Skip to main content
Science Advances logoLink to Science Advances
. 2022 Oct 26;8(43):eabq0611. doi: 10.1126/sciadv.abq0611

Rapid and precise measurement of carbonate clumped isotopes using laser spectroscopy

Nitzan Yanay 1,*, Zhennan Wang 1, David L Dettman 1,2, Jay Quade 1, Katharine W Huntington 3, Andrew J Schauer 3, David D Nelson 4, J Barry McManus 4, Kaustubh Thirumalai 1, Saburo Sakai 5, Anna Rebaza Morillo 1, Ananya Mallik 1
PMCID: PMC12419128  PMID: 36288314

Abstract

Carbonate clumped isotope abundance is an important paleothermometer, but measurement is difficult, slow, and subject to cardinal mass (m/z) interferences using isotope ratio mass spectrometry (IRMS). Here, we describe an optical spectroscopic measurement of carbonate clumped isotopes. We have adapted a tunable infrared laser differential absorption spectrometer (TILDAS) system to measure the abundances of four CO2 isotopologues used for clumped isotope thermometry. TILDAS achieves the same precision (0.01‰ SE) as IRMS measurements rapidly (∼50 min per carbonate analysis) and using small samples (<2 mg of calcite), without making assumptions about 17O abundance in the sample. A temperature calibration based on 406 analyses of CO2 produced by digestion of 51 synthetic carbonates equilibrated at 6° to 1100°C is consistent with results for natural carbonates and previous calibrations. Our system results were indistinguishable from IRMS systems after replicating the InterCarb interlaboratory calibration. Measurement by TILDAS could change the landscape for clumped isotope analysis.


A fast, automated, high-throughput spectroscopic system is implemented for carbonate clumped isotope thermometry.

INTRODUCTION

In recent years, measurement of rare isotopologues with more than one heavy isotope has been a subject of growing interest. Perhaps the most ubiquitous application of this “clumped isotope” research is carbonate clumped isotope thermometry. With this method, the relative abundance of multiply-substituted isotopologues containing 18O-13C bonds is used to estimate the formation temperature of carbonate minerals (1, 2). The key advantage of this method is that no knowledge of the water isotopic composition from which the mineral was precipitated is required, also allowing one to calculate the δ18O composition of the parent water using a conventional oxygen isotope fractionation relationship.

The field of carbonate clumped isotope geochemistry has advanced rapidly since it was first demonstrated as a practical paleothermometer by (3). Interlaboratory standardization efforts have greatly improved reproducibility (4, 5), discrepancies between empirical temperature-clumping calibrations have been markedly reduced (6, 7), and automation and development of new isotope ratio mass spectrometry (IRMS) techniques have shortened analysis times and decreased the sample size required for analysis (810). Despite these substantial improvements, the low natural abundance of clumped isotopologues and their nonunique mass continue to hinder IRMS clumped isotope measurements. Measurement remains difficult and costly and requires 17O corrections that introduce significant uncertainty (11, 12). Because of these limitations, the method is underused compared to conventional stable isotope methods, with a relatively small number of laboratories producing reliable data.

To address these issues, we developed a tunable infrared (IR) laser differential absorption spectrometer (TILDAS) for clumped isotope analysis of CO2. It has been previously demonstrated that this spectrometer is capable of accurately and precisely (0.01‰ SE) measuring clumping in CO2 gas in 20 to 25 min per analysis and with sample sizes of 1.9 mg calcite equivalent (13). By using mid-IR absorption instead of mass to measure the relative abundance of isotopologues, the TILDAS system determines clumping more directly than IRMS because it measures the molecular species rather than the cardinal mass (m/z) and hence does not require, for example, a 17O correction for mass interference or removal of contaminant isobaric molecules (13). However, in previous work, only CO2 samples from gas cylinders were analyzed, not gas derived from acid digestion of carbonate mineral samples. This additional step is required for a direct comparison to IRMS systems and to demonstrate the applicability of TILDAS clumped isotope analysis to carbonates. Spectroscopic clumped isotope measurements of carbonate-derived CO2 have been previously reported by other researchers (14). However, (14) and (13) were proof-of-concept papers by design and did not attempt to compare performance to IRMS using carbonates.

Here, we report the results from a new, fully automated carbonate clumped isotope analyzer system that extends previously published TILDAS instrument capabilities to measure carbonates. We present results from 1678 analyses performed over a 61-day period. Using this dataset, we establish an empirical temperature versus clumping calibration based on a set of carbonates synthesized or equilibrated in the laboratory. The calibration is validated using natural samples that have known formation temperatures. The synthetic samples are also used to establish a temperature versus calcite-water oxygen isotope fractionation (αcalcite-water) relationship. In addition, we compare TILDAS-derived data to IRMS data using a series of CO2 gases equilibrated with waters at different temperatures [carbon dioxide equilibrium scale (CDES)] and a carbonate-based method [InterCarb-CDES (I-CDES)] for projecting raw data into an interlaboratory reference frame. Clumping values of seven interlaboratory standards are compared to accepted values measured using IRMS systems. We demonstrate that the TILDAS system reliably produces clumped isotope data that are competitive in quality with the best IRMS systems but with a much higher sample throughput and small sample size.

Notation

IRMS-based carbonate clumped isotope thermometry involves defining a measurable temperature-dependent clumping Δ47 value for CO2 (1) derived from carbonate. In this study, we adopt the analogous TILDAS-based measure of clumping in CO2 of (13). This laser-specific notation is based on the Air Force Geophysics Laboratory shorthand notation, which has been used in the spectroscopic high-resolution transmission (HITRAN) database for decades (15). In this notation, 16O12C16O, 16O13C16O, 16O12C18O, and 16O13C18O are referred to as 626, 636, 628, and 638, respectively. Isotopologues are identified by the second digit of the atoms’ atomic mass. We define the composition of isotopologue i as

δi=([isample][626sample][istandard][626standard]1)× 1000 (1)

The molecular ratios δ636 and δ628 are good approximations of the atomic isotopic composition [see (16) for definitions of “molecular” and “atomic”], δ13C and δ18O, respectively, whereas δ638 is approximately equal to (δ636 + δ628) and thus to (δ13C + δ18O) and is used in place of IRMS δ4747 = [(R47,sample/R47,working reference gas) − 1] × 1000) (17). R47 is the abundance ratio of mass-47 CO2 to the isotopically normal isotopologue, 16O12C16O. These ratios are expected to be nearly equal to the atomic isotopic composition because the 636 and 628 isotopologues are naturally three and two orders of magnitude more abundant than the next most common isotopologues with 13C and 18O, respectively (18). Given that the differences between the atomic and molecular ratios are expected to be negligible (16) and that projection into the Vienna Pee Dee Belemnite (VPDB) reference frame using International Atomic Energy Agency (IAEA) values for carbonate standards effectively eliminates minor discrepancies, all results given in the VPDB reference frame are reported as δ18O and δ13C to avoid introduction of additional and potentially confusing notation. This assumption is in fact shown to be valid in this study. Only raw values (defined below) are reported as δ628 or δ636.

The following isotope exchange reaction is referred to in the clumped isotope calculation

O16C12O16+O16C13O18O16C13O16+O16C12O18 (2)

The equilibrium constant of isotope exchange reaction (Eq. 2) can be written as

Keq=[O16C13O16]× [O16C12O18][O16C12O16]× [O16C13O18] (3)

The abundance of isotopologue i relative to that expected if all isotopes were randomly distributed (stochastic distribution) among all isotopologues is defined as

Δi=(RiRi1)×1000 (4)

where Ri is the abundance of isotopologue i relative to the nonsubstituted isotopologue and the * in Ri stands for the random distribution condition. A previous study (1) derived the approximation

ΔO16C13O181000ln(KeqKeq*)=1000ln([O16C13O16]× [O16C12O18][O16C12O16]× [O16C13O18]) (5)

where Keq* is the equilibrium constant for the fully randomized stochastic distribution and, in the case of Eq. 2, Keq* = 1. We therefore denote the CO2 clumping relative to stochastic distribution as measured by TILDAS as Δ638. While similar to the Δ47 notation used by the IRMS community, Δ638 is different because (i) its measurement does not involve (nor necessitate) a correction for isobaric interferences and (ii) Δ638 includes only the isotopologues in Eq. 3. In contrast, Δ47 includes contributions from all m/z-47 isotopologues (1), although the 16O13C18O isotopologue contributes roughly 97% of the 47 mass (18), so differences between Δ638 and Δ47 are expected to be minor with respect to analytical precision. A “raw” subscript (e.g., δ628raw) indicates that values are reported relative to a working reference (WR) gas. Following analogous conventions for IRMS-based clumped isotope measurements, values presented as “CDES” are projected into an interlaboratory reference frame using equilibrated/heated gas anchors (4, 13). Values presented as “I-CDES” are projected into an interlaboratory reference frame using ETH carbonate standards (Materials and Methods) (5). We follow the definitions of (19) for “sample,” “analysis,” and “anchor.”

RESULTS

Sample size and analytical throughput

The automated carbonate clumped isotope analyzer system is composed of an autosampling carousel for carbonate digestion in phosphoric acid, a CO2 extraction and purification line, a bellows pressure-control system, and the TILDAS instrument previously described by (13) (Fig. 1) (Materials and Methods). The analytical setup does not require a gas chromatographic or equivalent cleanup step and can process up to 36 carbonate powders in a single fully automated run. However, this would require more than 24 hours, so we chose to load roughly 25 carbonate powder samples (standards and unknowns) per 24-hour period, in addition to three water-equilibrated or heated gas CO2 samples that were analyzed daily. Using this procedure, we ran approximately 28 analyses total every 24 hours, including the time required to reload the carousel.

Fig. 1. System schematic.

Fig. 1.

A diagram of the fully automated carbonate clumped isotope analyzer. The system is composed of a CO2 extraction and cleanup section, a sample collection, mixing and delivery system, and an inlet and spectroscopic analysis component. P sensor, pressure sensor; LN2, liquid nitrogen Dewar; N2, nitrogen gas; WR, working reference gas; concentric circles, critical orifice. See Materials and Methods for a detailed description of the analysis protocol.

The carbonate samples analyzed included 45 synthetic carbonates we precipitated at known temperature from water of known δ18O composition, analyzed 5 to 11 times each; 8 carbonates re-equilibrated to temperatures of 450°C to 1100°C, analyzed 3 to 10 times each; and 17 natural carbonates formed at known temperature, analyzed 3 to 21 times each. We also typically analyzed six different carbonate standards, four interlaboratory ETH standards and two intralaboratory standards, in each carousel run. The MK standard (Mallinckrodt AR analytical reagent, a synthetic calcium carbonate powder) was analyzed roughly twice in each run (n = 130), and the others approximately once per run (nETH-1 = 40, nETH-2 = 41, nETH3 = 41, nETH-4 = 41, ncar-2 = 82). Car-2 is a Carrera marble. In addition, commercially available primary IAEA standards NBS-18, NBS-19, IAEA-CO-8, IAEA-603, and LSVEC were analyzed 4 to 10 times each to establish VPDB values for car-2 and MK. Last, we analyzed a set of carbonate standards used in the InterCarb project (5) and distributed to the University of Washington (UW) IsoLab, ETH-1, ETH-2, ETH-3, ETH-4, IAEA-C1, IAEA-C2, and MERCK (n = 6) (table S4).

During the analytical session that produced the data reported in this study, we performed 1678 analyses in 61 days, averaging 27.5 analyses per day (including downtime). Of these, 1563 analyses are presented here, including 318 gas sample analyses and 1245 carbonate analyses (data S1). To our knowledge, this sample throughput is a major improvement on the capabilities of current IRMS systems. The TILDAS system performs an entire analysis, including acid digestion, in ∼50 min, while the newest IRMS systems require 6 to 10 hours to achieve comparable precision. The minimum sample size required for a single analysis (Δ638 or Δ47 0.01‰ SE) is 1.9 mg of pure carbonate, competitive with modern automated clumped IRMS systems. This amount of carbonate produces sufficient CO2 for a measurement at 60 torr and a CO2/N2 mixing ratio of 0.35%.

Dependence of Δ638 on δ628 and projection of raw Δ638 values into an interlaboratory reference frame

An important step in clumped isotope analysis is the projection of raw Δ47 or Δ638 values for CO2, measured relative to a specific WR gas, into an interlaboratory reference frame. For IRMS systems, this requires a bulk isotope correction that accounts for the dependence of Δ47 on δ47. This dependence is due to pressure baseline signals in the Faraday cups of the smaller ion beams (20, 21) and scrambling of ions during ionization in the instrument source (17). As the TILDAS system does not ionize the CO2 gas, there is no reason to expect a similar dependence of Δ638raw on δ638raw. To test for bulk dependence and project our raw data into an interlaboratory reference frame, we prepared and analyzed a set of CO2 gases with different δ636 and δ628 compositions equilibrated at different temperatures (Materials and Methods). As expected, measurements of CO2 gas samples with different δ636 compositions equilibrated at the same temperature show a δ636raw dependence indistinguishable from zero for gases equilibrated at 6°C and 23°C (0.0003 ± 0.0005 Δ638raw/1‰ δ636raw and −0.0003 ± 0.0004 Δ638raw/1‰ δ636raw, respectively) but display a weak dependence for gases equilibrated at 60°C (−0.0011 ± 0.0004 Δ638raw/1‰ δ636raw) (Fig. 2A). Although the common slope of all gas samples combined is statistically distinguishable from zero (−0.00062‰ ± 0.00005 Δ638raw/1‰ δ636raw), we did not apply a δ636 correction because the magnitude of this dependence is very small, is inconsistent between equilibration temperatures, and is more likely to introduce noise (by adding a poorly quantified correction factor) than to improve the accuracy of the results. However, measurements of gas samples equilibrated at the same temperature but with different δ628raw compositions demonstrate that the TILDAS system has a linear dependence of measured Δ638raw on the δ628raw composition of the sample (0.0034‰ Δ638raw/1‰ δ628raw) (Fig. 2B). The possible causes of the dependence of Δ638raw on δ628raw are discussed below (Discussion).

Fig. 2. Bulk isotope dependence.

Fig. 2.

The empirical relationship between ∆638raw and δ636raw (A) and between ∆638raw and δ628raw (B) in the equilibrated and heated gases. Symbols: individual measurement (four sample-WR comparisons); vertical and horizontal error bars: 1 SD uncertainty in ∆638raw, δ636raw, and δ628raw measurements; r: Pearson correlation coefficient. In (A), only samples with equal δ628raw composition and different δ636raw composition are shown. In (B), only samples with equal δ636raw composition and different δ628raw composition are shown.

We adopted the method used in IRMS studies (4, 17) for monitoring the δ628 dependence by regularly measuring equilibrated and heated CO2 gases that bracket a range of 77.1‰ in δ628 space and a temperature range of 6° to 1000°C, bracketing the compositions of all the carbonate-derived CO2 analyzed. Although the magnitude of the δ628 dependence can vary over time [note lack of dependence in (13)], it remained stable enough for the duration of this calibration effort (22 July 2021 to 2 November 2021) such that a single average correction was sufficient for the entire carbonate dataset.

From theoretical considerations, the more a sample differs from the WR gas in its δ628 composition, the more susceptible it is to this dependence and to potential shifts in the dependence magnitude. For example, a change in δ628 dependence from 0.0034‰ Δ638raw/1‰ δ628raw to 0.0024‰ Δ638raw/1‰ δ628raw would induce an apparent shift of 0.02‰ in Δ638 for a sample with a δ628raw value of 20‰ versus WR but would only cause a 0.0005‰ shift in Δ638 for a sample with a δ628raw value of 5‰ versus WR. Therefore, to minimize the impact of this dependence, we selected a WR gas that has a bulk isotope composition similar to CO2 derived from most of the carbonates analyzed (δ13C = −4.701 ± 0.0178‰ VPDB and δ18O = −1.707 ± 0.012‰ VPDB). In this way, the δ628 dependence correction (with its associated uncertainty) was multiplied by a smaller δ628raw factor.

We projected the sample Δ638 data into the interlaboratory CDES reference frame using as anchors (i) only equilibrated CO2 gas measurements (4), which are referenced to theoretical clumping values for Δ638, and (ii) only ETH carbonate standards, which necessitate a direct comparison of IRMS-derived Δ47 values to TILDAS-derived Δ638 values (Materials and Methods) (5). The results demonstrate that such a comparison is valid (Fig. 3). Our mathematical approach is that of (12) and takes the form

Δraw47=a×Δ47+b×δ47+c (6)

or in our system

Δ638raw=a×Δ638+b×δ628+c (7)

where δ628 is used in place of δ638 because the TILDAS system Δ638 values only depend on δ628 (Fig. 2). Using the equilibrated CO2 approach (a = 1.0069 ± 0.0748; b = 0.0034 ± 0.000008; c = −0.9434 ± 0.0575), the “a” factor is close to unity, indicating excellent agreement between the theoretical and observed Δ638 values of the gas anchors, with a very slightly expanded scale for the TILDAS system, similar to (13). Using ETH carbonate standards as anchors (a = 1.0648 ± 0.2204; b = 0.0030 ± 0.00005; c = −0.9668 ± 0.0404) produces a larger “a” factor, indicating a more expanded scale using carbonate anchors compared to the gas anchor reference frame, but this difference is statistically insignificant given the uncertainties. The smaller δ628 dependence (b = 0.0034 ± 0.000008 versus 0.0030 ± 0.00005) may be due to the much narrower range in δ628 that is bracketed by the ETH standards (17.3‰) compared to the gas samples (77.1‰), although the dependence appears to be linear for the entire range (Fig. 2). Notably, the uncertainties in Δ638 and ultimately in temperature are larger using ETH anchors because they bracket a smaller range in δ628 and in Δ638 space (tables S3 and S4 and data S1) similar to the effect of bracketing a smaller range in δ47 for IRMS systems (5).

Fig. 3. InterCarb standards comparison.

Fig. 3.

Δ47 and Δ638 results from IRMS and TILDAS measurements of seven Interlaboratory standards: ETH-1, ETH-2, ETH-3, ETH-4, IAEA-C1, IAEA-C2, and MERCK. IRMS results are from the InterCarb study (5). In (A) to (D), raw data are projected into the CDES reference frame using equilibrated and heated gases as anchors. IRMS results are ordered from left to right in the order of laboratories A to J as reported in table 1 of (5) with regular intervals. Each circle represents the mean value for a standard reported by a single laboratory. The blue line represents the weighted mean value for all IRMS laboratories. Red diamonds represent TILDAS mean values for each standard for powders received in March 2019, analyzed 40 times for ETH-1 and 41 times for ETH-2,3,4. Purple diamonds are TILDAS mean values for the same standards but for the exact powders distributed to UW IsoLab for the InterCarb study, analyzed six times each. Error bars are ±1 SE. In (E) to (G), results are projected into the I-CDES reference frame using carbonate standards ETH-1,2,3 for IRMS results and ETH-1,2,3,4 for TILDAS results. The results from IRMS instruments 1 to 26 as reported in table 3 of (5) are ordered from left to right. Note that not all mass spectrometers 1 to 26 reported data for all three standards. In all panels, IRMS values are corrected to 70°C acid digestion by adding a factor of 0.022‰.

System stability

The ability of the TILDAS system to consistently produce accurate results can be assessed by regular measurement of homogeneous standard materials. To evaluate the stability and reproducibility of the system, we analyzed six homogeneous carbonates on a near-daily basis during the analytical session (Fig. 4). The SDs of Δ638 for the carbonate standards (nMK = 130, nETH-1 = 40, nETH-2 = 41, nETH3 = 41, nETH-4 = 41, ncar-2 = 82) ranged from 0.019 to 0.035‰. The “pooled reproducibility” is 0.023‰, calculated as described by (19) using all homogeneous carbonates, standards, and unknowns analyzed at least five times. For comparison, the SD of the WR gas measured against itself is 0.0154‰. This represents the expected reproducibility limit for the TILDAS system, as WR versus WR comparisons involve no acid digestion and do not go through the CO2 purification line, both of which could potentially add noise.

Fig. 4. Carbonate standards stability.

Fig. 4.

(A) Repeated analyses of intralaboratory (car-2, MK) and interlaboratory (ETH) carbonate standards between 22 July 2021 and 2 November 2021. Clumped isotope values are projected into CDES using equilibrated and heated gases. Carbonates are digested in 70°C acid, and Δ638CDES values are not corrected to 25°C. Error bars are ±1 SD of individual analyses. Number of analyses: car-2, 82; MK, 130; ETH-1, 40; ETH-2, 41; ETH-2, 41; ETH-4, 41. (B) Deviation from the mean value of each standard for the same analyses as (A). Error bars removed for clarity.

We used a homogeneous carbonate that has a δ628raw value close to that of the WR gas to assess the amount of noise contributed from acid digestion and CO2 purification. The Δ638 value of such a carbonate is minimally affected by the bulk correction, as explained above. Intralaboratory standard C64 has an SD of 0.0160‰ (n = 21), only 0.0006‰ larger than the WR gas, likely because its δ628raw composition is only −5.70‰ (table S4 and data S1) versus the working gas. Although the number of analyses is smaller for C64 than for other standards, this suggests that the acid digestion and gas cleanup steps have only a minor impact on the overall reproducibility of the system. For conventional stable isotope composition, the pooled reproducibility for δ13C is 0.0305‰ and for δ18O is 0.0367‰, both competitive with IRMS systems.

Comparison of interlaboratory standards to accepted IRMS values

A key question regarding the TILDAS system is whether the Δ638, δ628, and δ636 values it generates can be directly compared to IRMS Δ47, δ18O, and δ13C values, respectively. Because the TILDAS system measures molecular ratios, while IRMS systems measure atomic ratios, there could be inherent differences between the values produced by the two methods.

The recent InterCarb clumped isotope community project (5) provides an excellent dataset for comparison with TILDAS results, as it demonstrates the interlaboratory reproducibility of carbonate standards between IRMS systems and provides consensus values for seven standards (ETH-1, ETH-2, ETH-3, ETH-4, IAEA-C1, IAEA-C2, and MERCK). To compare TILDAS results to IRMS values, we analyzed four of these standards (ETH-1,2,3,4), received in March 2019, 40 to 41 times each, and separately analyzed the full set of InterCarb standards, using the exact same powders distributed to UW IsoLab, six times each. After correcting InterCarb values to a 70°C acid digestion temperature by adding a factor of 0.022‰ (5), the difference between TILDAS Δ638CDES and mean IRMS Δ47CDES values for the ETH standards ranges between −0.012 and +0.017‰, with an average absolute difference of 0.0095‰ (Fig. 3).

Three more standards, IAEA-C1, IAEA-C2, and MERCK, were analyzed in the InterCarb project and projected into the I-CDES reference frame using ETH standards as anchors. Since our Δ638CDES values of the ETH standards are comparable to IRMS values, we applied the same methodology using the InterCarb consensus values for ETH standards corrected to a 70°C acid temperature as anchors. It should be noted that the InterCarb IRMS laboratories used ETH-1, ETH-2, and ETH-3 as anchors and treated ETH-4 as an unknown, while we used all four ETH standards as anchors. After projecting the IAEA-C1, IAEA-C2, and MERCK results into the I-CDES reference frame, the differences between the mean IRMS Δ47,I-CDES from all InterCarb instruments and the TILDAS Δ638,I-CDES values for IAEA-C1, IAEA-C2, and MERCK were 0.0025, 0.0000, and −0.0174‰, respectively. As with the gas anchor method, these differences are smaller than those observed between IRMS laboratories (Fig. 3) and do not appear to be systematic, implying that a direct comparison of TILDAS CDES or I-CDES values to IRMS CDES or I-CDES values, respectively, is as valid as a comparison of results from different IRMS instruments.

TILDAS raw δ628 and δ636 values were projected into the VPDB scale using carbonate standards with consensus δ18O and δ13C values (Materials and Methods). Again, using the InterCarb dataset, we find that TILDAS δ636 and δ628 VPDB values are within the range of IRMS δ13C and δ18O VPDB values for the standards measured (fig. S3). A single exception is the δ636 value of MERCK, which is 0.14‰ lower than the lowest mean value for any InterCarb IRMS instrument. The MERCK standard has the most extreme δ636 value, −42.21‰ VPDB, so its abnormal value compared to IRMS could be a result of a slight inaccuracy in the measured VPDB value of our MK secondary standard, which anchors the low end of our δ636 scale. Alternatively, it could be a result of an inherent difference between δ636 and δ13C that is only manifested at extreme compositions. The current dataset does not allow us to resolve the issue, as more standards with low δ636 composition are required. In either case, these results indicate that comparing TILDAS δ628 values to IRMS δ18O values is equivalent to comparing δ18O values from different IRMS instruments. The same is true for TILDAS δ636 and IRMS δ13C for all but extremely low δ636 values. Although the TILDAS system measures δ636 and δ628, the normalization of these ratios to IAEA standards with assigned δ18O and δ13C values results in equivalent values. We therefore suggest that singly substituted carbonate isotope ratios can be reported using the traditional δ13C and δ18O VPDB notation.

Empirical temperature calibration

A fundamental tool in low-temperature isotope geochemistry and paleoclimate studies is the temperature dependence of oxygen fractionation between carbonate minerals and the water they precipitate from. To establish an empirical calibration based on TILDAS δ18O measurements of calcite, we precipitated 44 carbonate powders under controlled conditions at temperatures of 6.3° to 69.7°C (Materials and Methods and the Supplementary Materials). One sample was 3% aragonite and 97% calcite, and the rest were 100% calcite (data S2). By measuring the oxygen isotope composition of the carbonates and the water, we constructed a rigorous calcite-water fractionation-temperature calibration, resulting in the following relationship

1000lnαcalite-water=18.0508±0.0875×(1000T)32.1996±0.2850R2=0.996 (8)

This relationship is indistinguishable in slope and intercept from the widely cited IRMS calibration of (22) (Fig. 5) and improves on most low-temperature calibrations in the number of samples, mineralogic homogeneity, and temperature range and arguably in the method of carbonate precipitation. The experimental setup used for this series of precipitations (fig. S1) has advantages over most previous studies in that it resulted in highly reproducible experiments (fig. S2) at low calcite saturation indexes (table S1) in well-mixed solutions and with temperature control better than 0.2°C. Furthermore, solution chemistry (pH, electrical conductivity, and temperature) was tracked continuously throughout the experiments, allowing a better understanding of the precipitation process as opposed to only knowing the start and end conditions (Materials and Methods and the Supplementary Materials). This calibration can be coupled to the clumped isotope-temperature calibration below to calculate the δ18O value of the parent water of a carbonate of interest, with adjustments for noncalcite mineralogy. Although TILDAS δ18O values appear to be comparable to IRMS δ18O values (see above), we also measured the δ18O composition of the carbonates using conventional IRMS (Materials and Methods) to produce a purely IRMS-based calibration, which may be preferable for IRMS users. The resulting relationship is

1000lnαcalcite-water=18.0972±0.0870×(1000T)32.0952±0.2851R2=0.994 (9)

Fig. 5. Conventional calcite-water 18O fractionation-temperature calibration.

Fig. 5.

Black circles are 1000lnαcalcite-water values based on mean δ18O values of 41 synthetic calcite samples prepared for this study and the δ18O values of the water they precipitated from. Calcite powders were analyzed 5 to 11 times each. Horizontal and vertical error bars are smaller than markers. The red line is a linear regression based on TILDAS δ18O values of the carbonates. The brown line is the regression based on IRMS δ18O values of the same samples. Blue diamonds are data points from (22). The blue line is the linear regression of (22).

Both relationships are within error of each other in slope and intercept (Fig. 5).

We used the same set of 44 synthetic calcites to construct a Δ638CDES-temperature calibration, analyzing each sample 5 to 11 times for a total of 359 clumped isotope analyses. The resulting relationship is

Δ638CDES=0.405±0.0006×(106T2)+0.1822±0.0061R2=0.985 (10)

An additional set of seven carbonates reequilibrated to temperatures of 450° to 1100°C (Materials and Methods) was added to expand the range of the calibration to temperatures relevant to diagenetic and metamorphic processes. These were analyzed 4 to 10 times for a total of 47 analyses. It should be noted that the high-temperature (>70°C) part of the calibration is less robust than the low-temperature segment. There is a large sampling gap between 70° and 450°C and a second large gap between 600° and 1100°C. In addition, the solid-state reordering method used to equilibrate samples in the 450° to 600°C range was less reproducible, making it difficult to ascertain that the samples fully equilibrated to the desired temperature. We therefore regard the high-temperature end of the calibration with caution. The 1100°C sample, prepared using a piston cylinder apparatus, is more robust because this high-temperature, high-pressure experimental setup very likely induced complete recrystallization (Materials and Methods). The temperature-clumping relationship calculated using the full set of samples is given by the equation

Δ638CDES=0.0409±0.0003×(106T2)+0.1776±0.0031R2=0.997 (11)

Despite the shortcomings of the high-temperature segment of the calibration, the expanded range relationship is within the uncertainty of the low-temperature calibration. The two overlap in the 6° to 70°C range and only diverge slightly at the high-temperature end (Fig. 6), suggesting that the high-temperature carbonates are near equilibrium or, possibly, at the same disequilibrium as the low-temperature samples. The key difference is the much smaller uncertainty in slope and intercept in the full-range calibration, a result of the two data clusters (6° to 70°C and 450° to 1100°C) forming a pseudo–two-point regression. The reduced uncertainty using the full-range calibration propagates to temperature uncertainties that are roughly 1°C smaller for unknown samples formed at Earth surface temperatures (table S3). Both the low-temperature and full-range relationships are not within the uncertainty of a recent IRMS calibration compilation (6) (Fig. 6), but the difference between the TILDAS and IRMS relationships is no larger than the discrepancies observed between previous IRMS calibrations. Because IRMS Δ47 measurements include contributions from the 17O13C17O and 17O12C18O isotopologues, which have a stronger temperature dependence than the 16O13C18O isotopologue measured separately by TILDAS, the Δ47-reciprocal-temperature relationship should, in theory, have a slightly lower slope compared to Δ638 (1, 6), consistent with our result. However, the theoretical differences are too small to fully explain the discrepancies between the calibrations.

Fig. 6. Empirical temperature calibration.

Fig. 6.

(A) The black circles are mean Δ638 CDES values of synthetic calcite prepared for this study, plotted against their precipitation/reequilibration temperatures. Uncertainties in Δ638 and in temperatures are smaller than circles. The red line is a weighted linear regression based on the synthetic samples precipitated between 6° and 70°C, as shown in main figure. The dashed brown line is a regression based on both the low-temperature samples and samples reequilibrated at 450° to 1100°C, shown in the bottom left of the inset. The colored symbols are mean Δ638 values of natural carbonates plotted against their known formation temperatures. Vertical error bars are smaller than symbols, and horizontal error bars represent uncertainties in formation temperatures. The blue line is the IRMS calibration of (6) with the intercept adjusted to 70°C acid digestion temperature. (B) Same as (A) but samples are projected into the I-CDES reference frame using ETH anchors. The blue line is the calibration of (7).

Recent work has shown that using ETH carbonate anchors instead of gas anchors results in better agreement between temperature-clumping calibrations produced by different laboratories (7). Our results indicate that the same approach can be applied to project Δ638raw values into the I-CDES. Using this method, the temperature-clumping relationship based on only the low-temperature synthetic carbonates becomes

Δ638I-CDES=0.0395±0.0007×(106T2)+0.1800±0.0074R2=0.986 (12)

Expanding the calibration to include the 450° to 1100°C samples yields the following relationship

Δ638I-CDES=0.0386±0.0002×(106T2)+0.1902±0.0025R2=0.996 (13)

The low-temperature calibration is within the uncertainty of the IRMS calibration of (7). The full-range calibration slope is within the uncertainty of the calibration of (7) slope, and the intercept is slightly higher (Fig. 6). It is likely that the shallower slope in the carbonate anchor relationship compared to the gas anchor relationship is a result of the larger-scale compression introduced by projection into the I-CDES compared to projection into the CDES (larger “a” parameter in the conversion). Ultimately, both methods produce indistinguishable temperatures for the carbonates analyzed in this dataset, but temperature uncertainty is ∼0.5° to 1°C larger using the carbonate anchors (table S3 and data S1).

Natural samples

To test the ability of the calibration to predict the formation temperatures of natural carbonates, we analyzed 17 natural samples with known formation temperatures ranging from 7.8° to 39°C (Materials and Methods). These include 4 inorganic and 13 biogenic carbonates and a mix of calcite and aragonite (table S2). Using the low-temperature gas anchor calibration, the temperature predicted by the calibration and the measured Δ638CDES value of 13 of the samples is within the uncertainty of the known formation temperature (table S3). Temperature uncertainties are calculated on the basis of the 95% confidence interval (1.96 SE) of Δ638 values using full error propagation (19) also propagating the uncertainties in the Δ638-temperature calibration slope and intercept. The largest offset between predicted and known temperatures (39° ± 4°C and 33°C, respectively) is for the spring tufa sample. Spring tufas formed in similar settings have been shown to deviate from expected clumped composition (23), although not in all cases (24). In addition, the sample in question was collected from a shallow stream in Ash Meadows National Wildlife Refuge, Nevada, which may reach higher temperatures during the summer months compared to those measured when the sample was collected in January. However, the sample was collected directly upstream of a pool that has only a 2°C annual temperature cycle based on temperature logger data (fig. S4), so stream temperature fluctuations are likely limited. Another sample not within the uncertainty of predicted temperature is the Laghetto Basso calcite. This material was initially thought to deviate significantly from empirical clumping-temperature calibrations (25), but a recent calibration appears to have resolved the discrepancy (7). In either case, the known temperature is only 0.4°C outside of the predicted temperature range. Two other outliers are aquatic gastropods. The reason for their deviation is unclear because other samples of the same species collected at different springs and other species collected from the same springs did not deviate beyond uncertainty from expected temperatures. In total, the average absolute offset between predicted and measured values was 2.3 ± 1.7°C (1 SD) including all natural samples and 2.1 ± 1.3°C excluding the biggest outlier, the tufa sample (table S3).

Using the extended temperature range calibration produces at most a 0.2°C difference in expected temperatures of the natural samples but a roughly 1°C smaller uncertainty. While appealing, the smaller uncertainty means that four more samples have known temperatures that are not within the uncertainty of predicted values, so the smaller uncertainty may be statistically true but unrealistic. Using ETH carbonate anchors increases the average offset between predicted and measured values to 2.6° ± 1.9°C (2.3 ± 1.7°C excluding the tufa outlier), so it appears to be slightly less accurate based on the limited number of natural samples analyzed in this study. However, because of the larger temperature uncertainty, only 3 of 17 samples have known temperatures not within the uncertainty of the predicted temperatures based on the low-temperature calibration (table S3).

DISCUSSION

This study describes an analytical approach, an operational routine, and a data analysis method that produce precise clumped isotope data using the TILDAS instrument with a high sample throughput and small sample sizes. It also establishes temperature-clumping and temperature-oxygen isotope fractionation calibrations validated using natural samples. Although the system is performing well, a lingering issue is the bulk isotope dependence. The system has a linear dependence of Δ638 on δ628 and no dependence of Δ638 on δ636 (Fig. 2). Although seemingly similar to IRMS bulk dependence, the TILDAS instrument’s nonlinearity effects are driven by different mechanisms. In this TILDAS dual laser instrument, laser 1 is tuned to 2250.2 cm−1 to measure the absorbance of the 16O13C16O and 16O13C18O isotopologues, and laser 2 is tuned to 2285.1 cm−1 to measure the 16O12C16O and 16O12C18O absorbance peaks [see figure 1 in (13)]. Therefore, the spectral fitting, whose basis functions are laser intensity baseline forms and absorption line shapes, is conducted separately for each laser. The deviation of baseline fitting from “true baseline” is a possible explanation of the nonlinearity and its different magnitude in δ636raw and δ628raw. A better understanding of this bulk dependence could lead to its elimination and even better precision, as seen by the SDs of carbonate standards with small δ628raw values. Until a better solution is found, the method presented in this study is effective in monitoring and correcting for this dependence.

Both the equilibrated CO2 gas and carbonate standard approaches tested for projection of raw data into an interlaboratory reference frame work well. On the basis of a comparison to the InterCarb dataset, using either gas anchors or ETH carbonate anchors, both resulted in Δ638 values that are within the range of IRMS Δ47 values for the carbonate standards tested, meaning that a direct comparison of TILDAS and IRMS datasets is as accurate as comparing datasets from two IRMS instruments.

The advantages and disadvantages of each approach for IRMS analysis are discussed in detail in (5). Specifically for the TILDAS system (and as with select IRMS systems), the carbonate anchor approach has a distinct advantage because carbonate processing is entirely automated, allowing one to easily run more anchors, while gas samples have to be manually introduced into the gas purification line. The calibration based on carbonate anchors is also in better agreement with the most recent IRMS calibration (7) than the gas-based calibration is with the latest equivalent IRMS calibration (6), although this improved agreement may be a product of forcing the Δ638 values of ETH anchors to exactly match IRMS consensus values. Although based on theory (1, 6) we do expect a slight difference between the temperature dependence of Δ638 and Δ47, this theoretical difference is too small to fully explain the discrepancies between the gas anchor calibrations. However, using gas anchors reduces the analytical uncertainty in Δ638 and ultimately the temperature uncertainties. Moreover, ETH anchors cannot be used to monitor (or correct) for a potential future dependence of Δ638raw on δ636raw because they do not provide two carbonates with identical clumped temperatures and identical δ628raw compositions that bracket a wide range in δ636 space. We have not observed a significant δ636 dependence, but until the bulk dependence issue is better understood, the ability to check for changes in this dependence is important and necessitates the use of some gas anchors.

There are several limitations to the results presented in this study. While spectroscopic systems do not suffer from mass interferences, optical interferences can occur. An unrecognized absorbance peak overlapping one of the peaks used for our Δ638 measurement could potentially result in an erroneous value. We are not aware of any common gases that absorb in our narrow measured spectral region and have not observed these peaks, but optical interferences are possible for the TILDAS system. If these gases were produced by acid reaction with some natural samples, a gas chromatographic or equivalent purification step may become necessary, but presently no such purification seems to be required. A second limitation of the system is the need to thoroughly mix CO2 samples with nitrogen before measurement, a step that is time-consuming and makes it difficult to recapture a gas sample for repeated analyses. The mixing ratio of CO2 to nitrogen in the sample must closely match that of the WR gas for best performance. While this is relatively simple to do for the bulk CO2/N2 ratio, individual isotopologues can have very different mixing ratios from the WR gas depending on the bulk composition and clumping of the sample. This could be a source of measurement error or bias. As discussed above, the TILDAS system also requires a δ628 correction, the underlying causes of which are still not fully understood. Last, the conversion of δ636raw and δ628raw data into the VPDB reference frame may be improved by establishing more accurate VPDB values for our secondary standards. This could also resolve the discrepancy between TILDAS and IRMS δ13C VPDB values observed for the MERCK sample.

The results presented in this study open the door to exciting potential future applications. The high analytical throughput of the TILDAS system and the small sample size required can quickly produce large robust datasets with increased sample replication. This will allow higher resolution sampling of materials, critical to paleoclimatic reconstructions, and could expand the use of clumped isotope geochemistry to new users and fields. Future plans to couple the system to an additional single-laser spectrometer capable of high-precision 17O measurements could produce conventional, clumped, and triple oxygen data with a simple rapid analysis and a sample size of only <2 mg calcite equivalent (∼15 μmol CO2). Dual clumped isotope analyses, which have recently been shown to resolve kinetic effects in certain natural carbonates (26, 27), could conceivably be done using an additional dual-laser or single-laser spectrometer in a separate optical box with a different spectral range. With the TILDAS system, measurements would not be limited to isotopologue pairs with different masses (e.g., Δ47 and Δ48), as is the case with IRMS, and could focus on Δ638 and Δ828 directly.

In conclusion, the TILDAS system described in this study is comparable in sample size to the best IRMS systems and improves on IRMS systems in analysis time, sample throughput, and simplicity, as it does not require a gas chromatographic or equivalent CO2 purification step. In addition, it is capable of measuring individual isotopologues directly, regardless of their mass, eliminating the need for 17O corrections. On the basis of these results, the TILDAS system is now set to substantially advance the field of carbonate clumped isotope research and low-temperature geochemistry as a whole.

MATERIALS AND METHODS

Experimental design

The objectives of this study are to (i) evaluate the long-term reproducibility and stability of the TILDAS instrument for clumped isotope analysis, (ii) evaluate the projection of raw data into an interlaboratory reference frame using gas and carbonate anchors, (iii) determine whether clumped and conventional isotope data produced by the TILDAS system are directly comparable to IRMS data, (iv) construct an empirical temperature-clumping calibration, and (v) test the system’s ability to reconstruct formation temperatures of natural samples. We took the following steps to meet these objectives:

1) To test long-term reproducibility, we regularly analyzed a set of six homogeneous carbonates for the duration of the study. Reproducibility is assessed on the basis of the SD of all analyses for each sample and on observed variations in analysis results over time. The pooled reproducibility of all carbonates analyzed at least five times is also considered.

2) A set of gas anchors and carbonate anchors was routinely analyzed during the study, allowing us to use both to project data into interlaboratory reference frames. The methods are evaluated on the basis of the analytical uncertainty they produce for unknown samples, the technical ease of use for each method, and the interlaboratory reproducibility achieved with each method.

3) Once data are projected into an interlaboratory reference frame, results from seven interlaboratory standards are compared to a large dataset of IRMS analyses of the same standards. The difference between mean IRMS values and TILDAS values is compared to the variability of results between different IRMS laboratories.

4) Fifty-one carbonates are precipitated at or reequilibrated to controlled temperatures between 6.3° and 1100°C and analyzed. The results are used to construct a temperature-clumping calibration.

5) Seventeen natural samples with known formation temperatures are also analyzed. The results of the natural sample analyses are applied to the temperature-clumping calibration to calculate predicted temperatures for the natural samples. A comparison of the predicted temperatures to known temperatures is used to test the empirical calibration.

Gas extraction/cleanup line and laser system description

The fully automated carbonate clumped isotope analyzer has three parts: (i) gas extraction/dehydration line, (ii) CO2 sample mixing and delivery system, and (iii) TILDAS dual-laser instrument (Fig. 1). The gas extraction line follows common methods for carbonate-phosphoric acid reaction and cryogenic purification of the CO2 gas. The reaction is conducted at 70°C in a temperature-controlled container, in which an autosampler carousel enables up to 36 samples to be processed in unattended operation. This gas extraction/dehydration procedure takes ca. 20 min. The sample mixing and delivery system thoroughly mixes CO2 gas with N2 gas in a stainless steel bellows and loads the mixture into the TILDAS instrument for the CO2 isotopologue suite measurement. By diluting the CO2 in a buffer gas, the spectral lines are broadened, which improves the precision of the spectral fitting in the data processing software and reduces the impact of the laser jitter noise. As a result, the repeatability of the analysis is improved. The gas-mixing procedure takes ca. 20 min. The TILDAS dual laser instrument, described in (13), directly and simultaneously measures four CO2 isotopologues (16O12C16O, 16O13C16O, 16O12C18O, and 16O13C18O) in the mid-IR spectral regions of 2250.2 cm−1 and 2285.1 cm−1. Isotopologues associated with 17O are not present in the absorption peaks measured for the clumped isotope calculation. The automated sampler, gas extraction line, and the sample mixing apparatus were constructed in-house at the University of Arizona, whereas the TILDAS instrument was custom-built for this project by Aerodyne Research (Billerica, MA) based on existing laser designs. A detailed, step-by-step protocol for operating the system is given in the Supplementary Materials.

Equilibrated and heated gas samples

A set of gas anchors was used to project raw data into CDES. Equilibrated gases were prepared at the University of Arizona. Pyrex tubes with 10 μl of water were evacuated on a high vacuum line by first freezing the water with an ethanol/LN2 slurry. After evacuation, the ethanol/LN2 slurry was replaced by LN2 and sufficient CO2 for a single clumped isotope analysis was frozen into the tube, after which the tube was flame-sealed. The gas was then allowed to equilibrate at 6°, 23°, or 60°C for at least 2 weeks before measurement. Temperature was tracked using HOBO loggers at 1-hour intervals. To maximize the bulk composition range, we used one gas with δ13C = ∼−4‰ VPDB and δ18O = ∼−7‰ VPDB and a second gas with δ13C = ∼−40‰ VPDB and δ18O = ∼−20‰ VPDB. These were equilibrated with either Vostok water (δ18O = ∼−31‰ VPDB) or evaporated water (δ18O = ∼+32‰ VPDB), resulting in four gas types with bulk compositions spanning 36‰ in δ13C space and 77‰ in δ18O space. To verify the linearity of the bulk dependence, we prepared 12 samples with intermediate δ18O compositions (−15‰, +2‰, +6‰, and +9‰ VPDB), all equilibrated at 23°C.

Heated gases were prepared using the same procedure but without water and replacing Pyrex tubes with quartz tubes. The samples were heated to 1000°C for at least 24 hours before analysis.

Synthetic samples

We used carbonate samples prepared in the laboratory to establish temperature calibrations. Carbonates were precipitated at controlled temperatures using a custom-built apparatus composed of a reaction vessel, two reagent vessels, a peristaltic pump, a water chiller/heater circulator, and a pH/electric conductivity (EC)/temperature meter with probes (fig. S1). Each experiment began with pure water in the reaction vessel and 20 mM solutions of NaHCO3 and CaCl2 in the reagent vessels. The solutions were equilibrated at the desired temperatures by circulating water through the water-jacketed vessels. After equilibration, reagent solutions were slowly dosed into the reaction vessel and solution was removed from the reaction vessel at an equal rate using the peristaltic pump. In this way, the concentrations of Ca2+ and HCO3 in the reaction vessel gradually increased while maintaining a constant water volume until supersaturation was reached and carbonate was precipitated. When sufficient carbonate had precipitated for clumped isotope and x-ray diffraction (XRD) analyses, the experiments were ended and the carbonate was scraped off the vessel walls, filtered from the solution, dried, and stored for analysis. The pH, EC, and temperature in the reaction vessel were logged at 1-min intervals. The resulting pH/EC curves show that the experiments are highly reproducible (fig. S2) and that carbonate was precipitated at a low saturation state and therefore likely near equilibrium. On the basis of XRD analyses performed at the University of Arizona Department of Chemistry, all but one sample were 100% calcite. One sample contained trace amounts (∼3%) of aragonite. See the Supplementary Materials for a detailed description of the precipitation apparatus and the experiments.

High-temperature samples

To extend the range of the clumping-temperature calibration, carbonate samples were equilibrated at 450°, 500°, 550°, and 600°C by solid-state reordering. For each sample, sufficient calcite for multiple clumped isotope analyses was inserted into a quartz tube. The tube was evacuated and filled with enough CO2 gas for a single TILDAS analysis (2 mg calcite equivalent) before the tube was flame-sealed. Samples were then heated to the desired temperature for 21, 9, 13, and 4 days (for 450°, 500°, 550°, and 600°C, respectively) for solid-state reequilibration, following (28). After heating, the tubes were cracked, and the water vapor and CO2 pressure were measured to detect decarbonation of the calcite. In some cases, the isotope composition of gas was also measured for the same purpose. After extracting the gas, the carbonate was removed and stored for analysis. Our intralaboratory standards, car-2 and MK, as well as BB (vein calcite spar from Barrancas Blancas, northern Chile), were used as starting material because they have different grain sizes and initial clumped and bulk isotope values. In this way, we could test whether samples had equilibrated to the desired temperature. In several samples, including all samples with MK as a starting material, the heated powder was heterogenous with respect to clumping (differences of ∼0.1‰ between replicate analysis) likely due to incomplete resetting. These samples were not used in the calibration. A single sample prepared with car-2 as the starting material and heated to 700°C for 5 days produced a homogeneous powder but was a clear outlier in the temperature calibration. Adding it to the calibration substantially degraded the regression fit and the accuracy with which the calibration predicted the temperatures of natural samples, so it was omitted from the calibration.

One sample was recrystallized at 1100°C using a high-temperature, high-pressure system at the University of Arizona Experimental Petrology Laboratory. The sample was pressurized to 1 GPa and heated to 1100°C using a half-inch assembly in a piston cylinder apparatus. Twenty milligrams of reagent-grade calcite (Alfa Aesar, 11403; CaCO3 5 μm powder, minimum 95% purity) was placed in a 2-mm-diameter, 7.07-mm-high Au75-Pd25 capsule, surrounded by a MgO sleeve. A four-hole Al2O3 sleeve as thermocouple insulator, type C thermocouple, graphite furnace for heating, and BaCO3 pressure medium were used. The sample was pressurized to 1 GPa, and then the temperature was increased to 1100°C at a rate of 50°C/min. It was then allowed to equilibrate for 48 hours. At the end of the experiment, the sample was rapidly quenched to room temperature by shutting off power, reaching <50°C within 31 s.

IRMS analysis

Mass spectrometric measurements were used to determine the δ18O composition of the waters that the synthetic samples precipitated from and to calculate an oxygen fractionation–temperature calibration that is entirely IRMS-based. Water δ18O was measured on an automated gas-source isotope ratio mass spectrometer (Finnigan Delta S) at the University of Arizona Environmental Isotope Laboratory. Samples were equilibrated with CO2 gas at approximately 15°C in an automated equilibration device coupled to the mass spectrometer. Standardization is based on international reference materials Vienna Standard Mean Ocean Water (VSMOW) and Standard Light Antarctic Precipitation (SLAP). Analytical precision (1σ) is 0.1‰ or better for δ18O on the basis of repeated intralaboratory standards. Carbonate δ13C and δ18O values were measured using an automated carbonate preparation device (KIEL-III) coupled to a gas-ratio mass spectrometer (Finnigan MAT 252). Powdered samples were reacted with dehydrated phosphoric acid under vacuum at 70°C. The isotope ratio measurement is calibrated on the basis of repeated measurements of NBS-19 and NBS-18, and precision is ±0.10‰ for δ18O and ±0.08‰ for δ13C (1σ).

Natural samples

A set of natural samples with known formation temperatures was used to validate the temperature-clumping calibration. These included a ~13,000-year-old ostrich egg shell from Gona, Ethiopia [∼39° ± 1°C (29)] and a large freshwater bivalve collected in 1998 from Lake Tanganyika in central Africa. The shell was collected from a spit of land just south of Kigoma, Tanzania, latitude 4.912°S, 24° to 27°C (30); subaqueous mammillary calcite precipitated on the walls of the Devil’s Hole cave system [33.7°C (25, 31)]; subaqueous calcite coating formed at the bottom of Laghetto Basso, a small lake in Corchia Cave, Italy [7.9°C (25)]; aquatic gastropods and spring tufa collected in January 2020 from warm, high discharge springs in Ash Meadows National Wildlife Refuge, Nevada (32) (28° to 33°C; fig. S4); and modern non-encrusted deep mixed-layer dwelling planktic foraminifera Globorotalia truncatulinoides (dextral) that flourish during January and February (mean temperature = 18°C) picked from the topmost sediments in core 2010-GB2-MCC raised in the Garrison Basin, Western Gulf of Mexico (3335). More details are provided in table S2.

Statistical analysis

Bivariate linear regressions used for calculating the temperature calibrations, including uncertainties in slope and intercept, were performed using the method of (36), implemented by the MATLAB code of (37). For the Δ638-clumping calibration, the mean values of replicate analyses of each sample were used as single data points, with the 95% confidence interval, calculated as described below, used as uncertainty. Each point in the regression was weighted using the inverse square of the Δ638 uncertainty (1.96 SE) and the inverse square of the temperature uncertainty taken as either the temperature variation (SD) during sample preparation or the analytical uncertainty of the temperature probe, whichever was larger. For the oxygen isotope fractionation calibration, points were weighted by the inverse square of the uncertainty in 1000lnα propagated from the SD in δ18O values of replicate analyses of each sample.

The parameters used to project raw data into the CDES and I-CDES reference frames were calculated using the method of (19), adjusted to our system. Because our Δ638 values are only dependent on δ628, we replaced δ47 with δ628 in the calculation and Δ47 was replaced with Δ638. To calculate conversion parameters for the CDES reference frame, equilibrated and heated gas anchors are used. The theoretical Δ638 values for the gases at the equilibration temperatures (6.2°, 22.8°, 63.5°, and 1000°C) were calculated using a seventh-order polynomial fit to the calculated Δ638 values from (6), in which the International Union of Pure and Applied Chemistry (IUPAC) parameters were used (6, 38). For the I-CDES conversion, the InterCarb consensus Δ47 values for ETH-1,2,3,4 anchors are used (5), adjusted to 70°C acid digestion by adding a factor of 0.022‰. This method includes full error propagation of the uncertainties in the parameters of the transfer function and the analytical reproducibility of the system. We used a “pooled reproducibility” approach to calculate the reproducibility of Δ638, δ628, and δ636 as described by (19). All carbonates that were analyzed at least five times are included in the calculation. This method was implemented using a MATLAB script based in part on that used by (11) and in part on the “ClumpyCrunch” Python code from (19).

Clumped isotope (Δ638) uncertainties presented in this study for each sample are 95% confidence intervals calculated as 1.96 SE of the CDES or I-CDES value of the sample. The mean value of replicate analyses of the sample and the external pooled reproducibility are used. These uncertainties and the uncertainties of the slope and intercept of the temperature-clumping calibration are propagated to calculate the temperature uncertainty of the sample.

The VPDB values of our secondary standards, car-2 and MK, were determined by comparison to primary IAEA standards NBS-18, NBS-19, IAEA-CO-8, IAEA-603, and LSVEC (LSVEC was only used as a δ13C standard) analyzed on the same day. The conversion of δ628raw and δ636raw values to VPDB for unknown samples was done using a moving window approach with our secondary standards, car-2 (δ18O = −1.593 and δ13C = 1.969‰) and MK (δ18O = −22.608 and δ13C = −43.244‰). For each analysis, all car-2 and MK results in the previous 25 analyses and the next 25 analyses (51 analyses window centered around the analysis of interest) in the dataset are used to establish conversion factors to VPDB values. This window width typically corresponded to 2 to 3 days of analyses and two to four MK and car-2 analyses. For example, for analysis 100 in the dataset, the average δ628raw and δ636raw values for all MK and car-2 samples in the analysis range 75 to 125 are calculated. Then, a conversion function to VPDB, specific to analysis 100, is calculated on the basis of the raw δ628 and δ636 and established δ18O and δ13C VPDB values of the two standards; this is then applied to analysis 100. Expanding the window width or centering it asymmetrically (for example, previous 40 analyses and following 10 analyses) resulted in a degradation of the pooled reproducibility of δ18O and δ13C. Narrowing the window width resulted in some windows that had no MK or no car-2 analyses, so a VPDB conversion function could not be calculated.

An abrupt change in δ628raw and particularly in δ636raw values occurred after routine pump maintenance on 29 October 2021 (analysis 1595 of 1679 in the dataset). To account for this change, we used a cutoff in the moving window analysis such that windows for analyses before 1595 did not include MK or car-2 data after analysis 1595 and windows for analysis after 1595 did not include MK or car-2 data from before 1595.

Acknowledgments

We would like to thank M. Daëron for providing samples from Laghetto Basso and Devils Hole 2 as well as assisting with the statistical analysis. We also thank T. B. Coplen for providing material from Devils Hole. We thank C. Weise and D. M. Reeves from Ash Meadows National Wildlife Refuge for assisting us with obtaining permits to collect samples. The manuscript was significantly improved by insightful and helpful comments from two anonymous reviewers for which we are grateful.

Funding: This work was supported by NSF grants EAR-1649959 (D.L.D., J.Q., and D.D.N.), EAR-1933122 (D.L.D., J.Q., and D.D.N.), EAR-1933130 (K.W.H.), and EAR-1649986 (K.W.H.); Grants-in-Aid for Scientific Research MEXT/JSPS 23681006 (S.S.); Grants-in-Aid for Scientific Research MEXT/JSPS 15H03756 (S.S.); Grants-in-Aid for Scientific Research MEXT/JSPS 15KK0179 (S.S.); Start-up grant from the University of Arizona (A.M.); Technology and Research Initiative Fund (TRIF), Arizona Board of Regents, from the University of Arizona (K.T.); University of Arizona GPSC travel grant (N.Y.); and University of Arizona GPSC research grant (A.R.M.).

Author contributions: Conceptualization: N.Y., Z.W., D.L.D., J.Q., K.W.H., A.J.S., D.D.N., J.B.M., and S.S. Methodology: N.Y., Z.W., D.L.D., J.Q., K.W.H., A.J.S., S.S., K.T., A.R.M., and A.M. Investigation: N.Y., Z.W., D.L.D., and J.Q. Visualization: N.Y. and Z.W. Supervision: D.L.D. and J.Q. Writing—original draft: N.Y. and Z.W. Writing—review and editing: D.L.D., J.Q., K.W.H., A.J.S., D.D.N., J.B.M., K.T., A.M., and A.R.M.

Competing interests: D.D.N. and J.B.M. are employees of Aerodyne Research, which has commercialized the laser isotope monitor used in this work. D.L.D. is director of the Environmental Isotope Laboratory, which plans to process clumped isotope samples on a fee-for-service basis. All other authors declare that they have no competing interests.

Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials.

Correction (2 December 2022): The previous version of Figure 1 inadvertently included a typo, with the label “TILDAS” displaying incorrectly. This has now been corrected. In addition, the red equation in Figure 6B has been corrected. This equation previously included typos and has been corrected to match Equation 12 as intended.

Supplementary Materials

This PDF file includes:

Supplementary Text

Figs. S1 to S5

Tables S1 to S4

References

sciadv.abq0611_sm.pdf (1.8MB, pdf)

Other Supplementary Material for this manuscript includes the following:

Data S1 to S4

REFERENCES AND NOTES

  • 1.Wang Z., Schauble E. A., Eiler J. M., Equilibrium thermodynamics of multiply substituted isotopologues of molecular gases. Geochim. Cosmochim. Acta 68, 4779–4797 (2004). [Google Scholar]
  • 2.Schauble E. A., Ghosh P., Eiler J. M., Preferential formation of 13C-18O bonds in carbonate minerals, estimated using first-principles lattice dynamics. Geochim. Cosmochim. Acta 70, 2510–2529 (2006). [Google Scholar]
  • 3.Ghosh P., Adkins J., Affek H., Balta B., Guo W., Schauble E. A., Schrag D., Eiler J. M., 13C–18O bonds in carbonate minerals: A new kind of paleothermometer. Geochim. Cosmochim. Acta 70, 1439–1456 (2006). [Google Scholar]
  • 4.Dennis K. J., Affek H. P., Passey B. H., Schrag D. P., Eiler J. M., Defining an absolute reference frame for “clumped” isotope studies of CO2. Geochim. Cosmochim. Acta 75, 7117–7131 (2011). [Google Scholar]
  • 5.Bernasconi S. M., Daëron M., Bergmann K. D., Bonifacie M., Meckler A. N., Affek H. P., Anderson N., Bajnai D., Barkan E., Beverly E., Blamart D., Burgener L., Calmels D., Chaduteau C., Clog M., Davidheiser-Kroll B., Davies A., Dux F., Eiler J., Elliott B., Fetrow A. C., Fiebig J., Goldberg S., Hermoso M., Huntington K. W., Hyland E., Ingalls M., Jaggi M., John C. M., Jost A. B., Katz S., Kelson J., Kluge T., Kocken I. J., Laskar A., Leutert T. J., Liang D., Lucarelli J., Mackey T. J., Mangenot X., Meinicke N., Modestou S. E., Müller I. A., Murray S., Neary A., Packard N., Passey B. H., Pelletier E., Petersen S., Piasecki A., Schauer A., Snell K. E., Swart P. K., Tripati A., Upadhyay D., Vennemann T., Winkelstern I., Yarian D., Yoshida N., Zhang N., Ziegler M., InterCarb: A community effort to improve interlaboratory standardization of the carbonate clumped isotope thermometer using carbonate standards. Geochem. Geophys. Geosyst. 22, e2020GC009588 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Petersen S. V., Defliese W. F., Saenger C., Daëron M., Huntington K. W., John C. M., Kelson J. R., Bernasconi S. M., Colman A. S., Kluge T., Olack G. A., Schauer A. J., Bajnai D., Bonifacie M., Breitenbach S. F. M., Fiebig J., Fernandez A. B., Henkes G. A., Hodell D., Katz A., Kele S., Lohmann K. C., Passey B. H., Peral M. Y., Petrizzo D. A., Rosenheim B. E., Tripati A., Venturelli R., Young E. D., Winkelstern I. Z., Effects of improved 17O correction on interlaboratory agreement in clumped isotope calibrations, estimates of mineral-specific offsets, and temperature dependence of acid digestion fractionation. Geochem. Geophys. Geosyst. 20, 3495–3519 (2019). [Google Scholar]
  • 7.Anderson N. T., Kelson J. R., Kele S., Daëron M., Bonifacie M., Horita J., Mackey T. J., John C. M., Kluge T., Petschnig P., Jost A. B., Huntington K. W., Bernasconi S. M., Bergmann K. D., A unified clumped isotope thermometer calibration (0.5–1,100°C) using carbonate-based standardization. Geophys. Res. Lett. 48, (2021). [Google Scholar]
  • 8.Meckler A. N., Ziegler M., Millán M. I., Breitenbach S. F. M., Bernasconi S. M., Long-term performance of the Kiel carbonate device with a new correction scheme for clumped isotope measurements. Rapid Commun. Mass Spectrom. 28, 1705–1715 (2014). [DOI] [PubMed] [Google Scholar]
  • 9.Müller I. A., Fernandez A., Radke J., van Dijk J., Bowen D., Schwieters J., Bernasconi S. M., Carbonate clumped isotope analyses with the long-integration dual-inlet (LIDI) workflow: Scratching at the lower sample weight boundaries. Rapid Commun. Mass Spectrom. 31, 1057–1066 (2017). [DOI] [PubMed] [Google Scholar]
  • 10.Mackey T. J., Jost A. B., Creveling J. R., Bergmann K. D., A decrease to low carbonate clumped isotope temperatures in Cryogenian Strata. AGU Adv. 1, e2019AV000159 (2020). [Google Scholar]
  • 11.Schauer A. J., Kelson J., Saenger C., Huntington K. W., Choice of 17O correction affects clumped isotope (Δ47) values of CO2 measured with mass spectrometry. Rapid Commun. Mass Spectrom. 30, 2607–2616 (2016). [DOI] [PubMed] [Google Scholar]
  • 12.Daëron M., Blamart D., Peral M., Affek H. P., Absolute isotopic abundance ratios and the accuracy of Δ47 measurements. Chem. Geol. 442, 83–96 (2016). [Google Scholar]
  • 13.Wang Z., Nelson D. D., Dettman D. L., McManus J. B., Quade J., Huntington K. W., Schauer A. J., Sakai S., Rapid and precise analysis of carbon dioxide clumped isotopic composition by tunable infrared laser differential spectroscopy. Anal. Chem. 92, 2034–2042 (2020). [DOI] [PubMed] [Google Scholar]
  • 14.Prokhorov I., Kluge T., Janssen C., Laser absorption spectroscopy of rare and doubly substituted carbon dioxide isotopologues. Anal. Chem. 91, 15491–15499 (2019). [DOI] [PubMed] [Google Scholar]
  • 15.Rothman L. S., Gordon I. E., Barbe A., Benner D. C., Bernath P. F., Birk M., Boudon V., Brown L. R., Campargue A., Champion J. P., Chance K., Coudert L. H., Dana V., Devi V. M., Fally S., Flaud J. M., Gamache R. R., Goldman A., Jacquemart D., Kleiner I., Lacome N., Lafferty W. J., Mandin J. Y., Massie S. T., Mikhailenko S. N., Miller C. E., Moazzen-Ahmadi N., Naumenko O. V., Nikitin A. V., Orphal J., Perevalov V. I., Perrin A., Predoi-Cross A., Rinsland C. P., Rotger M., Šimečková M., Smith M. A. H., Sung K., Tashkun S. A., Tennyson J., Toth R. A., Vandaele A. C., Vander Auwera J., The HITRAN 2008 molecular spectroscopic database. J. Quant. Spectrosc. Radiat. Transf. 110, 533–572 (2009). [Google Scholar]
  • 16.E. Kerstel, Isotope ratio infrared spectrometry, in Handbook of Stable Isotope Analytical Techniques (Elsevier, 2004), pp.759–787. [Google Scholar]
  • 17.Huntington K. W., Eiler J. M., Affek H. P., Guo W., Bonifacie M., Yeung L. Y., Thiagarajan N., Passey B., Tripati A., Daëron M., Came R., Methods and limitations of “clumped” CO2 isotope (Δ47) analysis by gas-source isotope ratio mass spectrometry. J. Mass Spectrom. 44, 1318–1329 (2009). [DOI] [PubMed] [Google Scholar]
  • 18.Eiler J. M., “Clumped-isotope” geochemistry-The study of naturally-occurring, multiply-substituted isotopologues. Earth Planet. Sci. Lett. 262, 309–327 (2007). [Google Scholar]
  • 19.Daëron M., Full propagation of analytical uncertainties in Δ47 measurements. Geochem. Geophys. Geosyst. 22, (2021). [Google Scholar]
  • 20.He B., Olack G. A., Colman A. S., Pressure baseline correction and high-precision CO2 clumped-isotope (∆47) measurements in bellows and micro-volume modes. Rapid Commun. Mass Spectrom. 26, 2837–2853 (2012). [DOI] [PubMed] [Google Scholar]
  • 21.Bernasconi S. M., Hu B., Wacker U., Fiebig J., Breitenbach S. F. M., Rutz T., Background effects on Faraday collectors in gas-source mass spectrometry and implications for clumped isotope measurements. Rapid Commun. Mass Spectrom. 27, 603–612 (2013). [DOI] [PubMed] [Google Scholar]
  • 22.Kim S.-T., O’Neil J. R., Equilibrium and nonequilibrium oxygen isotope effects in synthetic carbonates. Geochim. Cosmochim. Acta 61, 3461–3475 (1997). [Google Scholar]
  • 23.Kato H., Amekawa S., Kano A., Mori T., Kuwahara Y., Quade J., Seasonal temperature changes obtained from carbonate clumped isotopes of annually laminated tufas from Japan: Discrepancy between natural and synthetic calcites. Geochim. Cosmochim. Acta 244, 548–564 (2019). [Google Scholar]
  • 24.Kele S., Breitenbach S. F. M., Capezzuoli E., Meckler A. N., Ziegler M., Millan I. M., Kluge T., Deák J., Hanselmann K., John C. M., Yan H., Liu Z., Bernasconi S. M., Temperature dependence of oxygen- and clumped isotope fractionation in carbonates: A study of travertines and tufas in the 6-95°C temperature range. Geochim. Cosmochim. Acta 168, 172–192 (2015). [Google Scholar]
  • 25.Daëron M., Drysdale R. N., Peral M., Huyghe D., Blamart D., Coplen T. B., Lartaud F., Zanchetta G., Most Earth-surface calcites precipitate out of isotopic equilibrium. Nat. Commun. 10, 429 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Fiebig J., Bajnai D., Löffler N., Methner K., Krsnik E., Mulch A., Hofmann S., Combined high-precision ∆48 and ∆47 analysis of carbonates. Chem. Geol. 522, 186–191 (2019). [Google Scholar]
  • 27.Bajnai D., Guo W., Spötl C., Coplen T. B., Methner K., Löffler N., Krsnik E., Gischler E., Hansen M., Henkel D., Price G. D., Raddatz J., Scholz D., Fiebig J., Dual clumped isotope thermometry resolves kinetic biases in carbonate formation temperatures. Nat. Commun. 11, 4005 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Stolper D. A., Eiler J. M., The kinetics of solid-state isotope-exchange reactions for clumped isotopes: A study of inorganic calcites and apatites from natural and experimental samples. Am. J. Sci. 315, 363–411 (2015). [Google Scholar]
  • 29.Eugene C., Crawford, Knut Schmidt-Nielsen, ostrich body temp paper. Am. J. Physiol. 212, 347–353 (1967). [DOI] [PubMed] [Google Scholar]
  • 30.J. Descy, P. Plisnier, B. Leporcq, S. Sténuite, S. Pirlot, J. Stimart, V. Gosselain, L. André, L. Alleman, D. Langlet, W. Vyverman, C. Cocquyt, A. de Wever, M. Stoyneva, E. Deleersnijder, J. Naithani, D. Chitamwebwa, A. Chande, I. Kimirei, B. Sekadende, S. Mwaitega, S. Muhoza, D. Sinyenza, L. Makasa, C. Lukwessa, I. Zulu, H. Phiri CLIMLAKE, “CLIMLAKE Climate Variability As Recorded In Lake Tanganyika Final Report (2001–2005)” (2005).
  • 31.Coplen T. B., Calibration of the calcite-water oxygen-isotope geothermometer at Devils Hole, Nevada, a natural laboratory. Geochim. Cosmochim. Acta 71, 3948–3957 (2007). [Google Scholar]
  • 32.Shanahan T. M., Pigati J. S., Dettman D. L., Quade J., Isotopic variability in the aragonite shells of freshwater gastropods living in springs with nearly constant temperature and isotopic composition. Geochim. Cosmochim. Acta 69, 3949–3966 (2005). [Google Scholar]
  • 33.Reynolds C. E., Richey J. N., Fehrenbacher J. S., Rosenheim B. E., Spero H. J., Environmental controls on the geochemistry of Globorotalia truncatulinoides in the Gulf of Mexico: Implications for paleoceanographic reconstructions. Mar. Micropaleontol. 142, 92–104 (2018). [Google Scholar]
  • 34.Spear J. W., Poore R. Z., Quinn T. M., Globorotalia truncatulinoides (dextral) Mg/Ca as a proxy for Gulf of Mexico winter mixed-layer temperature: Evidence from a sediment trap in the northern Gulf of Mexico. Mar. Micropaleontol. 80, 53–61 (2011). [Google Scholar]
  • 35.Thirumalai K., Quinn T. M., Okumura Y., Richey J. N., Partin J. W., Poore R. Z., Moreno-Chamarro E., Pronounced centennial-scale Atlantic Ocean climate variability correlated with Western Hemisphere hydroclimate. Nat. Commun. 9, 392 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.York D., Evensen N. M., Martínez M. L., de Basabe Delgado J., Unified equations for the slope, intercept, and standard errors of the best straight line. Am. J. Phys. 72, 367–375 (2004). [Google Scholar]
  • 37.Thirumalai K., Singh A., Ramesh R., A MATLAB™ code to perform weighted linear regression with (correlated or uncorrelated) errors in bivariate data. J. Geol. Soc. India 77, 377–380 (2011). [Google Scholar]
  • 38.Brand W. A., Assonov S. S., Coplen T. B., Correction for the 17O interference in δ(13C) measurements when analyzing CO2 with stable isotope mass spectrometry (IUPAC Technical Report). Pure Appl. Chem. 82, 1719–1733 (2010). [Google Scholar]
  • 39.McManus J. B., Nelson D. D., Zahniser M. S., Design and performance of a dual-laser instrument for multiple isotopologues of carbon dioxide and water. Opt. Express 23, 6569–6586 (2015). [DOI] [PubMed] [Google Scholar]
  • 40.Levitt N. P., Eiler J. M., Romanek C. S., Beard B. L., Xu H., Johnson C. M., Near equilibrium 13C—18O bonding during inorganic calcite precipitation under chemo-stat conditions. Geochem. Geophys. Geosyst. 19, 901–920 (2018). [Google Scholar]
  • 41.Blue C. R., Giuffre A., Mergelsberg S., Han N., de Yoreo J. J., Dove P. M., Chemical and physical controls on the transformation of amorphous calcium carbonate into crystalline CaCO3 polymorphs. Geochim. Cosmochim. Acta 196, 179–196 (2017). [Google Scholar]
  • 42.Saldi G. D., Jordan G., Schott J., Oelkers E. H., Magnesite growth rates as a function of temperature and saturation state. Geochim. Cosmochim. Acta 73, 5646–5657 (2009). [Google Scholar]
  • 43.D. L. Parkhurst, C. A. J. Appelo, User’s guide to PHREEQC (version 2)—A computer program for speciation, batch-reaction, one-dimensional transport, and inverse geochemical calculations (1999); www.xs4all.nl/~appt/index.html.
  • 44.Griffith D. W. T., Deutscher N. M., Caldow C., Kettlewell G., Riggenbach M., Hammer S., A Fourier transform infrared trace gas and isotope analyser for atmospheric applications. Atmos. Meas. Tech. 5, 2481–2498 (2012). [Google Scholar]
  • 45.Kim S. T., Coplen T. B., Horita J., Normalization of stable isotope data for carbonate minerals: Implementation of IUPAC guidelines. Geochim. Cosmochim. Acta 158, 276–289 (2015). [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary Text

Figs. S1 to S5

Tables S1 to S4

References

sciadv.abq0611_sm.pdf (1.8MB, pdf)

Data S1 to S4


Articles from Science Advances are provided here courtesy of American Association for the Advancement of Science

RESOURCES