Skip to main content
Wiley Open Access Collection logoLink to Wiley Open Access Collection
. 2022 Apr 19;36(12):e9301. doi: 10.1002/rcm.9301

Temperature dependence of isotopic fractionation in the CO2‐O2 isotope exchange reaction

Getachew Agmuas Adnew 1,, Evelyn Workman 1, Christof Janssen 2, Thomas Röckmann 1
PMCID: PMC9285609  PMID: 35318757

Abstract

Rationale

Oxygen isotope exchange between O2 and CO2 in the presence of heated platinum (Pt) is an established technique for determining the δ 17O value of CO2. However, there is not yet a consensus on the associated fractionation factors at the steady state.

Methods

We determined experimentally the steady‐state α 17 and α 18 fractionation factors for Pt‐catalyzed CO2‐O2 oxygen isotope exchange at temperatures ranging from 500 to 1200°C. For comparison, the theoretical α 18 equilibrium exchange values reported by Richet et al. (1997) have been updated using the direct sum method for CO2 and the corresponding α 17 values were determined. Finally, we examined whether the steady‐state fractionation factors depend on the isotopic composition of the reactants, by using CO2 and O2 differing in δ18O value from −66 ‰ to +4 ‰.

Results

The experimentally determined steady‐state fractionation factors α 17 and α 18 are lower than those obtained from the updated theoretical calculations (of CO2‐O2 isotope exchange under equilibrium conditions) by 0.0024 ± 0.0001 and 0.0048 ± 0.0002, respectively. The offset is not due to scale incompatibilities between isotope measurements of O2 and CO2 nor to the neglect of non‐Born‐Oppenheimer effects in the calculations. There is a crossover temperature at which enrichment in the minor isotopes switches from CO2 to O2. The direct sum evaluation yields a θ value of ~0.54, i.e. higher than the canonical range maximum for a mass‐dependent fractionation process.

Conclusions

Updated theoretical values of α 18 for equilibrium isotope exchange are lower than those derived from previous work by Richet et al. (1997). The direct sum evaluation for CO2 yields θ values higher than the canonical range maximum for mass‐dependent fractionation processes. This demonstrates the need to include anharmonic effects in the calculation and definition of mass‐dependent fractionation processes for poly‐atomic molecules. The discrepancy between the theory and the experimental α 17 and α 18 values may be due to thermal diffusion associated with the temperature gradient in the reactor.

1. INTRODUCTION

Measurements of Δ′17O (defined in Equation 1) of oxygen‐containing molecules can provide insight into, for example, the carbon cycle and its link to the hydrological cycle 1 , 2 , 3 , 4 , 5 , 6 , 7 , 8 , 9 , 10 and are also used to reconstruct paleoclimate. 10 , 11 , 12 , 13 , 14 The Δ′17O of CO2 is a promising tracer to study the CO2 exchange flux between the biosphere/hydrosphere and the atmosphere. 3 , 5 , 6 , 15 However, due to the mass interference of 13C16O16O and 12C17O16O, it is difficult to measure δ 17O of CO2 with existing isotope ratio mass spectrometers, because the required mass resolving power of about 55,000 (Δm/m) is a challenge for even the highest resolution instruments. 16 High precision and accuracy are needed for measurements of both δ 18O and δ 17O in order to capture the small, but non‐zero deviations to use Δ′17O as a tracer. Since the measurement of δ 17O directly from CO2 with isotope ratio mass spectrometry (IRMS) is impossible due to the interference of δ 13C, except from the fragment ions formed in the ion source, 16 indirect methods have been developed that enable the measurement of δ 17O of CO2. These techniques are either based on isotope exchange or chemical conversion of CO2 and are described in more detail in Adnew et al 16 and Hofmann and Pack. 17

Δ17O=lnδ17O+10.528lnδ18O+1 (1)

The CO2‐O2 exchange method was first developed by Mahata et al 18 and has been established in several laboratories, including our own. 11 , 12 , 16 , 19 Recently, the method has been extended to measure the Δ′17O of water. 20 The CO2‐O2 exchange method to measure Δ′17O of CO2 is less labor‐intensive compared with other conversion/exchange methods and is capable of an impressive precision of < 0.01 ‰. Such precision allows the subtle seasonal Δ′17O variability of tropospheric CO2 to be quantified, with reported seasonal amplitudes between 0.02 ‰ and 0.13 ‰. 2 , 6 , 21

The CO2‐O2 exchange method is based on facilitating complete oxygen isotope exchange between CO2 and O2 at high temperature in the presence of a platinum catalyst. This way, the oxygen isotopic composition of the CO2 is imprinted on O2 and the 17O content of the CO2 sample can be calculated from measurement of the isotopic composition of O2 before and after the isotope exchange. Important parameters in the calculation of the Δ′17O of CO2 are the molar ratio of CO2 and O2 and the isotopic fractionation factors (α).

αICO2/O2=RIfCO2RIfO2=δIOf+1CO2δIOf+1O2 (2)

The index I is a placeholder for either 17 or 18. The fractionation factors α 17(CO2/O2) and α 18(CO2/O2) determine the distribution of the heavy isotopes between O2 and CO2 at isotopic steady state.

The CO2‐O2 exchange technique has been used in different laboratories to study the carbon cycle 1 , 2 , 3 , 4 , 6 , 22 , 23 and precise measurement of Δ′17O is used to correct for biases in clumped isotope measurements of CO2. 24 , 25 , 26 , 27 However, the fractionation factor in CO2‐O2 isotope exchange is not well established, and different values have been reported by different laboratories.

One explanation for these differences may be variations in the effective isotope exchange temperature, as the fractionation factor is expected to be temperature dependent. Furthermore, the mechanism of the CO2‐O2 isotope exchange in the presence of a platinum catalyst is unknown. 16 , 18 , 19 , 26 , 28 , 29 In this study we determined α 17(CO2/O2) and α 18(CO2/O2) for the CO2‐O2 exchange method over a temperature range from 500°C to 1200°C. We investigated whether there is a dependency on the initial oxygen isotopic composition of CO2 and O2. This also addresses a potential effect of an isotope scale difference between CO2 and O2. We revisit and update theoretical calculations of the isotope equilibrium by Richet et al. (1997) 30 and calculated the three‐isotope exponent (θ) of the CO2‐O2 equilibrium isotope exchange for a temperature range of 0°C to 1200°C.

2. MATERIALS AND METHODS

2.1. Theoretical calculations of CO2‐O2 isotope exchange at equilibrium

Fractionation factors 17 α and 18 α for the CO2‐O2 isotope exchange can be calculated from statistically corrected equilibrium constants of O‐atom exchange reactions with CO2 and O2, respectively. The latter are conveniently designated as β‐factors and were originally introduced as fractionation factors between the fully substituted molecule and O‐atom (I β A  = ([IO]/[16O]) A /([IO]/[16O])Atomic−O, where A is either CO2 or O2). 30 , 31 , 32 , 33 Since in the measurements of the equilibrium constants multi‐substituted species have been neglected, we define quantities α and β slightly differently:

αI=QCO16OIQCO216QO216QO16OI=QCO16OIQO16QCO216QOI/QO16OIQO16QO216QOI=βCO2IβO2I, (3)

where Q are the total partition functions. For the calculation of these exchange equilibria, molecular partition functions were divided into internal and translational parts. Total internal partition functions and ratios of diatomic molecules were calculated using the formalism of Urey 34 and Bigeleisen and Mayer, 35 including anharmonic corrections and zero‐point energy expressions based on Dunham's 36 expression of energy levels (see Richet et al. (1997) 30 ). Molecular constants of the 16O2 species were taken from Huber and Herzberg 37 in combination with recent atomic masses 38 and those of the isotopic variants were calculated using the standard mass dependence of Dunham coefficients. This choice of molecular constants is compatible with a recent recommendation 39 and an agreement to better than five significant digits is obtained across the temperature range. However, our current results are slightly different from a previously published parametrization 40 – the less than 0.2 ‰ difference being mainly due to the identification and removal of an error in the calculation of a correction term.

Unfortunately, application of Dunham's approach to larger molecules is not straightforward, because the isotope scaling of anharmonic vibrational constants is neither theoretically well constrained nor experimentally known, and the constant zero‐point energy offset G 0 is not known either. 30 We have therefore calculated the total internal partition functions of the 16O‐, 17O‐, and 18O‐containing isotopologues of IO12C16O as direct sums of overall energy levels obtained from the spectroscopically adjusted ab‐initio line list calculations of Huang et al 41 , 42 (see Prokhorov et al 43 for more details).

Our new calculations were verified by comparing atom‐diatom isotope exchange equilibria (β‐values) with the tables of Richet et al. (1997) 30 using several isotopes and diatomic molecules (13C exchange in CO and CS, 15N exchange in N2, and 18O exchange in CO and O2). These results agreed across the temperature range between 0°C and 1200°C, to better than 0.7 ‰ for all molecules except for O2, where our values were higher by 4 ‰ at 0°C and 1 ‰ at 1200°C (see Table S1, supporting information). We have discovered a subtle difference in the molecular constants used in this study and that of Richet et al. (1997) 30 but such slight differences cannot account for the observed discrepancy. Interestingly, Richet et al. (1997) 30 report the same 4 ‰ discrepancy when comparing their 25°C results with the calculations of Urey 32 (see Table 16 in Richet et al. (1997) 30 ), while the β‐values for O‐, C‐, or N‐exchange with other diatomic molecules always agree with the work of Urey to within 1 ‰. It needs be pointed out that the discrepancy is not due to our slightly different definition of β and our neglect of excess factors, since it even increases if we calculate β‐values from doubly substituted O2. This indicates that an error must have occurred in the O2 calculations of Richet et al. (1997). 30

When comparing our atom‐CO2 exchange equilibria with Richet et al. (1997) 30 the agreement is better, but still worse compared with diatomic gases when O2 is excluded from the comparison. For the 18O exchange, our β‐values are 2 ‰ higher at 0°C and less than 1 ‰ for relevant temperatures above 400°C (see supporting information). For 13C exchange, our values at 0°C are 1.3 ‰ higher and the discrepancy becomes already smaller than 1 ‰ at temperatures above 100°C.

The comparison of α 18(CO2/O2) between our calculation and that of Richet et al. (1997) 30 for the temperature range between 0°C and 1200°C is also shown in Table S2 (supporting information). Our α18(CO2/O2) value agrees with that reported by Richet et al. (1997) 30 to within 1 ‰, for temperatures higher than 450°C. However, at low temperature the difference is higher than 1 ‰ (for instance 2.2 ‰ at 0°C; see Table S2, supporting information). This is because our β‐factors for both O2 and CO2 are different from the values reported by Richet et al. (1997) 30 the difference being largest in O2, where we suspect the calculation of Richet et al. (1997) 30 is erroneous (see Table S1, supporting information).

We have compared partition function ratios Q(13CO2)/Q(12CO2) of 16O‐ and 18O‐containing isotopologues also with results from path integral Monte Carlo calculations of Webb and Miller 44 over the 300 to 600 K range and found deviations of less than 0.5 ‰ and 0.3 ‰ at 300 K and 600 K, respectively. This confirms that CO2‐related uncertainties of 18 α should be smaller than 0.3 ‰ for most of the temperatures in this study. Higher electronic states of both molecules have been neglected in the calculations. At transition energies of 5.7 eV (11A1 CO2 45 ) and 0.98 eV (a1g O2 37 ), the population of electronically excited CO2 is completely negligible and the O2(1 ) ground state is only marginally populated (by less than 8 parts in 105), even at the highest temperature. The effect on isotope ratios is much smaller because geometric changes in excited oxygen isotopes are small, and vibrational and rotational constants of the excited state are similar to the ground state values (changing by 7 % at most). 37

2.2. CO2‐O2 exchange system

A detailed description of the CO2‐O2 exchange method established at Utrecht University is given elsewhere. 1 , 16 In brief, equal amounts of CO2 and O2 were allowed to equilibrate in a quartz reactor in the presence of platinum sponge as a catalyst (purity ≥ 99.9 %, Sigma Aldrich, USA) at the bottom of the reactor. The CO2‐O2 exchange experiments were performed at temperatures between 500 and 1200°C. The temporal evolution of the isotope exchange was determined at 650°C, 750°C, and 850°C by varying the duration of the CO2‐O2 exchange. After the exchange reaction, CO2 and O2 were separated cryogenically and the O2 collected at liquid nitrogen temperature on molecular sieve 5 Å pellets (3 pellets, 5 mm long, 1.8 mm o.d., Sigma Aldrich, USA). The isotope measurements were performed on a Delta V Plus isotope ratio mass spectrometer (Thermo Scientific, Germany) in dual inlet mode. A schematic of the exchange reactor is shown in Figure 1. For comparison, the reactor that was used in Adnew et al. (2019) 16 is shown in Figure S1 (supporting information). The main difference between the current setup and the one used in the previous study 16 is an electropolished stainless steel extension.

FIGURE 1.

FIGURE 1

Geometry and dimensions of the CO2‐O2 reactor

2.3. Samples

Three commercial O2 gases with different isotope compositions were used for the measurements (Table 1). The working reference gas (GO2‐1) has δ 17O = 9.235 ± 0.011 ‰ and δ 18O = 18.514 ± 0.011 ‰ (Table 1), determined by Eugeni Barkan, Hebrew University of Jerusalem. For CO2 we used two gases from high‐pressure cylinders (GCO2‐3 and GCO2‐4) and two additional gases (GCO2‐1 and GCO2‐2) that were prepared by combusting a graphite rod with two isotopically different O2 gases 16 , 46 , 47 , 48 , 49 (GO2‐1 and GO2‐2, respectively). The isotopic compositions of the CO2 and O2 used in this study are shown in Table 1. The experimental procedure for the combustion experiment is described in detail elsewhere. 16 In summary, the graphite rod (99.9995% purity, 3.05 mm × 20 mm, Alfa Aesar, Part No: 40765, Thermo Fisher Scientific) was wrapped in a sheet of platinum foil and platinum wire and placed inside a quartz reactor. The combustion with O2 was performed at 800°C. Before starting the experiment, the graphite rod was cleaned for more than 24 h by heating at 1000°C in high vacuum. After each conversion experiment, the graphite rod was again cleaned by heating at 1000°C for at least 1 h.

TABLE 1.

Overview of names, suppliers, and isotopic compositions of the CO2 and O2 working standards used in this study. All the CO2 gases used had a purity of > 99.995 % except for the ones that were produced in the laboratory by combustion of a graphite rod. O2 gases had a purity of > 99.998 %. All isotopic values are given in ‰, relative to VPDB for δ13C and relative to VSMOW for δ17O and δ18O. The errors are reported as one standard deviation (σ)

CO2 gases used in this study
Name Source/supplier δ 13C δ 18O Δ′17O
GCO2‐1 Combustion −26.160 ± 0.030 17.921 ± 0.042 −0.465 ± 0.018
GCO2‐2 Combustion −26.154 ± 0.059 −38.205 ± 0.051 −0.518 ± 0.016
GCO2‐3 Linde Gas, The Netherlands −31.665 ± 0.005 34.655 ± 0.011 −0.242 ± 0.003
GCO2‐4 Air Products, Germany −10.332 ± 0.005 30.082 ± 0.011 −0.169 ± 0.002
O2 gases used in this study
Name Source/supplier δ17O δ18O Δ′17O
GO2‐1 Air Products, The Netherlands 9.235 ± 0.011 18.514 ± 0.011 −0.494 ± 0.011
GO2‐2 Air Liquide, The Netherlands −20.734 ± 0.017 −37.862 ± 0.012 −0.573 ± 0.017
GO2‐3 Linde Gas, The Netherlands 7.040 ± 0.021 14.095 ± 0.011 −0.375 ± 0.020

3. RESULTS

The differences in isotopic composition of the two working standards (GCO2‐3 and GCO2‐4) are 4.440 ± 0.015 ‰ for δ 18O and −21.556 ± 0.007 ‰ for δ 13C. respectively (Table 1). In our laboratory, we measured relative differences of 4.479 ± 0.003 ‰ for δ 18O and −21.448 ± 0.001 ‰ for δ 13C, respectively, from six replicates. For each measurement new gas was introduced to the bellows. Similarly, for GO2‐1 and GO2‐2, the relative difference was −29.690 (−29.746) ‰ for δ 17O and −55.456 (−55.432) ‰ for δ 18O, respectively. The values in parentheses are the assigned differences based on the measurements at Hebrew University (E. Barkan). Overall, the assigned and measured values agree well.

3.1. CO2‐O2 isotope steady state

Figure 2 shows the temporal evolution of the CO2‐O2 isotope exchange for temperatures of 650°C, 750°C, and 850°C. The exchange proceeds faster at higher temperatures, as expected.

FIGURE 2.

FIGURE 2

Temporal evolution of δ 17O (O2) (A) and δ 18O (O2) (B) during CO2‐O2 exchange on a platinum sponge catalyst in a quartz reactor, for reaction temperatures of 650°C, 750°C, and 850°C. Δδ17O(O2) and Δδ18O (O2) are the differences in δ17O and δ18O, respectively, between the initial and final isotopic compositions of O2 [Color figure can be viewed at wileyonlinelibrary.com]

The data sets at each of the three reaction temperatures are fitted to an exponential curve of the form y = A + B × e −k × t , where y is the isotopic composition (δ17O or δ18O) of O2 at time t, A is the isotopic composition at long time scales (Figure 2), A + B is the isotopic composition of O2 at t = 0, k is the rate constant of the reaction, and t is the reaction time. In our experiments we defined steady state as having been achieved when k determined from the fit for each temperature e kt < 10−6, i.e. t = 13.816/k. From these fits, the respective steady state values of δ 17Of and δ 18Of of the O2 are 2.684 ‰ and 5.032 ‰ at 650°C; 3.018 ‰ and 5.648 ‰ at 750°C; and 3.278 ‰ and 6.100 ‰ at 850°C. The enrichments of 17O and 18O in the O2 at steady state increase with reaction temperature (Figures 2 and 3A).

FIGURE 3.

FIGURE 3

A, δ 18O(O2) and δ 18O(CO2) after CO2‐O2 exchange at isotope exchange steady state for reaction temperatures ranging from 500°C to 1200°C. B, The corresponding fractionation factors α 17(CO2/O2) and α 18(CO2/O2). The vertical dashed line shows the temperature at which α 17(CO2/O2) and α 18(CO2/O2) cross the value of 1 (horizontal dashed line). In the fit function, T is the temperature in °C [Color figure can be viewed at wileyonlinelibrary.com]

Figure 3A shows the steady state isotopic composition δ 18Of (O2) and δ 18Of (CO2) for reaction temperatures ranging from 500 to 1200°C. In Figure 3B the fractionation factors α 17(CO2/O2) and α 18(CO2/O2) derived from these steady state enrichments are shown. As can be seen, α 17 and α 18 decrease with increasing reaction temperature. The exchange reactions were carried out using gases GCO2‐4 (δ 18O = 30.082 ‰) and GO2‐3 (δ 18O = 14.095 ‰), see Table 1. Figure 3 shows that below 810°C, δ 18Of (CO2) > δ 18Of (O2) and α 18(CO2/O2) > 1; thus there is a slight preference for the CO2 to be relatively enriched in 17O and 18O. At higher temperatures, this preference is reversed, with α 18 < 1, i.e. δ 18Of (CO2) < δ 18Of (O2).

3.2. Dependence of α 17(CO2/O2) and α 18(CO2/O2) on the relative difference between the initial oxygen isotopic compositions of CO2 and O2

The isotopic steady state should be a property of the reaction system and independent of the isotopic composition of the initial reagents. To confirm this, we used O2 and CO2 with a wide range of isotopic compositions, with a δ 18O difference of the initial O2 and CO2 by up to −65.96 ‰ in exchange experiments at 650°C, 750°C, and 850°C.

Figure 4 demonstrates that the same values for α 17(CO2/O2) and α 18(CO2/O2) are reached in experiments where the isotopic composition of the starting gases is very different, confirming that α 17(CO2/O2) and α 18(CO2/O2) values are independent of the initial isotopic composition of CO2 and O2.

FIGURE 4.

FIGURE 4

Dependence of α 17(CO2/O2) (blue symbols) and α 18(CO2/O2) (black symbols) on the initial isotopic composition of CO2 and O2 (see legend for the relative difference in δ 18O between CO2 and O2). The rectangle and star symbols are for CO2 samples produced by the combustion of a graphite rod with O2. Δδ 18O is the initial δ 18O difference between CO2 and O2 [Color figure can be viewed at wileyonlinelibrary.com]

3.3. Comparison with theoretical calculations

Figure 5A shows that α 17(CO2/O2) and α 18(CO2/O2) depend on temperature. This holds for both the experimental steady state results and the theoretically calculated thermodynamic equilibrium values, with the shapes of the temperature dependence being similar between theory and experiment. Interestingly, the calculated results are higher than the experimental data. As shown in Figure 6, the offset between the experimental values and theoretical equilibrium calculation is not constant but has a slight temperature dependence.

FIGURE 5.

FIGURE 5

A, Dependence of measured α 17(CO2/O2) and α 18(CO2/O2) on temperature, together with data from theoretical equilibrium calculations. In B) the temperature scale has been adjusted to that of the experiments and the theoretically derived equilibrium values for α 17(CO2/O2) and α 18(CO2/O2) (solid lines) have been corrected for the respective mean offsets in α 17(CO2/O2) and α 18(CO2/O2) (dashed lines) [Color figure can be viewed at wileyonlinelibrary.com]

FIGURE 6.

FIGURE 6

Temperature dependence of the experimental α 17(CO2/O2) and α 18(CO2/O2) offsets from the theoretical equilibrium values

The average offset between the theoretically calculated and the experimental values is 0.0024 ± 0.001 and 0.0048 ± 0.002 for α 17(CO2/O2) and α 18(CO2/O2), respectively. This offset is very close to mass‐dependent fractionation with a triple oxygen isotope fractionation (θ) value of 0.5. For the CO2‐O2 exchange experiment, θ is calculated as shown in Equation 4.

θCO2/O2=lnα17CO2/O2lnα18CO2/O2 (4)

In Figure 5B we have shifted the theoretical α 17(CO2/O2) and α 18(CO2/O2) by the average offset from the experimental value, which results in a very good agreement with the experimentally determined values. The difference between the theoretical and measured α 17(CO2/O2) and α 18(CO2/O2) values show a slight increase with temperature (Figure 6). This trend diminishes at temperatures above 800°C.

According to the theoretical calculations, the fractionation at higher temperatures should approach unity. However, as mentioned above, experimentally we observe a fractionation of less than unity at higher temperatures, where the δ 18O of the exchanged O2 is higher than the δ 18O of CO2 (Figures 3 and 4).

3.3.1. Scale difference between O2 and CO2

To check if the discrepancy of α 17(CO2/O2) and α 18(CO2/O2) between the experimental results and theoretical equilibrium estimates is due to the scale difference between the CO2 and O2, we used CO2 formed from combustion of a graphite rod as described in section 2.3. The difference in δ 18O between GO2‐3 and GO2‐1 was −55.335 ‰. After combustion of the graphite rod with both GO2‐1 and GO2‐2, the relative difference between the produced GCO2‐1 and GCO2‐2 was −55.169 ‰. A potential scale difference is thus of the order 0.2 ‰, but these small differences between δ 18O of the oxygen used for combustion and the δ 18O of the resulting CO2 might also be due to incomplete combustion and/or CO formation. 16 , 46 , 47 , 48 , 49 After assessing the potential scale discrepancies, we performed the same equilibration experiments with these gases at a reaction temperature of 750°C. As shown in Figure 4, the discrepancy between the theoretically estimated equilibrium fractionation in δ 17O and δ 18O, and the experimental results after isotopic exchange between CO2 and O2 at steady state, is similar for all combinations of CO2 and O2, including the gases where scale discrepancies were specifically assessed. This provides evidence that the offset is not due to scale differences between the O2 and CO2 isotope scales.

3.4. Dependence of θ (CO2/O2) on CO2‐O2 equilibration temperature

As shown in Figure 7, θ (CO2/O2) is dependent on the CO2‐O2 equilibration temperature. θ (CO2/O2) takes on extreme values close to the crossover (−∞ to ∞). In addition, for α values close to 1, small experimental errors will lead to large variability in θ (CO2/O2). The theoretical calculation does not show any crossover temperature, unlike the experimental values. Due to the crossover in the experiments between 750 and 800°C, θ values are largely outside the canonical range between 0.5 and 0.5305. 31 , 50 , 51 , 52 , 53 In contrast, the calculated θ (CO2/O2) value systematically exceeds the canonical upper limit θ HTL = 0.5305 and remains close to 0.54 across the temperature range between 0°C and 1200°C. We suspect that the drop in θ (CO2/O2) at temperatures >1250°C in Figure 7 is an artefact, likely to be due to not including some high‐energy states in the calculations. 43 , 54 , 55 This concerns mostly the rarest (17O containing) isotopologue, on which the least experimental information is available.

FIGURE 7.

FIGURE 7

Dependence of θ (CO2/O2) on reaction temperature and α 18(CO2/O2) value for experimental data (A and C) and theoretical calculation (B and D). In (A) and (C), the dashed line is the θ (CO2/O2) value calculated from the fit function of Figure 3: α 18(CO2/O2) = 0.024 exp(−0.002 T) + 0.996; α 17(CO2/O2) = 0.013exp(−0.002 T) + 0.998

4. DISCUSSION

4.1. Reaction rate constant of the CO2‐O2 isotope exchange reaction

The rate for the CO2‐O2 isotope exchange reaction in the presence of a platinum catalyst depends on the exchange temperature. As shown in Figure 2, isotope steady state between CO2 and O2 was reached in less than 1 h at reaction temperatures of 650°C, 750°C, and 850°C. However, at 500°C, the results indicated that the isotopic steady state was not reached within 2 h (α 17(CO2/O2) = 1.0010 and α 18(CO2/O2) = 1.0187), see Table S3 (supporting information). Mahata and co‐authors 28 reported an increase in the δ 18O value of oxygen with temperature which they attributed to an enhancement in the catalytic activity of the platinum. Our data on the temperature dependence of α 18(CO2/O2) show (after accounting for an offset, discussed later) very good agreement with the theoretical equilibrium calculations over the entire temperature range of 500 to 1200°C. This demonstrates that the increase in the δ 18O value of the oxygen is not due to the efficiency of the catalyst but reflects the thermodynamically expected behavior. However, we did observe a decrease in the efficiency of the platinum sponge catalyst after it was exposed to a higher temperature. For example, after exposure to 1200°C, it took a long time to reach isotopic steady state at a lower temperature (e.g. at 650°C), which most likely indicates some kind of sintering of the platinum sponge, leading to a reduction in active surfaces and thus a decrease in catalytic efficiency. The efficiency of the catalyst only affects the duration required to attain CO2‐O2 isotope exchange, not the final isotopic composition.

4.2. α 17(CO2/O2) and α 18(CO2/O2) of the CO2‐O2 isotope exchange reaction

As described in section 3.3, the experimental values for α 17(CO2/O2) and α 18(CO2/O2) are lower than the corresponding theoretical equilibrium estimates, and they cross from values above 1 to values below 1 at higher temperatures, in contrast to the theoretical equilibrium calculations. One shortcoming of the current work is that the calculations of α values involve the assumption of the Born‐Oppenheimer approximation in the analysis of spectral data leading to the partition functions in Equation 2. However, this assumption is only approximate and it is possible that neglecting these effects could at least partly explain the difference between experiment and theory. Based on the work of Born and Huang, 56 Zhang and Liu 57 give an expression for the correction of a partition function ratio for a heavy and light isotope pair, which essentially implies that there is an additional constant shift (ΔE) in the energy levels between the heavy and the light molecules of an isotope pair, leading to an additional term exp(−ΔE/k B T) in the calculation of the partition function ratios and the equilibrium constant (where k B is Boltzmann's constant). These effects are strongly temperature dependent and vanish at high temperatures. Such behavior is not compatible with our observation of a temperature‐independent offset and we consequently exclude the notation non‐Born‐Oppenheimer effects could explain the observed discrepancy.

Mahata and co‐authors 18 also reported α 18(CO2/O2) < 1, ranging from 0.998128 to 0.99463, even at 680°C. Table 2 shows that all recently published α 17(CO2/O2) and α 18(CO2/O2) values are lower compared with the theoretically calculated equilibrium value for CO2‐O2 exchange temperatures of 600°C to 800°C. The presence of the platinum sponge, the geometry of the experimental setups, and the temperature gradient between the cold and hot zones of the reactor were postulated as possible sources of the discrepancy between the measured fractionation value and the theoretically estimated thermodynamic equilibrium value.

TABLE 2.

Theoretically calculated equilibrium values (this study) and experimentally determined values of α 17(CO2/O2) and α 18(CO2/O2) from this study (*) and from recent publications. All the experimentally determined values are lower than the theoretically calculated equilibrium ones

Temp. [°C] Theoretical values (this study) Experimental values Ref.
α 17(CO2/O2) α 18(CO2/O2) α 17(CO2/O2) α 18(CO2/O2) θ (CO2/O2)
600 1.0036 1.0067 1.0031–1.00785 30
670 1.0032 1.0059 0.99977 0.9990 0.229911 29
680 1.0031 1.0058 0.998 to 0.995 19
700 1.0030 1.0056 0.9997–1.00414 30
750 1.0028 1.0052 1.00135 1.00227 0.594987 20
1.00125 1.00215 0.581657 12
1.00082 1.00141 0.581732 13
1.000666 1.000998 0.667445 17
1.00048 1.00051 0.941191 *
800 1.0026 1.0048 0.9998–1.001269 30
0.9990 0.9977 0.4345 28
1.0002065 0.999988 −17.2065 *

Based on the results of the O2‐CO2 conversion experiments, the discrepancy between the experimental and theoretically calculated equilibrium values of α 18(CO2/O2) and α 17(CO2/O2) is not due to an isotopic scale difference between O2 and CO2. In addition, α 18(CO2/O2) and α 17(CO2/O2) values are independent of the initial oxygen isotope composition of the CO2 and O2 and also do not depend on the relative difference in the oxygen isotope composition between the initial CO2 and O2. Thus, the discrepancy must have other causes.

Oxygen isotope exchange between CO2 or O2 and the quartz reactor tube has been shown to be negligible in our experiments at 750°C. 16 , 19 We did not investigate whether it occurs at higher temperatures, but Barkan and Luz 46 reported that isotope exchange between oxygen and quartz at temperatures of up to 950°C is negligible. Similarly, Saeger et al 26 reported no significant exchange between quartz and O2 nor between quartz and CO2 at 800°C. The effect of oxygen isotope exchange with quartz will be higher if the δ 18O value of the quartz is very different from that of the CO2 or O2. Quartz glass tubing is mostly made from silica sand having δ 18O values ranging from 10 ‰ to 20 ‰, 58 which is relatively close to the δ 18O value of the O2 used for characterizing the CO2‐O2 exchange reaction in this study (GO2‐3). The potential for isotope exchange should depend on the relative difference in oxygen isotopic composition between the quartz, CO2, and O2. However, even if there is isotope exchange with quartz, it only affects the δ 17O and δ 18O values of CO2 and O2 after exchange, not the α 17(CO2/O2) and α 18(CO2/O2) values.

A possible explanation for the discrepancy between the calculation and the experimental results is isotopic separation due to thermal diffusion. 59 , 60 , 61 , 62 , 63 , 64 In this study, one end of the reactor is positioned inside an oven at a specific temperature (500°C to 1200°C), whereas the other end is at room temperature (see Figure 1). The hot zone of the reactor is only 35 cm3 in volume and approximately half that is outside the oven (Figure 1). Due to thermal diffusion, the heavier isotopes will accumulate in the cold zone of the reactor compared to the hot zone. 59 , 60 , 61 , 65 , 66 The relative separation of the isotopologues is proportional to their mass difference. The thermal diffusion factor of the heavy isotopologue relative to the light isotopologue is (α 21) = α o(m 2 − m 1)/(m 1 + m 2) where m 1 and m 2 are the molecular masses of the light and heavy isotopologues, respectively, and α o is the reduced thermal diffusion factor which depends only on temperature and the intermolecular force between isotopes. 60 , 67 A detailed description of thermal diffusion and the thermal diffusion factor is presented in the literature 60 , 67 and references cited therein. The isotope separation between the cold and hot zones for O2 isotopologues is higher than for CO2 isotopologues. 60 , 65 Thus, the O2 in the cold zone of the reactor will be more enriched in heavy isotopes compared with CO2, resulting in α 17(CO2/O2) and α 18(CO2/O2) being lower than as calculated for thermodynamic equilibrium.

The effect of thermal diffusion should depend on the geometry and the temperature gradient between the cold and hot zones of the reactor. Horizontally oriented reactors are less convective and could result in larger isotope fractionation compared to a vertically oriented reactor. Mahata and co‐authors 28 used a horizontal reactor, and in fact the results from their study show a higher discrepancy with respect to the theoretical calculations (3.4 ‰ and 6.9 ‰ for α 17(CO2/O2) and α 18(CO2/O2), respectively; Table 2). If thermal diffusion is indeed responsible for the discrepancies between the theoretical equilibrium values and the experimental results, the offset is also expected to depend on the relative volumes of the cold and hot zones of the reactor. Indeed, at 750°C, Adnew et al 16 reported a slightly higher (thus closer to the theoretical value) value of α 18(CO2/O2), 1.000998 vs. 1.00051 (see Table 1), using a similar experimental setup except that in the previous study the volume of the cold zone of the reactor was smaller and without a stainless‐steel extension (see Figure 1 and Figure S1, supporting information). As shown in Figure 6, the offset between theory and experiment depends on the temperature gradient between the hot and cold zones of the reactor. The cross‐over temperature and the discrepancy between the theoretically calculated equilibrium and experimentally determined steady‐state α values might depend on the geometry of the reactor, the pressure in of the reactor, the volume of the reactor, and the proportion of the cold and hot part of the reactor.

Oxygen three‐isotope phenomena due to thermal diffusion have been investigated previously using pure gases 63 , 64 and the experiments indicate that fractionation effects are well established in static reactors exposed to spatial temperature gradients. Due to mass balance, fractionation effects in these studies disappeared when the temperature gradient was removed after the experiment and the system was allowed to equilibrate. This is different to our situation, where cooling the reactor to room temperature implies that the O2‐CO2 exchange shuts down.

The fact that the experimental α values differ from theoretically equilibrium calculated data demonstrates that dynamical effects drive the system out of thermodynamic equilibrium. While kinetic isotope fractionation might occur in processes on the surface or at the solid–gas interface, the presence of a catalytic surface alone cannot change the thermodynamic equilibrium in the gas phase, by its very definition. 68 This implies that fractionation effects involving the catalytical substance are annihilated in forward and reverse reactions. 68 Similar to isotope fractionation in evaporation, surface chemical kinetics can only induce fractionation between O2 and CO2 if rapid removal processes in the gas phase prevent equilibrium being established. 69 The thermal gradient inducing a disequilibrium in the gas phase could possibly provide the required kinetic removal and thus potentially allow for surface kinetics to impact the O2‐CO2 equilibration. However, without further quantitative analysis, such effects cannot be distinguished from the isotope fractionation due to thermal diffusion, which also come into play once a thermal gradient is established. Contrary to previous claims, surface‐induced isotope effects during O2‐CO2 equilibration 25 cannot be studied without considering the impact of thermal diffusion.

4.3. θ (CO2/O2) of the CO2‐O2 isotope exchange

Experimentally determined θ (CO2/O2) values are larger than the canonical high temperature limit (θ >0.5305), an observation already made in previous studies. 12 , 19 , 70 Fosu et al 12 interpreted this as additional evidence that the CO2‐O2 isotope exchange must be kinetically controlled. The experimentally determined θ (CO2/O2) takes extreme values when 18 α (CO2/O2) is close to unity (near the crossover temperature where the sign of fractionation between CO2 and O2 changes, see Figure 7). These observations are a consequence of a singular behavior of θ close to the crossover, a well described anomaly in isotope equilibria. 32 , 71 , 72 , 73 , 74 , 75 Nevertheless, the experimentally determined θ values may not represent the CO2‐O2 equilibrium exchange due to the possible thermal diffusion effect (see above) and other heterogenous process involving adsorption, desorption, and mixing. 12 , 19 , 28 , 29 As a result, the experimentally determined θ values are steady‐state values.

It is interesting to investigate further the calculated θ (CO2/O2). Theoretically calculated values for 17 α (CO2/O2) and 18 α (CO2/O2) lead to values of θ = ln(17 α)/ln(18 α) which are outside the “canonical” range, i.e. θ > 0.5305 (see Figure 7). The upper bound of this conventional range arises from the high temperature limit (HTL) of the harmonic oscillator approximation (e.g. Young et al, 50 Cao and Liu, 31 Dauphas and Schauble, 53 Matsuhisa et al, 51 Kaiser et al 76 ).

θHTL=m161m171m161m181=0.5305 (5)

Examples in support of the approximation usually come from considering diatomic molecules, such as CO or O2, for example (Cao and Liu, 31 Wang et al 54 ). Here, we provide evidence that this empirical rule may not hold for more complex molecules like CO2. The calculated temperature dependence of the three‐isotope exponents for O + O2 (θOO2) and O + CO2 (θOCO2) isotope exchange reactions is shown in Figure S2 (supporting information). The three‐isotope exponent for atom‐molecule exchange (κ) is calculated as shown in Equation 6) following Cao and Liu. 31

θOXO=κXO=lnβXO17lnβXO18 (6)

where κXO or θOXO is the three‐isotope exponent for atom‐molecule exchange, which should also be restricted to the parameter range < θHTL. This is indeed the case for O2, but κCO2 clearly exceeds the expected range. According to our calculations, it reaches a HTL of around 0.533. We suppose that some structure in the results above 800°C is likely to be due to the smallness of the isotopic signatures, as well as to the precision and potential incompleteness of excited states in the line lists. While our κO2 value further agrees with the calculation of Cao and Liu, 31 to the fourth decimal place (or even better), our κCO2 value is higher by 0.0027 to 0.0035 over the full T range and it falls well above the canonical range, even at room temperature, where κCO2(25°C) = 0.5317 (see Figure S2, supporting information).

Cao and Liu 31 have shown that the three‐isotope exponent (θ ab ) for isotope exchange between two molecules a and b can also be calculated as a linear combination of the three‐isotope exponents κa and κb of the atom–molecule exchange equilibria for molecules a and b (see Equation 6), respectively:

θab=1cκa+cκb,wherec=lnβb18lnαab18 (7)

Interestingly, the coefficient c in Equation 7 does not depend on 17O, it only depends on 18O via the 18 β‐value of molecule b and the (statistically corrected) equilibrium constant 18 α a−b between molecules a and b.

We can use this relationship and the well‐studied H2O‐CO2 system for an independent estimate of κCO2. The three‐isotope relationship between liquid water and gaseous CO2 (θCO2H2Ol) was experimentally determined to be 0.5229 at 25°C (α18CO2H2Ol = 1.041036). 77 Using κH2O = 0.5300 calculated for water vapor by Cao and Liu 31 (a similar κH2O value was derived by the same authors from literature data on the gas vapor equilibrium) and βCO218 = 1.1194 (this study), κCO2 can be calculated using Equation (7), yielding 0.5324 at 25°C. This also is clearly larger than the canonical HTL. The result is only slightly different (+0.0001) when βCO218 = 1.1171 (from Richet et al 30 ) is used, in contrast to κCO2 = 0.5280 as obtained by Cao and Liu. 31 As a cautionary note, it needs to be pointed out that the κH2O value of Cao and Liu 31 was obtained without taking into account non‐Born‐Oppenheimer effects that seem to be important for water at room temperature. 57 When we use our value of κCO2(25°C) = 0.5317 to determine κH2Ol via Equation 7, from the measurements of α18CO2H2Ol = 1.041036 77 and θCO2H2Ol= 0.5229 77 and βCO218 = 1.1194, we obtain κH2Ol= 0.5294. Interestingly, Barkan and Luz 78 have also determined the equilibrium fractionation factors between the gaseous and liquid phases of H2O, from which the three‐isotope coefficient θCO2H2Og= 0.5241 for the equilibrium fractionation of water and CO2 in the gaseous phase can be derived. The corresponding α18CO2H2Og = 1.05082 implies that κH2Og= 0.5377 using our κCO2=0.5317, or κH2Og= 0.5311 when using the lower κCO2 from Cao and Liu. 31 Independently of the way of calculating κCO2, available data on the CO2‐H2O system seem to indicate that κH2Og also exceeds the canonical range limit. This implies that κ values of a chemical compound in different states might not be identical.

The independently derived κCO2 value of 0.5324 is somewhat higher than the value calculated directly from the experimentally improved ab‐initio data from Huang et al, 42 κCO2= lnβCO217/ln(βCO2)18 = 0.5317, and supports our finding that κCO2(slightly) exceeds the value of 0.5305 even at room temperature. At high temperatures (~1000°C), κCO2 should be even higher, by 0.01 to 0.02. This is in striking contrast to the κCO2 values estimated by Cao and Liu 31 which never exceed 0.5305, but their underlying calculations of partition function ratios seem to include only a limited number of anharmonic or higher order corrections to the simple harmonic oscillator approximation. Such corrections seem to be mostly negligible in the calculation of κ for diatomic molecules, such as O2, for which the κO2 values calculated in this study (κO2 at 25°C = 0.5282, for example) are in perfect agreement with the values estimated by Cao and Liu. 31 Nevertheless, they appear to be important for triatomic molecules (H2O, CO2, etc.) and probably for more complex molecules too.

Finally, Equation 7 can be used to determine θCO2O2 independent of our evaluation of 17 α (if it is suspected that the 17O data are less reliable than the other data, for example) or even independent of 18 α. Using κCO2(25°C) = 0.5324 (see previous paragraph), κO2(25°C) = 0.5282, 18 α = 1.0352 (1.0372) and 18 β O2 = 1.0813 (1.0773), results in θCO2O2= 0.5419 (0.5410). The values in parentheses are the corresponding values of Richet et al. 30 For comparison, the calculation based on the direct sum method for CO2 used in this study yields θCO2O2 = 0.5396 (0.5388) at 25°C, with the value in parentheses again being from the compilation of Richet et al. 30 Both routes of calculation yield values around 0.540, well beyond the “canonical range”.

According to Equation 7, the combination of both the unusually large κCO2> 0.5305 (exceeding κO2) and the relatively large and opposite‐sign coefficients of c ≃ −2 and (1 − c) ≃ 3 leads to the exceptionally high value of θCO2O2 = 0.5396. Nevertheless, the O + CO2 exchange alone with κCO2≃ 0.533 at temperatures above 300°C must be regarded as the origin of this exception since standard values of κCO2 close to κO2 would not lead to such departures, regardless of the coefficient c in Equation (7). The solid theoretical and experimental evidence for κCO2> 0.5305 and θ (CO2/O2) ~ 0.54 implies that analysis of three‐isotope effects in CO2 needs to go beyond the simple rigid‐rotor harmonic oscillator analysis. This might also apply to other non‐diatomic molecules and clumped isotope effects. 79 The above analysis of the CO2‐H2O system seems to support this hypothesis concerning the oxygen isotopic fractionation of the H2O molecule.

5. CONCLUSIONS

The steady‐state fractionation factors, α 17(CO2/O2) and α 18(CO2/O2), for the CO2‐O2 isotope exchange reaction depend on temperature, largely explaining differences between previously reported values. Nevertheless, all our experimentally determined values are lower than those obtained from the updated theoretical calculations of equilibrium fractionation, and the offset is strictly mass‐dependent. The discrepancy between the theoretical and experimental fractionation factors might be caused by thermal diffusion effects. While this requires further investigation, it provides an explanation for the inconsistent values reported in the literature.

We have shown that the CO2‐O2 isotopic exchange under equilibrium conditions is characterized by unusual oxygen three‐isotope behavior. Calculated θ (CO2‐O2) values based on the direct sum method for CO2, which includes anharmonic corrections, are larger than the canonical range limit for mass‐dependent fractionation processes (0.5305), which possibly needs to be extended to 0.54. The result can be derived independently using the experimentally determined θ (CO2/H2O) and α 18(CO2/H2O) values, together with the theoretically calculated fractionation factor (β) for oxygen exchange between CO2 and atomic O. Our analysis shows that the unusual behavior is linked to the energy levels of CO2 isotopologues. This could imply the need for more accurate representation of energy terms (beyond the simple Bigeleisen‐Mayer model) for poly‐atomic molecules.

6.

PEER REVIEW

The peer review history for this article is available at https://publons.com/publon/10.1002/rcm.9301.

Supporting information

Figure S1. Geometry and dimensions of the CO2‐O2 exchange reactor used by Adnew et al (2019) 1.

Figure S2. Three isotope exponents κ of O + O2 and O + CO2 isotope exchange equilibria as a function of temperature.

Table S1. β factors, κ and θ values calculated in this study for CO2 and O2. β factors (i.e. the 18O fractionation between the compound [in this case oxygen and carbon dioxide] and a dissociated and non‐interacting atom O 2−7) are also compared to the previously reported values by Richet et al,(1977) 2. Reldev stands for the relative deviation of β factors for O2 and CO2 calculated in this study and Richet et al,(1977)2 in per mill (‰) (β (Richet et al)/β (This study) – 1).

Table S2. Comparison of α18(CO2/O2) values calculated in this study to the previously reported values by Richet et al 2.

Table S3. The CO2−O2 exchange time to reach an isotopic steady state at 500°C.

ACKNOWLEDGEMENTS

The authors thank Eugeni Barkan and Rolf Vieten from the Hebrew University of Jerusalem for calibration of their O2 and CO2 working gases. GAA is supported by EU Horizon 2020 ERC‐ASICA project with a research grant number 64908 and project ALWPP.2016.013 of the Dutch Science Foundation NWO. The authors thank Wouter Peters from Wageningen University for collaboration and funding through ASICA. The authors appreciate the two anonymous reviewers for their constructive feedback which improved the manuscript.

Adnew GA, Workman E, Janssen C, Röckmann T. Temperature dependence of isotopic fractionation in the CO2‐O2 isotope exchange reaction. Rapid Commun Mass Spectrom. 2022;36(12):e9301. doi: 10.1002/rcm.9301

DATA AVAILABILITY STATEMENT

All the data used in this study are reported in the form of Figures and Tables.

REFERENCES

  • 1. Adnew GA, Pons TL, Koren G, Peters W, Röckmann T. Leaf‐scale quantification of the effect of photosynthetic gas exchange on Δ17O of atmospheric CO2 . Biogeosciences. 2020;17(14):3903‐3922. doi: 10.5194/bg-17-3903-2020 [DOI] [Google Scholar]
  • 2. Hofmann MEG, Horváth B, Schneider L, Peters W, Schützenmeister K, Pack A. Atmospheric measurements of Δ17O in CO2 in Göttingen, Germany reveal a seasonal cycle driven by biospheric uptake. Geochim Cosmochim AC. 2017;199:143‐163. doi: 10.1016/j.gca.2016.11.019 [DOI] [Google Scholar]
  • 3. Liang M‐C, Mahata S, Laskar AH, Thiemens MH, Newman S. Oxygen isotope anomaly in tropospheric CO2 and implications for CO2 residence time in the atmosphere and gross primary productivity. Sci Rep. 2017;7(1):13180 doi: 10.1038/s41598-017-12774-w [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4. Laskar AH, Mahata S, Liang M‐C. Identification of anthropogenic CO2 using triple oxygen and clumped isotopes. Environ Sci Technol. 2016;50(21):11806‐11814. doi: 10.1021/acs.est.6b02989 [DOI] [PubMed] [Google Scholar]
  • 5. Hoag KJ, Still CJ, Fung IY, Boering KA. Triple oxygen isotope composition of tropospheric carbon dioxide as a tracer of terrestrial gross carbon fluxes. Geophys Res Lett. 2005;32(2):L02802 doi: 10.1029/2004GL021011 [DOI] [Google Scholar]
  • 6. Koren G, Schneider L, van der Velde IR, et al. Global 3‐D simulations of the triple oxygen isotope signature Δ17O in atmospheric CO2 . J Geophys Res Atmos. 2019;124(15):8808‐8836. doi: 10.1029/2019JD030387 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7. Luz B, Barkan E, Bender ML, Thiemens MK, Boering KA. Triple‐isotope composition of atmospheric oxygen as a tracer of biosphere productivity. Nature. 1999;400(6744):547‐550. doi: 10.1038/22987 [DOI] [Google Scholar]
  • 8. Thiemens MK, Chakraborty S, Jackson TL. Decadal Δ17O record of tropospheric CO2: Verification of a stratospheric component in the troposphere. J Geophys Res Atmos. 2014;119(10):6221‐6229. doi: 10.1002/2013JD020317 [DOI] [Google Scholar]
  • 9. Landais A, Barkan E, Yakir D, Luz B. The triple isotopic composition of oxygen in leaf water. Geochim Cosmochim ac. 2006;70(16):4105‐4115. doi: 10.1016/j.gca.2006.06.1545 [DOI] [Google Scholar]
  • 10. Landais A, Barkan E, Luz B. Record of δ18O and 17O‐excess in ice from Vostok Antarctica during the last 150,000 years. Geophys Res Lett. 2008;35(2):L02709 doi: 10.1029/2007GL032096 [DOI] [Google Scholar]
  • 11. Sha L, Mahata S, Duan P, et al. A novel application of triple oxygen isotope ratios of speleothems. Geochim Cosmochim Acta. 2020;270:360‐378. doi: 10.1016/j.gca.2019.12.003 [DOI] [Google Scholar]
  • 12. Fosu BJ, Subba R, Peethambaran R, Bhattacharya SK, Ghosh P. Developments and applications in triple oxygen isotope analysis of carbonates. ACS Earth Space Chem. 2020;4(5):702‐710. doi: 10.1021/acsearthspacechem.9b00330 [DOI] [Google Scholar]
  • 13. Liljestrand F. The application of silica Δ17O and δ18O towards paleo‐environmental reconstructions. Doctoral dissertation, Harvard University, Graduate School of Arts & Sciences. 2019.
  • 14. Crockford PW, Hayles JA, Bao H, et al. Triple oxygen isotope evidence for limited mid‐Proterozoic primary productivity. Nature. 2018;559(7715):613‐616. doi: 10.1038/s41586-018-0349-y [DOI] [PubMed] [Google Scholar]
  • 15. Thiemens MH. History and applications of mass‐independent isotope effects. Annu Rev Earth Planet Sci. 2006;34(1):217‐262. doi: 10.1146/annurev.earth.34.031405.125026 [DOI] [Google Scholar]
  • 16. Adnew GA, Hofmann MEG, Paul D, et al. Determination of the triple oxygen and carbon isotopic composition of CO2 from atomic ion fragments formed in the ion source of the 253 Ultra high‐resolution isotope ratio mass spectrometer. Rapid Commun Mass Spectrom. 2019;33(17):1363‐1380. doi: 10.1002/rcm.8478 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17. Hofmann MEG, Pack A. Technique for high‐precision analysis of triple oxygen isotope ratios in carbon dioxide. Anal Chem. 2010;82(11):4357‐4361. doi: 10.1021/ac902731m [DOI] [PubMed] [Google Scholar]
  • 18. Mahata S, Bhattacharya SK, Wang CH, Liang M‐C. Oxygen isotope exchange between O2 and CO2 over hot platinum: An innovative technique for measuring Δ17O in CO2 . Anal Chem. 2013;85(14):6894‐6901. doi: 10.1021/ac4011777 [DOI] [PubMed] [Google Scholar]
  • 19. Barkan E, Musan I, Luz B. High‐precision measurements of δ17O and 17O‐excess of NBS19 and NBS18. Rapid Commun Mass Spectrom. 2015;29(23):2219‐2224. doi: 10.1002/rcm.7378 [DOI] [PubMed] [Google Scholar]
  • 20. Affek HP, Barka E. A new method for high‐precision measurements of 17O/16O ratios in H2O. Rapid Commun Mass Spectrom. 2018;32(23):2096‐2097. doi: 10.1002/rcm.8290 [DOI] [PubMed] [Google Scholar]
  • 21. Liang M‐C, Mahata S. Oxygen anomaly in near surface carbon dioxide reveals deep stratospheric intrusion. Sci Rep. 2015;5(1):11352 doi: 10.1038/srep11352 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22. Laskar AM, Mahata S, Bhattacharya SK, Liang M‐C. Triple oxygen and clumped isotope compositions of CO2 in the middle troposphere. Earth Space Sci. 2019;6(1):1205‐1219. doi: 10.1029/2019EA000573 [DOI] [Google Scholar]
  • 23. Laskar HA, Abhayanand SM, Vishvendra S, Bhola RG, Liang M‐C. A new perspective of probing the level of pollution in the megacity Delhi affected by crop residue burning using the triple oxygen isotope technique in atmospheric CO2 . Environ Pollut. 2020;263(Pt A):114542 doi: 10.1016/j.envpol.2020.114542 [DOI] [PubMed] [Google Scholar]
  • 24. Daëron M, Blamart D, Peral M, Affek HP. Absolute isotopic abundance ratios and the accuracy of Δ47 measurements. Chem Geol. 2016;442:83‐96. doi: 10.1016/j.chemgeo.2016.08.014 [DOI] [Google Scholar]
  • 25. Olack G, Colman AS. Modeling the measurement: Δ47, corrections, and absolute ratios for reference materials. Geochem Geophys Geosyst. 2019;20(7):3569‐3587. doi: 10.1029/2018GC008166 [DOI] [Google Scholar]
  • 26. Saenger CP, Schauer AJ, Heitmann OE, Huntington KW, Steig EJ. How 17O excess in clumped isotope reference‐frame materials and ETH standards affects reconstructed temperature. Chem Geol. 2021;563:120059 doi: 10.1016/j.chemgeo.2021.120059 [DOI] [Google Scholar]
  • 27. Adnew GA, Hofmann MEG, Pons TL, et al. Leaf scale quantification of the effect of photosynthetic gas exchange on Δ47 of CO2. Sci Rep. 2021;11:14023. doi: 10.1038/s41598-021-93092-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28. Mahata S, Bhattacharya SK, Liang MC. An improved method for high precision determination of ∆17O by oxygen isotope exchange between O2 and CO2 over hot platinum. Rapid Commun Mass Spectrom. 2016;30(1):119‐131. doi: 10.1002/rcm.7423 [DOI] [PubMed] [Google Scholar]
  • 29. Prasanna K, Bhattacharya SK, Ghosh P, Mahata S, Liang M‐C. Isotopic homogenization and scrambling associated with oxygen isotopic exchange on hot platinum: Studies on gas pairs (O2, CO2) and (CO, CO2). RSC Adv. 2016;6(56):51296‐51303. doi: 10.1039/C6RA08286F [DOI] [Google Scholar]
  • 30. Richet P, Bottinga Y, Javoy M. A review of hydrogen, carbon, nitrogen, oxygen, sulphur, and chlorine stable isotope fractionation among gaseous molecules. Annu Rev Earth Planet Sci. 1977;5(1):65‐110. doi: 10.1146/annurev.ea.05.050177.000433 [DOI] [Google Scholar]
  • 31. Cao X, Liu Y. Equilibrium mass‐dependent fractionation relationships for triple oxygen isotopes. Geochim Cosmochim Acta. 2011;75(23):7435‐7445. doi: 10.1016/j.gca.2011.09.048 [DOI] [Google Scholar]
  • 32. Hayles JA, Cao X, Bao H. The statistical mechanical basis of the triple isotope fractionation relationship. Geochem Perspect Lett. 2017;3:1‐11. [Google Scholar]
  • 33. Schauble EA. Applying stable isotope fractionation theory to new systems. Rev Mineral Geochem. 2004;55(1):65‐111. doi: 10.2138/gsrmg.55.1.65 [DOI] [Google Scholar]
  • 34. Urey HC. The thermodynamic properties of isotopic substances. J Chem Soc. 1947;562‐581:562 doi: 10.1039/jr9470000562 [DOI] [PubMed] [Google Scholar]
  • 35. Bigeleisen J, Mayer MG. Calculation of equilibrium constants for isotopic exchange reactions. J Chem Phys. 1947;15(5):261‐267. doi: 10.1063/1.1746492 [DOI] [Google Scholar]
  • 36. Dunham JL. The energy levels of a rotating vibrator. Phys Ther Rev. 1932;41(6):721‐731. doi: 10.1103/PhysRev.41.721 [DOI] [Google Scholar]
  • 37. Huber KP, Herzberg G. Molecular Spectra and Molecular Structure. Springer Science & Business Media; 2013. [Google Scholar]
  • 38. Wang M, Huang WJ, Kondev FG, Audi G, Naimi S. The AME 2020 atomic mass evaluation (II). Tables, graphs and references. Chinese Phys C. 2021;45(3):030003 doi: 10.1088/1674-1137/abddaf [DOI] [Google Scholar]
  • 39. Irikura KK. Experimental vibrational zero‐point energies: Diatomic molecules. J Phys Chem. 2007;36(2):389‐397. doi: 10.1063/1.2436891 [DOI] [Google Scholar]
  • 40. Janssen C, Tuzson B. Isotope evidence for ozone formation on surfaces. J Phys Chem. 2010;A114:9709‐9719. [DOI] [PubMed] [Google Scholar]
  • 41. Huang X, Gamache RR, Freedman RS, Schwenke DW, Lee TJ. Reliable infrared line lists for 13CO2 isotopologues up to E'=18,000 cm−1 and 1500 K, with line shape parameters. J Quant Spectrosc Radiat Transfer. 2014;147:134‐144. [Google Scholar]
  • 42. Huang X, Schwenke DW, Freedman RS, Lee TJ. Ames‐2016 line lists for 13 isotopologues of CO2: Updates, consistency, and remaining issues. J Quant Spectrosc Radiat Transfer. 2017;203:224‐241. doi: 10.1016/j.jqsrt.2017.04.026 [DOI] [Google Scholar]
  • 43. Prokhorov I, Kluge T, Janssen C. Optical clumped isotope thermometry of carbon dioxide. Sci Rep. 2019;9(1):4765 doi: 10.1038/s41598-019-40750-z [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44. Webb MA, Miller TF III. Position‐specific and clumped stable isotope studies: Comparison of the Urey and path‐integral approaches for carbon dioxide, nitrous oxide, methane, and propane. J Phys Chem A. 2014;118(2):467‐474. doi: 10.1021/jp411134v [DOI] [PubMed] [Google Scholar]
  • 45. Grebenshchikov SY. Photodissociation of carbon dioxide in singlet valence electronic states. I. Six multiply intersecting ab initio potential energy. J Chem Phys. 2013;138:224106 [DOI] [PubMed] [Google Scholar]
  • 46. Barkan E, Luz B. Conversion of O2 into CO2 for high‐precision oxygen isotope measurements. Anal Chem. 1996;68(19):3507‐3510. doi: 10.1021/ac9602938 [DOI] [PubMed] [Google Scholar]
  • 47. Kroopnick P, Craig H. Oxygen isotope fractionation in dissolved oxygen in the deep sea. Earth Planet Sci Lett. 1976;32(2):375‐388. doi: 10.1016/0012-821X(76)90078-9 [DOI] [Google Scholar]
  • 48. Taylor HP, Epstein S. Relationship between 18O/16O ratios in coexisting minerals of igneous and metamorphic rocks: Part 1: Principles and experimental results. Geol Soc Am Bull. 1962;73(4):461‐480. doi: 10.1130/0016-7606(1962)73[461:RBORIC]2.0.CO;2 [DOI] [Google Scholar]
  • 49. Taylor HP. 18O/16O ratios in coexisting minerals of igneous and metamorphic rocks. PhD thesis, California Institute of Technology, Pasadena, CA, USA. 1959:40–43.
  • 50. Young ED, Galy A, Nagahara H. Kinetic and equilibrium mass‐dependent isotope fractionation laws in nature and their geochemical and cosmochemical significance. Geochim Cosmochim AC. 2002;66(6):1095‐1104. doi: 10.1016/S0016-7037(01)00832-8 [DOI] [Google Scholar]
  • 51. Matsuhisa Y, Goldsmith JR, Clayton RN. Mechanisms of hydrothermal crystallization of quartz at 250°C and 15 kbar. Geochim Cosmochim Acta. 1978;42(2):173‐182. doi: 10.1016/0016-7037(78)90130-8 [DOI] [Google Scholar]
  • 52. Weston RE. Anomalous or mass‐independent isotope effects. Chem Rev. 1999;99(8):2115‐2136. doi: 10.1021/cr9800154 [DOI] [PubMed] [Google Scholar]
  • 53. Dauphas N, Schauble EA. Mass fractionation laws, mass‐independent effects, and isotopic anomalies. Annu Rev Earth Planet Sci. 2016;44(1):709‐784. doi: 10.1146/annurev-earth-060115-012157 [DOI] [Google Scholar]
  • 54. Wang Z, Schauble EA, Eiler JM. Equilibrium thermodynamics of multiply substituted isotopologues of molecular gases. Geochim Cosmochim Acta. 2004;68(23):4779‐4797. doi: 10.1016/j.gca.2004.05.039 [DOI] [Google Scholar]
  • 55. Gamache RR, Roller C, Lopes E, et al. Total internal partition sums for 166 isotopologues of 51 molecules important in planetary atmospheres: Application to HITRAN2016 and beyond. J Quant Spectrosc Radiat Transfer. 2017;203:70‐87. doi: 10.1016/j.jqsrt.2017.03.045 [DOI] [Google Scholar]
  • 56. Born M, Huang K. Dynamical Theory of Crystal Lattices. International series of monographs on physics. Oxford: Clarendon Press; 1956. [Google Scholar]
  • 57. Zhang Y, Liu Y. The theory of equilibrium isotope fractionations for gaseous molecules under super‐cold conditions. Geochim Cosmochim Acta. 2018;238:123‐149. doi: 10.1016/j.gca.2018.07.001 [DOI] [Google Scholar]
  • 58. Blatt H. Oxygen isotopes and the origin of quartz. J Sediment Res. 1987;57(2):373‐377. doi: 10.1306/212F8B34-2B24-11D7-8648000102C1865D [DOI] [Google Scholar]
  • 59. Abernathey JR, Rosenberger F. Soret diffusion and convective stability in a closed vertical cylinder. Phys Fluids. 1981;24(3):377 doi: 10.1063/1.863381 [DOI] [Google Scholar]
  • 60. Jones RC, Furry WH. The separation of isotopes by thermal diffusion. Rev Mod Phys. 1946;18(2):151‐223. doi: 10.1103/RevModPhys.18.151 [DOI] [Google Scholar]
  • 61. Ageev EP, Panchenkov GM. The separation of oxygen isotopes by thermal diffusion. Soviet Atomic Energy. 1964;14(5):518‐520. doi: 10.1007/BF01121902 [DOI] [Google Scholar]
  • 62. Müller G, Vasaru G. The Clusius‐Dickel thermal diffusion column − 50 years after its invention. Isotopenpraxis. 1988;24(11‐12):455‐464. doi: 10.1080/10256018808624027 [DOI] [Google Scholar]
  • 63. Sun T, Bao H. Thermal‐gradient‐induced non‐mass‐dependent isotope fractionation. Rapid Commun Mass Spectrom. 2011;25(6):765‐773. doi: 10.1002/rcm.4912 [DOI] [PubMed] [Google Scholar]
  • 64. Sun T, Bao H. Non‐mass‐dependent 17O anomalies generated by a superimposed thermal gradient on a rarefied O2 gas in a closed system. Rapid Commun Mass Spectrom. 2011;25(1):20‐24. doi: 10.1002/rcm.4825 [DOI] [PubMed] [Google Scholar]
  • 65. Hirschfelder OJ, Curtiss CF, Bird RB. Molecular theory of gases and liquids. New York: John Wiley and Sons, Inc.; 1964. [Google Scholar]
  • 66. Olson JM, Rosenberger F. Convective instabilities in a closed vertical cylinder heated from below. Part 2. Binary gas mixtures. J Fluid Mech. 1979;92(4):609‐629. doi: 10.1017/S0022112079000781 [DOI] [Google Scholar]
  • 67. Anav A, Friedlingstein P, Beer C, et al. Spatiotemporal patterns of terrestrial gross primary production: A review. Rev Geophys. 2015;53(3):785‐818. doi: 10.1002/2015RG000483 [DOI] [Google Scholar]
  • 68. Laidler KJ, Meiser JH, Sanctuary BC. Physical Chemistry. Houghton Mifflin; 2002. [Google Scholar]
  • 69. Mook W. Environmental isotopes in the hydrological cycle: principles and applications. Vol. 1. International Atomic Energy Agency and United Nations Educational, Scientific and Cultural Organization; 2000:39. [Google Scholar]
  • 70. Passey BH, Levin NE. Triple oxygen isotopes in meteoric waters, carbonates, and biological apatites: Implications for continental paleoclimate reconstruction. Rev Mineral Geochem. 2021;86(1):429‐462. doi: 10.2138/rmg.2021.86.13 [DOI] [Google Scholar]
  • 71. Kotaka M, Okamoto M, Bigeleisen J. Anomalous mass effects in isotopic exchange equilibria. J Am Chem Soc. 1992;114(16):6436‐6445. doi: 10.1021/ja00042a022 [DOI] [Google Scholar]
  • 72. Deines P. A note on intra‐elemental isotope effects and the interpretation of non‐mass‐dependent isotope variations. Chem Geol. 2003;199(1‐2):179‐182. doi: 10.1016/S0009-2541(03)00076-7 [DOI] [Google Scholar]
  • 73. Otake T, Lasaga AC, Ohmoto H. Ab initio calculations for equilibrium fractionations in multiple sulfur isotope systems. Chem Geol. 2008;249(3‐4):357‐376. doi: 10.1016/j.chemgeo.2008.01.020 [DOI] [Google Scholar]
  • 74. Spindel W, Stern MJ, Monse EU. Further studies on temperature dependences of isotope effects. J Chem Phys. 1970;52(4):2022‐2035. doi: 10.1063/1.1673255 [DOI] [Google Scholar]
  • 75. Vogel PC, Stern MJ. Temperature dependences of kinetic isotope effects. J Chem Phys. 1971;54(2):779‐796. doi: 10.1063/1.1674912 [DOI] [Google Scholar]
  • 76. Kaiser J, Röckmann T, Brenninkmeijer CAM. Contribution of mass‐dependent fractionation to the oxygen isotope anomaly of atmospheric nitrous oxide. J Geophys Res. 2004;109:D03305 [Google Scholar]
  • 77. Barkan E, Luz B. High‐precision measurements of 17O/16O and 18O/16O ratios in CO2 . Rapid Commun Mass Spectrom. 2012;26(23):2733‐2738. doi: 10.1002/rcm.6400 [DOI] [PubMed] [Google Scholar]
  • 78. Barkan E, Luz B. High precision measurements of 17O/16O and 18O/16O ratios in H2O. Rapid Commun Mass Spectrom. 2005;19(24):3737‐3742. doi: 10.1002/rcm.2250 [DOI] [PubMed] [Google Scholar]
  • 79. Liu Q, Tossell JA, Liu Y. On the proper use of the Bigeleisen–Mayer equation and corrections to it in the calculation of isotopic fractionation equilibrium constants. Geochim Cosmochim Acta. 2010;74(24):6965‐6983. doi: 10.1016/j.gca.2010.09.014 [DOI] [Google Scholar]

Associated Data

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

Supplementary Materials

Figure S1. Geometry and dimensions of the CO2‐O2 exchange reactor used by Adnew et al (2019) 1.

Figure S2. Three isotope exponents κ of O + O2 and O + CO2 isotope exchange equilibria as a function of temperature.

Table S1. β factors, κ and θ values calculated in this study for CO2 and O2. β factors (i.e. the 18O fractionation between the compound [in this case oxygen and carbon dioxide] and a dissociated and non‐interacting atom O 2−7) are also compared to the previously reported values by Richet et al,(1977) 2. Reldev stands for the relative deviation of β factors for O2 and CO2 calculated in this study and Richet et al,(1977)2 in per mill (‰) (β (Richet et al)/β (This study) – 1).

Table S2. Comparison of α18(CO2/O2) values calculated in this study to the previously reported values by Richet et al 2.

Table S3. The CO2−O2 exchange time to reach an isotopic steady state at 500°C.

Data Availability Statement

All the data used in this study are reported in the form of Figures and Tables.


Articles from Rapid Communications in Mass Spectrometry are provided here courtesy of Wiley

RESOURCES