subglacial karst system on IS-S timescales.
figure 3

Host rock and cave shown in grey and brown, respectively. During stadials a thicker glacier likely resulted in a higher contribution of \({{{\rm{O}}}}_{2}\)-depleted basal meltwater due to the pressure-induced lowering of the melting point. Crevasses might have enabled \({{{\rm{O}}}}_{2}\)-saturated surface meltwater to reach the base of the glacier. During interstadials, thinner ice likely increased the melting point and, thus, reduced the amount of basal meltwater contribution. Additionally, the \({{{\rm{O}}}}_{2}\)-rich surface water contribution was likely higher due to higher precipitation together with enhanced summer meltwater (Fig. S15).

Northern Alps speleothem δ18O reflects North Atlantic seawater δ18O

Previous studies of speleothems of glacial age in the Northern Alps suggested that the speleothem calcite δ18O reflects δ18Oppt7,13, which today is strongly correlated with air temperature64,65. Here, we present results from an isotope-enabled climate model, COSMOS-wiso (ref. 23, see “Methods” section), which confirms higher δ18Oppt during interstadials compared to stadials for Western Europe and the North Atlantic (Fig. 4). Positive IS-S differences associated with millennial-scale AMOC oscillations23 prevail in the North Atlantic which is the main moisture source of the Alps (area between 30 to 60°N)10,17. Our model simulations suggest δ18Oppt values for the Alps which are 0.5 to 1.5‰ higher during interstadials compared to stadials (Fig. 4). We applied a PCP correction on the speleothem δ18O (Fig. 1a, see “Methods” section and Supplementary Information for details) to obtain an improved estimate of δ18Oppt. Indeed, after the PCP correction the amplitude of interstadial-stadial differences in speleothem calcite is close to IS-S differences found for earlier glacials4,7,9,12,13,16 and in model simulated δ18Oppt over the Alps (Fig. 4).

Fig. 4: Map of simulated interstadial-stadial (IS-S) difference in δ18O of meteoric precipitation (δ18Oppt) and vertically integrated moisture transport for the North Atlantic and Europe.
figure 4

Colour scale refers to the IS-S mean annual δ18Oppt difference and arrows depict the moisture transport, where longer arrows indicate larger IS-S difference. Also shown is the Last Glacial Maximum (LGM) ice extent in the European Alps74. Inset shows enlarged Alpine LGM ice sheet and the location of the study site (Melchsee-Frutt, black star). Colour bar refers to both the main map and the inset.

We find that the δ18Oppt IS-S difference in the Northern Alps is the result of the IS-S difference of North Atlantic surface seawater δ18O. Seawater δ18O of the main moisture source region of the Alps (area between 30 to 60°N) was ~0.5 to 2 ‰ higher during interstadials compared to stadials (Fig. S15g), potentially due to increased evaporation. Thus, it is likely the main driver of the positive IS-S difference in simulated δ18Oppt over the Alps as recorded by Northern Alps speleothems. This is corroborated by δ18O values of planktonic foraminifera from sediment cores in the northern and eastern North Atlantic66,67. Their δ18O values indicate a ~1–2‰ IS-S difference in seawater δ18O when accounting for the ~10 °C calcification temperature difference on IS-S timescales, which is often found in sea surface temperature reconstructions from this region68.

A previous study explains speleothem δ18O variation on glacial and interglacial timescales by zonal shifting of the westerlies leading to enhanced moisture advected to the Northern Alps from the South during the LGM, which is depleted in δ18O due to orographic rain out10. However, the COSMOS-wiso model indicates a different pattern on IS-S timescales by simulating a larger contribution of southerly-sourced moisture during warmer compared to colder periods (indicated by IS-S difference in moisture transort, Fig. 4). As this is based on simulations of one single model, further investigations should be conducted to verify this result. However, based on that we conclude that shifts in the position of the westerlies and resulting change in moisture source and pathway are unlikely to explain the higher δ18Oppt over the Alps during interstadials relative to stadials.

Conclusions and Implications

The presented δ13C record is punctuated with extremely positive values. They occur contemporaneously with δ18O maxima. We propose that these are the result of an increased proportion of \({{{\rm{O}}}}_{2}\)-rich annual precipitation and/or surface summer snow melt compared to a smaller proportion of \({{{\rm{O}}}}_{2}\)-depleted glacier meltwater during interstadials relative to stadials. During interstadials, the glacier ice above the Alpine caves was increasingly melting and potentially comparably thin. This could have reduced pressure on the basal ice and thereby the melting point which would lead to less basal meltwater contribution to the karst. Larger contribution of \({{{\rm{O}}}}_{2}\)-rich waters then enabled more pyrite to oxidise and carbonate rock to dissolve, leading to higher initial \({{{{\rm{Ca}}}}}^{2+}\) concentrations and, thus, enhanced PCP.

For stadials, speleothem δ13C values indicate relatively cold conditions throughout the year resulting most likely in a comparably thick glacier. Thus, annual precipitation and/or summer snow melt amount were supposed to be small, whereas potentially relatively larger amounts of basal meltwater entered the karst (Fig. 3). Larger contribution of \({{{\rm{O}}}}_{2}\)-depleted basal meltwater lead to reduced pyrite oxidation, resulting in less carbonate rock dissolution and decreased \({{{\rm{Ca}}}}^{2+}\) concentrations and, hence, reduced PCP compared to interstadials.

According to our interpretation, speleothem δ13C values below the host rock δ13C value of ~2.7 ‰ provide evidence of carbonate dissolution due to soil \({{{\rm{CO}}}}_{2}\) contribution, and speleothem δ13C values above ~2.7 ‰ might indicate the presence of a glacier above the cave site. Thus, the δ13C speleothem record would suggest that the catchment of the caves became glaciated already at the end of MIS 9 (Fig. 1b, where δ13C values show no isotopic signs of soil influence at the end of MIS 9). However, IS-S variability is only visible in the speleothem δ18O and not in δ13C and Mg/Ca, which suggests that changes in water sources on IS-S timescales are negligible at that time, i.e. potentially indicating the absence of a glacier. Speleothem δ13C suggests that the cave site was continuously glaciated throughout MIS 8 with strong changes in meltwater contributions on IS-S timescales. During MIS 7e, c and b the catchment was likely ice-free as suggested by the low δ13C values indicating a soil C contribution. During MIS 7d, an interglacial substage, the glacier re-advanced and buried the cave site (indicated by a diminishing soil signal in speleothem δ13C) and we find rapid variations in δ13C and PCP-corrected δ18O values during MIS 7d (Fig. 1b) which we consider to result from similar mechanisms as for glacial times.

Our study provides new insights into several relevant and emerging aspects of speleothem science. We show that not only the duration for PCP, determined by the air vs. water content in the karst, but also the amount of PCP (i.e., the difference between initial \({{Ca}}^{2+}\) of the cave drip water, which is determined by the soil \({{p}}{{{\rm{CO}}}}_{2}\) and equilibrium \({{Ca}}^{2+}\) concentrations, when the solution is in equilibrium with cave air \({{p}}{{{\rm{CO}}}}_{2}\)) need to be considered when interpreting speleothem stable isotope records. This has important implications for speleothem-based climate proxy interpretation. In the vast majority of studies, PCP-sensitive proxies have been interpreted to reflect the amount of infiltration based on the negative relationship between the duration of PCP and water availability in the karst system. However, the amount of PCP might follow an opposite pattern whereby wetter conditions promote soil bioproductivity69 and, thus, increases the initial \({{Ca}}^{2+}\) concentration of drip water with respect to equilibrium \({{Ca}}^{2+}\) concentrations and thereby the amount of PCP.

Our data suggests a strong PCP influence not only on speleothem δ13C but also on δ18O which previously has been thought to be limited due to fast re-equilibration of the dissolved bicarbonate with the drip water as initially suggested by Hendy70. Since then, more accurate equilibration times have been reported71 implying relatively slow equilibration with the drip water as demonstrated also by isotope modelling studies (refs. 48,49, Fig. S14). Oxygen isotopes are the single most extensively used speleothem-based climate proxy and are of special interest to compare with climate models simulating δ18Oppt72,73. We disentangled the two main signals on IS-S timescales recorded in speleothem δ18O (δ18Oppt vs. PCP) and find a much better match with the isotope-enabled climate model IS-S differences in δ18Oppt after correcting for PCP. This also corroborates the isotope-enabled model results that Northern Alps δ18Oppt recorded in speleothem archives reflect North Atlantic seawater δ18O changes on IS-S timescales, but only when accounted for PCP. Thus, it is crucial to focus future research on the impact of PCP on speleothem δ18O (PCP in the vadose zone vs. on the stalactite, impact of \({{{\rm{CO}}}}_{2}\) exchange) and advancing methods and models to disentangle the PCP component from (isotope) proxies.

If we would have not accounted for PCP in our study, climate-related δ18Oppt changes would have been enormously overestimated. This would be difficult to explain with climatic processes alone (e.g., interstadial speleothem δ18O values as high as interglacial values, Fig. 2). However, it has to be pointed out that the effect of PCP can vary for cave and drip sites and for each case it has to be evaluated how to account for it (e.g., when PCP occurs already in the karst leaving much more time for re-equilibration48). PCP corrections involve parameter assumptions which can also introduce errors and such corrections are not always recommended. Furthermore, the strong effect of PCP at our site is also an advantage as it results in more pronounced interstadial-stadial transitions in our speleothem isotope records and thereby improves accuracy when trying to detect them. A major uncertainty of our correction approach arises from the kinetic fractionation effect which we only approximate in our correction. This underscores the need for more laboratory experiments in order to derive robust empirical relationships for quantifying cave-internal processes in speleothem records.

Our study highlights the potential of high-elevation speleothems as an unique archive for studying palaeoglacier dynamics. Such studies of Alpine speleothems which grew during glacials should be extended to other glacials and potentially other mountain regions providing a test bed to validate models of mountain glacier dynamics on millennial to orbital timescales74,75.

Methods

Stable isotope analyses

Samples for stable C and O isotope measurements were milled along the main growth axis. We hand-milled at 0.2 and 0.5 mm-resolution for M5_14_44B and M5_14_42, respectively. For 0.5 mm resolution we milled samples 0.2 mm wide to maintain a similar length of time over which each sample integrates, but at 0.5 mm steps. Stalagmite M43_401_14 was milled continuously at 0.1 and 0.3 mm-resolution. We obtained 2246, 1103 and 2113 samples for stalagmite M5_14_44B, M5_14_42 and M43_401_14, respectively. Sampling resolution for all stalagmites was guided by the changing growth rate obtained from U-Th ages (mean growth rates are 70, 49 and 42 mm/ka for M5_14_44B, M5_14_42 and M43_401_14, respectively).

Isotopes for stalagmites M5_14_44B and M5_14_42 were measured using an automated carbonate-extraction system (KIEL IV) interfaced with a MAT253 isotope ratio mass spectrometer (IRMS; ThermoFisher Scientific) at the Helmholtz Centre GFZ in Potsdam. Samples of 60-90 µg are dissolved in 103 % H3PO4 at 72 °C for 12 minutes and the isotopic composition was measured on the released and cryogenically purified \({{{\rm{CO}}}}_{2}\). Replicate analysis of reference material (NBS19) reported relative to VPDB yielded 1 sigma uncertainties of better than 0.06 ‰ for both δ18O and δ13C. Isotope measurements of M43_401_14 were conducted using a Delta V plus IRMS at the University of Innsbruck. This mass spectrometer is linked to an on-line Gasbench II carbonate preparation system. Details on the measurement procedure can be found in Spötl and Vennemann76. Long-term 1 sigma precision is better than 0.08‰ for both δ13C and δ18O. All isotopic ratios are expressed in the delta notation relative to VPDB.

Trace element analyses

The elements Ca, Mg, and Sr were measured along the isotope transects of all three stalagmites using a M4 Tornado micro-XRF scanner at the Helmholtz Centre GFZ in Potsdam. Element intensities were acquired as counts per second [cps], but log-ratios of element intensities are linear functions of element concentrations and provide robust information of relative element concentrations77 (Figs. S12 and S13). For comparison with the stable isotope data the trace elemental records were down-sampled using Gaussian Kernels with a width of 0.5 mm.

U/Th dating

Ages for all stalagmites were determined via U/Th disequilibrium dating. Sub-samples of ~200 mg were drilled along growth layers close to the growth axis (Fig. S2). A total of 16 samples were measured for stalagmite M5_14_44B, 18 for M5_14_42, and 25 for M43_401_14. Measurements were conducted with a MC-ICP-MS (ThermoFisher Neptune Plus) at Heidelberg University following the analytical protocol of Wefing et al. 78.

Samples are first dissolved, then spiked with 229Th, 233U, and 236U and subsequently U and Th purified using wet-column chemistry with Eichrom UTEVA resin. Isotope ratios were used to calculate U-series ages according to the decay equations79 and error propagation of statistical uncertainty. For this, 234U and 230Th half-lives published in Cheng et al. 80 were used. Correction for detrital contamination assumes a 232Th/238U weight ratio of 3.8 ± 1.981 and 230Th, 234U and 238U in secular equilibrium, and is minor for most samples (Tab. S2). All U and Th isotope measurements have been collected simultaneously and data was corrected for mass fractionation and tailing as mentioned by Wefing et al. 78.

Age-depth modelling

Age-depth modelling was performed using iscam26, a programme designed to align records from the same cave (i.e., assuming that a similar signal is recorded) into a master record, which improves the time control provided by individual time series by reducing the age uncertainties within the overlapping periods and increases signal-to-noise ratios of the combined record. The algorithm correlates dated proxy signals from several contemporaneously growing stalagmites by varying their age-depth relation within age uncertainties. It determines the most probable age-depth model and calculates the age uncertainty for the combined record. We run 30,000,000 Monte-Carlo iterations on the time series to find the best option for the age-depth realisation, i.e. the iteration with the highest correlation coefficient between both time series. Significance of the highest correlation coefficient is tested by surrogate time series, which have similar characteristics than the measured isotope records (e.g., same dating information, proxy resolution, memory of the time series26).

For each iteration, we generate 2000 surrogate time series. To calculate the significance limits, we tested each of these 2000 time series in the same way as the measured time series, except that we used only 1000 randomly selected iterations due to computational constraints. For each of the 2000 artificially constructed time series, the highest correlation, found within the 1000 iterations, was used to construct a probability density function, which defines significance limits for the correlation coefficient of the measured time series. All found correlation coefficients of the best-suited age-depth realisations for the individual stacking steps exceed the 95% confidence limits. The resulting isotope records plotted over age agree perfectly with each other (Fig. S3).

Within iscam we applied a pointwise linear interpolation between dated depths to produce the δ18O composite record. We stacked the individual δ13C records to a composite record in the same way as for δ18O and placed it on the age-model obtained with the δ18O data. Thus, each data point in the composite records of δ18O and δ13C was generated by the exact same stalagmite(s) samples. The composite records are shown in Figs. S5 and S6.

PCP correction

As explained in the main text, speleothem δ18O is likely strongly modified by PCP. Thus, we aimed to disentangle the speleothem δ18O in its PCP and δ18Oppt components using the corresponding speleothem δ13C as PCP proxy. Since laboratory experiments, isotope modelling (including our sensitivity test using an speleothem isotope model, see Supplementary Information, Fig. S14) and studies investigating speleothem isotope data suggest that kinetic fractionation needs to be accounted for in speleothem isotope studies82,83,84,85,86,87, we utilise an estimate of the kinetic fractionation for δ18O relative to the fractionation of δ13C87 (Fig. S16, Supplementary Information for details) to correct our speleothem δ18O data with an estimate of the δ13C-PCP component derived from our speleothem δ13C data. We apply the PCP correction on the measured speleothem δ18O data for each speleothem isotope data point where the δ13C value is above the host rock value (2.7 ‰). We use the δ13C enrichment above the host rock value (difference to 2.7 ‰) as estimate of δ13C increase solely due to PCP and convert it to the expected PCP-driven δ18O enrichment using the estimate of the kinetic fractionation for δ18O relative to the fractionation of δ13C (Figs. S16 and S17 in the Supplementary Information).

Alternative mechanisms cannot explain enhanced PCP during interstadials

To our knowledge other mechanisms than the one proposed in the main text cannot lead to the larger degree of PCP during interstadials compared to stadials. This section is intended to discuss these other mechanisms and demonstrates that they cannot explain the data.

We expect negligible differences between interstadials and stadials in karst temperature and in infiltrating (melt-) water temperature. When a glacier covered the cave site temperatures below the glacier and in the karst must have been always close to 0 °C. Same can be expected for surface meltwater. Thus, temperature-dependent mechanisms like rate of pyrite oxidation or oxygen saturation in water can be ruled out as explanation.

Carbonate dissolution is comparatively faster than the residence time of water in the karst (saturation achieved in a few hours up to several days88,89,90), so that equilibrium between \({{{\rm{CO}}}}_{2}\) dissolved in the water and \({{Ca}}^{2+}\) ions should be always attained irrespective of the hydrological conditions (except possibly for drip sites dominated by fracture flow following heavy rainfall).

Since the δ13C data indicates no soil and/or vegetation above the cave during glacial times (Fig. 2), also cave air \({{p}}{{{\rm{CO}}}}_{2}\) changes can likely be ruled out as a mechanism for IS-S PCP variability as near-constant atmospheric \({{p}}{{{\rm{CO}}}}_{2}\) values are expected in the cave. As there is only the atmosphere as the main \({{{\rm{CO}}}}_{2}\) reservoir, variability in cave ventilation would not result in substantial varations in cave air \({{p}}{{{\rm{CO}}}}_{2}\).

Climate model simulations

To investigate responses of δ18Oppt to S-IS transitions during glacial periods, we used model simulation experiments from Zhang et al. 23, which successfully simulated glacial millennial-scale climate oscillations using the climate model COSMOS-wiso. The model COSMOS-wiso is a water isotope-enabled fully-coupled atmosphere-ocean general circulation model91. The atmospheric model is ECHAM592, complemented by the land surface component JSBACH93, and was used at T31 resolution (3.75°), with 19 vertical layers. The ocean model is MPI-OM94 and includes sea-ice dynamics modelled with viscous-plastic rheology95. It has a resolution of GR30 (3° × 1.8°) in the horizontal, with 40 uneven vertical layers.

We chose modelling results from the experiment E40ka_Orb which characterises unforced millennial-scale changes in Atlantic Meridional Overturning Circulation under fixed glacial conditions, resembling glacial S-IS transitions. More details about experimental design can be found in Zhang et al. 23. We show differences in meteoric precipitation δ18O (Fig. 4) and seawater δ18O, precipitation amount and temperature over the eastern North Atlantic and Europe (annual, summer and winter) between the simulated interstadial and stadial states (IS-S, Fig. S15).