钻石的压力和温度数据

Paolo Nimis
{"title":"钻石的压力和温度数据","authors":"Paolo Nimis","doi":"10.2138/rmg.2022.88.10","DOIUrl":null,"url":null,"abstract":"One of the key scientific questions about diamonds is “how are they formed?” To answer this question, we need to know the diamond-forming reactions and the physicochemical conditions under which these reactions take place. The pressure (P) and temperature (T) of diamond formation are an essential part of this knowledge and their assessment is pivotal to develop predictive scenarios of diamond distribution in the Earth interior. These scenarios may contribute to our understanding of global Earth processes, such as the long-term carbon cycle, and might also eventually improve our capability to select potential targets for diamond exploration (Shirey et al. 2013; Nimis et al. 2020).The evaluation of the P and T of diamond formation can be carried out at two levels of investigation. The first is concerned with formation conditions for individual diamonds or small populations of diamonds from specific sources. This approach has been so far the most widely practiced. The second level considers the statistical distribution of P–T conditions for diamond formation at either local or global scale. This type of investigation is hampered by the difficulty of obtaining large sets of suitable samples from a specific locality or for a statistically significant number of localities, and is therefore unavoidably affected to some degree by sampling bias. Despite inherent limitations, the latter approach is the most appropriate to reveal systematics in diamond P–T distributions and, ultimately, in diamond depth distribution within the Earth.Early reviews of P–T distributions for lithospheric diamonds were made by Nimis (2002), based on thermobarometry of chromian diopside inclusions, and by Stachel and Harris (2008) and Stachel (2014), using a more comprehensive set of thermobarometers. More recently, Nimis et al. (2020) investigated diamond depth distributions for a set of South African kimberlites and provided evidence for systematic trends of likely global significance. The depth distribution for sublithospheric diamonds worldwide was reviewed by Harte (2010). In this contribution, I first describe the methods that can be used to estimate the P–T conditions of diamond formation, highlighting their respective strengths and weaknesses. I then review existing diamond P–T data and their implications for diamond distribution with depth from both a local and a global perspective.Thermobarometry of diamonds can be carried out by estimating P–T conditions of chemical or elastic equilibrium of mineral inclusions contained within them. With some assumptions, the aggregation state of nitrogen substituting for carbon in the diamond lattice can also be used as a thermometer. In some cases, it is possible to derive both P and T estimates for a diamond by combining independent thermobarometric methods. In most instances, however, either P or T estimates can be directly retrieved with sufficient confidence. Below is a list of currently available methods for diamond thermobarometry. Those based on chemical equilibria are specific for different minerals or mineral pairs. Since many inclusions in diamonds occur as isolated grains, detached from their original mineral assemblage, particular emphasis is given in this review to methods that use single minerals to retrieve P and/or T. The following descriptions emphasize thermobarometry applications rather than principles, and the reader is referred to the original publications for details of the specific thermobarometers.Pyroxenes. Many ultramafic mantle rocks contain both clinopyroxene and orthopyroxene. This mineral pair forms one of the most widely used geological thermometers, which is based on the net-transfer reaction of the enstatite component between the two pyroxenes MgSiO3(Opx) ↔ MgSiO3(Cpx). The very small P-dependency of this reaction limits propagation of errors when T is calculated by iteration in combination with an independent barometer. Several versions of this thermometer have been proposed in the literature. Tests based on experimental results and pyroxene compositions in natural xenoliths suggest that the calibration of Taylor (1998) is the most reliable for mantle peridotites and pyroxenites in a T range of 700 to at least 1400 °C (Nimis and Grütter 2010), which covers the entire range of the lithospheric diamond window. The much more popular Brey and Köhler (1990) calibration gives similar results for clinopyroxenes with relatively low Na contents around 0.05 atoms per 6-oxygen formula unit (apfu), but was proven to overestimate with increasing NaCpx (e.g., +150 °C at NaCpx = 0.25 apfu; Nimis and Grütter 2010). For inclusions in diamonds and diamond-bearing peridotitic xenoliths (average NaCpx = 0.12 apfu, median = 0.09 apfu, standard deviation = 0.08 apfu), the Brey and Köhler (1990) calibration will yield estimates that are on average ~50 °C higher that those using the Taylor (1998) calibration. Larger discrepancies may be expected for unusually Na-rich clinopyroxenes.Since clinopyroxene–orthopyroxene inclusion pairs are very rare in diamonds, it is generally more practical to use the alternative single-clinopyroxene version of Nimis and Taylor (2000). The two-pyroxene Taylor (1998) and single-clinopyroxene Nimis and Taylor (2000) methods provide almost indistinguishable results when applied to mantle ultramafic rocks that contain both pyroxenes (Nimis and Grütter 2010). Although the Nimis and Taylor (2000) thermometer uses only one pyroxene to calculate T, its application expressly requires that both pyroxenes were part of the mineral assemblage and were in chemical equilibrium. Therefore, this method is only suitable for diamonds belonging to the lherzolitic or websteritic suites and containing clinopyroxene inclusions, either isolated or associated with other minerals. If orthopyroxene was not part of the original mineral assemblage, as is the case for wehrlitic inclusions, single-clinopyroxene thermometry would only provide a minimum T estimate. An anomalously low T estimate (e.g., one that lies much below the local geotherm) may be a sign that the clinopyroxene is wehrlitic. Ziberna et al. (2016) suggested that a Ca/(Ca + Mg) molar ratio of > 0.5 should also be considered as suspicious, as only ~1% of orthopyroxene-saturated mantle clinopyroxenes lie above this value. Nonetheless, discrimination of wehrlitic clinopyroxenes is not generally possible based merely on compositional criteria.Simakov (2008) proposed a different, more complex calibration of the single-clinopyroxene thermometer, which definitely improves performance above 1500 °C, though not over the T < 1400 °C range typical of ‘lithospheric’ temperatures (note that the TNimis and Taylor 2000 estimates reported in Fig. 10 of Simakov 2008 are incorrect and unduly suggest overestimation below 1300 °C).The Ca-in-Opx thermometer of Brey and Köhler (1990) is the single-orthopyroxene version of the pyroxene thermometer and can provide estimates that are complementary to, and independent from those obtained with the single-clinopyroxene thermometer. The two single-pyroxene methods were shown to be mutually consistent, provided a correction is applied below 900 °C (Nimis and Grütter 2010). However, since inclusions of likely harzburgitic (i.e., clinopyroxene-free) affinity are relatively common in diamonds (Stachel and Harris 2008), the possibility that TCa-in-Opx estimates for orthopyroxene inclusions are only minimum T values is much higher than for the single-clinopyroxene thermometer. Therefore, its usefulness is rather limited in diamond studies.Garnet. The Ni content of garnet in equilibrium with olivine is very sensitive to T and apparently independent of P (Ryan et al. 1996; Canil 1999). Since the Ni content of olivine shows small variations in both mantle xenoliths (O’Reilly et al. 1997) and inclusions in diamonds (Griffin et al. 1992; Sobolev et al. 2008), T can be retrieved from Nigarnet alone, by assuming an appropriate Ni content for coexisting forsteritic olivine. A useful reference values for Niolivine is the mean value for mantle olivines (mean ± standard deviation = 2900 ± 360 ppm; Ryan et al. 1996), which is also close to the average value for olivine inclusions in diamonds worldwide (Stachel and Harris 2008; Sobolev et al. 2009). If available, the mean Niolivine value for olivine inclusions in diamonds from the same locality may be used (e.g., 3150 ± 200 ppm for 51 inclusions from the Kalahari craton, Griffin et al. 1992; 2700 ppm for 88 inclusions from Arkhangel’sk province, Malkovets et al. 2011). Choosing one or the other value changes the final T estimate by at most a few tens of degrees (Fig. 1). The concentrations of Nigarnet can be measured with a laser-ablation inductively-coupled plasma mass spectrometer (LA-ICP-MS), an ion microprobe (SIMS) or a proton microprobe (PIXE), which typically ensure a precision within a few ppm. The assumption of equilibrium with olivine may require compositional filtering (e.g., Grütter et al. 2004) to exclude any non-peridotitic garnet.The Ni-in-garnet thermometer has seen wide application in studies of Cr-pyrope garnets included in diamonds (e.g., Griffin et al. 1992, 1993; Davies et al. 2004a; Viljoen et al. 2014; De Hoog et al. 2019). Its calibration, however, is somewhat controversial. Ryan et al. (1996) calibrated it against mantle xenoliths, using T values derived from a combination of the olivine–garnet Fe–Mg-exchange thermometer (O’Neill and Wood 1979) and the MacGregor (1974) and Brey and Köhler (1990) orthopyroxene–garnet barometers. Canil (1999) calibrated it against experiments at T ≥ 1200 °C. The two calibrations give identical results at ~1100 °C, but progressively diverge at lower and higher T (Fig. 1). As a result, estimates obtained through the Canil (1999) formulation will typically span over narrower T intervals. Considering the most typical range for Ni in garnet inclusions, the maximum difference TCanil–TRyan is ca. +100 °C at 20 ppm Ni and ca. –250 °C at 180 ppm Ni. It has been variously claimed that either calibration, or even their ‘average’, agree best with other independent thermometers when applied to mantle xenoliths (e.g., Ryan et al. 1996; Canil 1999; De Hoog et al. 2019; Czas et al. 2020; Nimis et al. 2020), but a definitive assessment using internally consistent thermobarometers as reference is still lacking. The latest attempt to refine the Ni-in-garnet thermometer was made by Sudholz et al. (2021a), who recalibrated it against new experiments in a relatively narrow T range (1100–1325 °C) and introduced correction terms for Ca and Cr contents in the garnet. When tested against independent estimates for xenoliths using the Nimis and Taylor (2000) enstatite-in-clinopyroxene thermometer, the Sudholz et al. (2021a) calibration shows improved overall accuracy above 1100 °C relative to the Canil (1999) experimental calibration, but poorer overall precision and slightly stronger progressive overestimation at lower T (see Fig. 7 in Sudholz et al. 2021a).The Mn content of garnet in equilibrium with mantle olivine is sensitive to T (Smith et al. 1991) and can be used as a single-mineral thermometer, assuming a constant Mn content in olivine. This thermometer relies on electron microprobe data and was proposed as a substitute for the Ni-in-garnet thermometer when trace element data are not available (Grütter et al. 1999; Creighton 2009). The declared precision is rather poor (mostly ±150 °C) and, in the absence of independent olivine data, severe underestimation may occur at high T. Therefore, this method may have some use in surveys of large populations of garnet xenocrysts recovered during diamond exploration (e.g., Grütter and Tuer 2009), but its utility for diamond thermobarometry is limited.Ashchepkov et al. (2010) calibrated two monomineral thermometers for garnets in equilibrium with clinopyroxene or olivine, respectively. These methods are simplified versions of Fe–Mg-exchange thermometers for clinopyroxene–garnet or olivine-garnet pairs, in which the clinopyroxene and olivine compositions are modeled from the composition of the garnet. Their accuracy cannot be better than that of the original two-mineral formulations, which are problematic themselves (see below). Also, P must be independently known to calculate T. In a test on mantle xenoliths from the Udachnaya kimberlite, these thermometers reproduced temperatures calculated with the pyroxene thermometer of Brey and Köhler (1990), at pressures calculated with the orthopyroxene–garnet barometer, to within ca. ±100 °C. Additional uncertainties due to problematic estimation of P in the absence of compositional data for coexisting orthopyroxene (see below) probably make these thermometers of little interest for diamond thermobarometry, especially when investigating small populations of diamonds.Olivine. The Al content in peridotitic olivine is sensitive to T and can be used as a thermometer. The first Al-in-olivine thermometer for peridotitic olivine at conditions relevant to diamond was empirically calibrated on mantle xenoliths by De Hoog et al. (2010), using P–T estimates derived by a combination of the two-pyroxene thermometer and orthopyroxene–garnet barometer of Brey and Köhler (1990). The method was recalibrated by Bussweiler et al. (2017) by analysis of olivine from high-P–T experimental charges. The Al content in cratonic mantle olivine ranges from ~1 ppm to ~250 ppm and is conveniently analyzed by techniques such as LA-ICP-MS or SIMS, which ensure a precision within a few ppm. Electron microprobe analysis (EMPA) using very high probe currents (e.g., ≥ 200 nA) and long count times (≥ 100 s) may be a valid alternative (Batanova et al. 2018; D’Souza et al. 2020).Al-in-olivine temperatures for spinel peridotites or strongly metasomatized garnet peridotites may be significantly underestimated or, respectively, overestimated. Compositional screening using an Al vs V diagram may help to discriminate these ‘unsafe’ olivines (Bussweiler et al. 2017). Screening is especially important if olivine is not physically associated with garnet, as occurs with many inclusions in diamonds. The Al-in-olivine thermometer gives systematically higher estimates (~50 °C, slightly increasing below 900 °C) than the single-clinopyroxene thermometer (Nimis and Taylor 2000) for natural mantle peridotites, probably reflecting slight suppression of Al incorporation in the experimental olivines due to Na loss (Bussweiler et al. 2017). The discrepancy is sufficiently small to ensure successful application of the method, but should be considered when comparing estimates obtained using different thermometers. The Al-in-olivine thermometer has not yet been widely used in diamond studies (see Korolev et al. 2018 for an example of application), owing to its recent development and the necessity of non-routine analysis of Al. Considering the great abundance of olivine among inclusions in diamonds (Stachel and Harris 2008), its application is likely to increase significantly in the future.Spinel. The Zn content of spinel in equilibrium with mantle olivine is sensitive to T. Since the Zn content of mantle olivine is nearly constant (mean ± standard deviation = 52 ± 14 ppm), the Zn content of spinel can directly be used as a thermometer (Ryan et al. 1996). The concentrations of Zn in spinel can be measured by SIMS or PIXE, which ensure a precision within a few ppm. The current version of the Zn-in-spinel thermometer was calibrated over the range 680–1180 °C against T for coexisting garnets using the Ryan et al. (1996) version of the Ni-in-garnet thermometer. Therefore, it is subject to at least the same uncertainties as the Ni-in-garnet thermometer (see above).Fe–Mg exchange thermometers. These popular thermometers are based on the distribution of Fe2+ and Mg between two mineral phases and can be used to derive T for inclusions in which garnet is associated with olivine, orthopyroxene or clinopyroxene. Tests using mineral compositions from mantle xenoliths and experiments showed that these thermometers suffer from either low precision or systematic errors or both (see review of Nimis and Grütter 2010). In peridotitic systems, the olivine-garnet and the clinopyroxene–garnet thermometers show the lowest precision, with possible discrepancies exceeding 200 °C. The Harley (1984) orthopyroxene–garnet formulation shows the highest precision, but systematically overestimates at T < 1100 °C and underestimates at T > 1100 °C (by up to ~150 °C, on average, for mantle xenoliths). All these discrepancies may reflect a modest sensitivity to T of the exchange reaction (particularly for olivine-garnet), oversimplification in the solid solution models used (especially for garnet and clinopyroxene), and Fe3+/ΣFe ratios in the minerals different from those in the calibration samples. Matjuschkin et al. (2014) experimentally demonstrated the large effect of neglected Fe3+ on orthopyroxene–garnet and olivine-garnet thermometers. Their experiments in the Na-free CaO–FeO–Fe2O3–MgO–Al2O3–SiO2 system at 5 GPa and 1000–1400 °C suggest that temperature estimates may be significantly improved if calculations are made using Fe2+ instead of total Fe in the garnet. Mössbauer data for orthopyroxene and garnet in peridotitic xenoliths, however, suggest that the partitioning of Fe3+ between these two minerals is affected by pressure and Na content in the orthopyroxene (Nimis et al. 2015). Therefore, a test over a range of relevant pressures in a Na-bearing system would be desirable. Nimis and Grütter (2010) empirically corrected the Harley (1984) thermometer using two-pyroxene thermometry of well equilibrated mantle xenoliths as calibration. The corrected version generally yields lower T estimates at low T and higher T estimates at high T relative to the uncorrected thermometer and may be more accurate for orthopyroxene–garnet pairs equilibrated under ‘average’ mantle redox conditions. On the other hand, it may produce large errors (> to ≫100 °C) if redox conditions are far from ‘average’, especially for strongly oxidized conditions at very high pressure and when both P and T are calculated by iteration in combination with an orthopyroxene–garnet barometer (Nimis et al. 2015). Considering that at deep lithospheric levels diamond is stable over a relatively wide range of redox conditions (∆logfO2 FMQ ≈ –5 to –2), this problem can be serious. Also, application to some xenolith suites tends to produce frequent high-P–T outliers (Grütter, pers. comm.). In spite of its recognized systematic discrepancies, the original Harley (1984) version may still be preferable for diamond thermobarometry due to its lower P dependence, which reduces error propagation during iterative P–T calculations. However, the systematic deviations of the Harley (1984) thermometer at T far from 1100 °C, especially if all Fe is treated as Fe2+, should be considered in data evaluation.The clinopyroxene–garnet thermometer represents the only viable method to determine T from major element compositions of eclogitic inclusions in diamonds. There are many versions of this thermometer in the literature. The most recent calibration (Nakamura 2009) is based on an expanded experimental database covering mafic and ultramafic compositions across the P and T ranges relevant to diamond (1.5–7.5 GPa, 800–1820 °C) and incorporates updated solid solution models for both clinopyroxene and garnet. Older, more simplified, but in some cases still very popular versions of this thermometer do not reproduce the same experiments equally well and often show systematic deviations with changing T and composition. The nominal uncertainty of the Nakamura (2009) thermometer (±74 °C at the 1-sigma level) is still about double that of the best performing thermometers for peridotitic inclusions. None of the available versions of the clinopyroxene–garnet thermometer include corrections for the content of jadeite in clinopyroxene, therefore application to clinopyroxenes with Na > 0.5 apfu is not recommended. This excludes 15% of reported compositions for inclusions in eclogitic diamonds. As regards the Fe3+ problem, Purwin et al. (2013) and Matjuschkin et al. (2014) measured Fe3+/ΣFe ratios of coexisting clinopyroxenes and garnets in a CaO–FeO–Fe2O3–MgO–Al2O3–SiO2 system at 2.5–5.0 GPa and 800–1400 °C and found no systematic deviations in the results of clinopyroxene–garnet Fe–Mg-exchange thermometry using the Krogh (1988) thermometer when all Fe was treated as Fe2+. However, since the system investigated was Na-free, it did not allow for the incorporation of Fe3+ in clinopyroxene as aegirine component. This mechanism was shown to be relevant in clinopyroxenes from garnet peridotites (Woodland 2009) and may be even more so in eclogitic clinopyroxenes. The effect of Na on Fe3+ uptake in clinopyroxene probably contributes to the lower precision of the clinopyroxene–garnet thermometer in applications to natural samples.Trace element-based clinopyroxene–garnet thermometers. The partitioning of REE between clinopyroxene and garnet is sensitive to T and has been used as a base for a clinopyroxene–garnet thermometer (Witt-Eickschen and O’Neill 2005; Sun and Liang 2015; Pickles et al. 2016). A possible advantage of REE thermometry over classical Fe–Mg-exchange thermometry is that REE diffusivity is small compared to that of Fe and Mg. Therefore, REE thermometry is predictably more robust against reequilibration of touching inclusions in diamonds during kimberlite transport or subsequent cooling. The version by Pickles et al. (2016) is particularly interesting, as it is independent of other thermometers and reproduces experimental temperatures moderately well (mean deviation of ±76 °C) with a small P dependence (~50 °C/1 GPa). This method is also independent of variations in Fe3+/ΣFe ratio and may thus pose a promising alternative to the traditional clinopyroxene–garnet Fe–Mg exchange thermometer. A practical disadvantage is the requirement of high-quality REE data, which cannot be obtained with routine electron microprobe analysis. To the author’s knowledge, there are no T data for diamonds using this method.Nitrogen-aggregation thermochronometry. Nitrogen is the most common impurity in natural diamonds. With time, nitrogen atoms in the diamond crystal lattice tend to aggregate, proceeding from initial C-centers (single nitrogen) in Type Ib diamond, to A-centers (nitrogen pairs) in Type IaA diamond and, finally, to B-centers (nitrogen in tetrahedral arrangement) in Type IaB diamond. The first transformation is completed in relatively short geological times (for a N content of 1000 ppm: ~100 yr at 1350 °C, and ~16 m.y at 950 °C; Taylor et al. 1996). Therefore, most natural diamonds are of the intermediate IaAB variety, the proportion of B-centers being a function of nitrogen content, temperature and time (Taylor et al. 1990; Mendelssohn and Milledge 1995). If the ages of the diamond and of its host kimberlite are known, one can calculate the integrated average temperature under which the diamond has resided in the mantle since its formation, i.e., the mantle residence temperature, Tres. Radiometric dating of inclusions within a particular diamond is rarely available. More commonly, the age of the diamond is assumed on the basis of radiometric dating of other diamonds from the same source. However, nitrogen aggregation is much more sensitive to temperature than to time and can be used as an effective thermometer, assuming the diamond has resided in the mantle at constant T since formation (Taylor et al. 1990, 1996). A diagram of nitrogen content vs the relative percentage of B-centers permits appreciation of age uncertainties on thermometry, since isotherms can be drawn for different mantle residence times (Fig. 2). Chronometry errors of one b.y. propagate thermometry errors of only a few tens of degrees on calculated Tres, unless the diamond is unusually young.Nitrogen content and aggregation can be measured by Fourier-Transform Infrared (FTIR) spectroscopy, using an infrared microscope in transmission mode. This method logically provides nitrogen content and aggregation data averaged over the volume of diamond analyzed. Working on thinned (ideally, a few hundred µm) samples at high spatial resolution may help to detect zoning and to obtain more robust estimates from distinct diamond growth zones. The reader is referred to Kohn et al. (2016) for practical guidelines on spatially resolved FTIR thermochronometry and for interpretation of data for zoned diamonds. Detection limits for nitrogen content in diamonds are typically on the order of 5–10 ppm for high-quality spectra. Nitrogen contents below detection limit (Type-II diamonds) are relatively rare in lithospheric diamonds, but are more common in sublithospheric diamonds and are clearly unsuitable for nitrogen-aggregation thermochronometry. It must be clear that, contrary to thermodynamic equilibria used in inclusion thermobarometry, nitrogen aggregation is a kinetic process and so does not provide a snapshot of temperature conditions at a certain time. For instance, a short period of elevated temperature, e.g., due to circulation of hot fluids in the mantle during or after diamond growth, may have a large effect on nitrogen aggregation and could lead to ambiguous interpretations of calculated Tres. In general, unless the diamond is very young relative to its kimberlite host, heating on a time scale of ~1 m.y. will have small effects on the calculated Tres, which will approach ambient mantle conditions rather than conditions during diamond formation (Fig. 3a). If the diamond experienced secular cooling over a b.y. time scale subsequent to formation, the calculated Tres may both significantly underestimate the formation T and overestimate the final ambient T (Fig. 3b).Orthopyroxene–garnet. The Al content of orthopyroxene in equilibrium with garnet is strongly sensitive to P and forms the basis of the most widely used barometer for garnet-bearing peridotites. In this barometer, P mostly depends on the Al content in the orthopyroxene, but the compositions of both minerals must be considered to obtain robust estimates. Of the many calibrations of this barometer, that of Brey and Köhler (1990) is by far the most popular, but the older Nickel and Green (1985) calibration reproduces experimental pressures for most peridotitic systems to somewhat better precision (Nimis and Grütter 2010). Still, the Nickel and Green (1985) version was formulated under simple model-system assumptions and requires correction of Al-component activity to avoid P underestimation for orthopyroxenes with non-negligible contents of jadeite component. Carswell and Gibb (1987) and Carswell (1991) proposed two simple correction schemes, based on different assumptions on how Ti is coupled with Na and Al in orthopyroxene. The two corrections only apply to orthopyroxenes with Na > (Cr + Fe3+ + 2·Ti) or, respectively, Na > (Cr + Fe3+ + Ti) atoms per formula unit, Fe3+ usually being neglected in practical applications. That of Carswell and Gibb (1987), which was also adopted by Brey and Köhler (1990) for their barometer, produces slightly smaller P variations, but there is as yet no solid experimental basis to prefer one correction over the other. The T-dependence of the orthopyroxene–garnet barometer is reasonably small (~0.2–0.3 GPa / 50 °C), but still sufficiently large to require an accurate independent estimate of T. This fact should be considered when this barometer is used in combination with the Harley (1984) orthopyroxene–garnet thermometer, which is known to overestimate at low T and underestimate at high T (see above): the artificial compression of T estimates produced by the Harley (1984) thermometer will necessarily result in non-natural compression also of P estimates. In general, errors in input T will displace the calculated P–T estimates roughly along a typical cratonic geotherm, thus limiting the ‘apparent’ scatter in P–T plots for samples equilibrated on the geotherm (Fig. 4). This may be an advantage if the aim is to define the mantle geotherm, but may give the misperception of a high overall precision by masking the true dispersion of the data.Clinopyroxene–garnet. Barometry of bimineralic eclogites and of eclogitic clinopyroxene–garnet inclusion pairs in diamonds has long been a problem for mantle scientists. There have been several attempts to develop a suitable barometer based on the P-sensitive equilibrium grossular + pyrope ↔ diopside + Ca-Tschermak. According to a recent test by Beyer et al. (2018), the Beyer et al. (2015) formulation reproduces pressures for experiments in natural systems significantly better than the earlier formulations of Simakov and Taylor (2000) and Simakov (2008). Nonetheless, in the Beyer et al. (2018) test on natural xenoliths from the Jericho kimberlite, Slave craton, P estimates for eclogites were systematically lower (by ~1 GPa) than orthopyroxene–garnet P estimates for peridotites recording similar T. As a result, the Jericho eclogites oddly appeared to fall on a hotter geotherm. The T-dependence of the clinopyroxene–garnet barometer is ca. 0.25 GPa / 50 °C, i.e., similar to that of the orthopyroxene–garnet barometer. Therefore, also in this case, errors in input T would displace P–T estimates roughly along the geotherm and cannot explain the observed discrepancy between peridotite-system and eclogite-system thermobarometry results.A further test of clinopyroxene–garnet thermobarometry is illustrated in Figure 5, for touching or non-touching inclusions in diamonds and diamond- or graphite-bearing xenoliths. The P–T estimates reflect iterations of the Beyer et al. (2015) barometer and the Nakamura (2009) Fe–Mg exchange thermometer. Since the uncertainty of the barometer increases rapidly with decreasing concentrations of [4]Al in clinopyroxene, application to clinopyroxenes with Si contents >1.985 atoms per 6-oxygen formula is not recommended by Beyer et al. (2015). These compositions are relatively common for inclusions in diamonds (35% of 224 inclusions) and diamondiferous xenoliths (42% of 233 xenoliths), thus numerous samples had to be excluded. Even if further severe filters were applied to discard potentially low-quality chemical analyses, the results are disconcerting: although graphite-bearing xenoliths plot in the graphite stability field, most diamond-bearing xenoliths and inclusions in diamonds yield too low pressures well outside the diamond stability field (Fig. 5). Iteration of earlier formulations of the same barometer and thermometer does not improve the results (cf. Fig. 10 in Shirey et al. 2013). The reason for this discrepancy is unclear. Simakov and Taylor (2000) cautioned that a barometer based on the Ca-Tschermak content in clinopyroxene may lead to incorrect estimates when applied to kyanite- or SiO2-oversaturated assemblages, but the Beyer et al. (2015) barometer shows no systematic deviations in calculated pressures for experiments in these systems. The large chemical complexity of eclogitic garnets and, especially, eclogitic clinopyroxenes might in part explain the poor performance of the barometer in applications to natural systems, in which variable amounts of Na, Fe3+, Cr, K and vacancy-bearing (i.e., Ca-Eskola) components occur. A further possible explanation is the high sensitivity of the barometer to even small errors in the chemical analyses of clinopyroxene. Beyer et al. (2015) provided a formula to predict relative errors on calculated pressures due to propagation of uncertainties on SiO2 contents, assuming an uncertainty of 1% relative in electron microprobe analysis: %err = 1.94 × 10–8e10.18398[Si apfu]. These model errors do not consider the additional errors that derive from iterative calculation of P and T (Fig. 5). A 1% relative increase in SiO2 contents may move many inaccurate P–T estimates into the diamond field, suggesting that application of the clinopyroxene–garnet barometer may require very high-quality, well-standardized electron microprobe analyses to yield meaningful results. Many of the analyses reported in the literature are possibly of routine quality and do not meet the necessary standards for robust clinopyroxene–garnet barometry. Still, analytical quality alone is unlikely to account for all of the observed discrepancies (Fig. 5).Sun and Liang (2015) proposed that REE partitioning between clinopyroxene and garnet could be used not only as a thermometer (see above) but also as a barometer. Unfortunately, the discrepancies between calculated PREE and run pressures for a set of independent validation experiments appear to be very large (up to ~3 GPa). A test conducted on twelve quartz eclogites, two graphite eclogites and nine diamond eclogites passed the appropriate stability-field constraints for all but two of the nine diamond eclogites; these fell at 0.2 and 0.5 GPa lower P than the diamond-graphite phase boundary. The proposed REE thermobarometry approach is promising, but further tests are necessary before this method can be used with confidence in diamond studies.If trace element data are available, the partitioning of Li between clinopyroxene and garnet can also be used as a barometer (Hanrahan et al. 2009). Pressures for fifteen calibration experiments on a Mg-rich eclogitic composition at 4–13 GPa and 1100–1500 °C are reproduced to a very reasonable ±0.2 GPa (1 sigma) and the T dependence of the barometer is small (~0.2 GPa / 50 °C). Preliminary tests yielded conditions compatible with diamond for three diamond-bearing xenoliths and for non-touching inclusion pairs in four diamonds, whereas two other diamonds yielded conditions up to 2 GPa lower than the diamond stability field. Hanrahan et al. (2009) suggested that the two spurious P estimates could reflect disequilibrium between the non-touching inclusions. However, the use of non-touching inclusions is essential for this barometer, as the high diffusivity of Li may lead to reequilibration of small touching inclusions during and after transport to surface. Similar to the REE barometer, further tests are needed before this method can be recommended for diamond studies.Monomineral barometers.Nimis and Taylor (2000) developed an empirical Cr-inclinopyroxene barometer for chromian diopsides in equilibrium with garnet. In combination with the Nimis and Taylor (2000) single-clinopyroxene thermometer, this barometer uniquely permits derivation of both P and T conditions based on the composition of a single mineral inclusion. The T-dependence of the Cr-in-clinopyroxene barometer (0.15–0.25 GPa / 50 °C for diamond inclusion compositions) is generally lower than that of the orthopyroxene–garnet barometer (~0.2–0.3 GPa / 50 °C) and the P-dependence of the complementary single-clinopyroxene thermometer is very small (Fig. 4). Therefore, if P and T are calculated by iteration, errors in input T or P will displace a clinopyroxene P–T estimate at an angle relative to a cratonic geotherm and will contribute to ‘apparent’ scatter in P–T results for samples equilibrated on a steady-state geotherm. The Nimis and Taylor (2000) Cr-in-clinopyroxene barometer tends to progressively underestimate above 4.5 GPa (by up to ~1 GPa at 7 GPa; Nimis 2002; Ziberna et al. 2016). This is a known artifact and the discrepancy should be considered when comparing Cr-in-clinopyroxene estimates with those obtained using other methods. Nimis et al. (2020) proposed an empirical correction to minimize the differences with P estimates obtained using the orthopyroxene–garnet barometer at high P, at the cost of somewhat lower precision. These authors cautioned that their correction represented a temporary measure to reduce inconsistencies between independent barometric estimates. More recently, Sudholz et al. (2021b) provided an experimental recalibration of the Cr-in-clinopyroxene barometer that yields results very similar to the empirically corrected calibration for clinopyroxenes associated with diamonds or hosted in well equilibrated peridotitic xenoliths at P > 5 GPa (Fig. 6a). At lower P the two revised calibrations diverge slightly, but the Nimis et al. (2020) calibration maintains somewhat better overall consistency with orthopyroxene–garnet barometry (Fig. 6b,c). For proper application of the Cr-in-clinopyroxene barometer, the analyses must undergo compositional filtering in order to select clinopyroxenes in equilibrium with garnet and eliminate certain compositions for which analytical uncertainties propagate excessive uncertainties on the calculated P. Ziberna et al. (2016) provided a useful cookbook to perform the necessary compositional screening, which slightly improved an earlier one proposed by Grütter (2009). According to this cookbook, some compositions require that electron microprobe analyses are carried out using beam currents and counting times higher than routine to minimize analytical uncertainties, whereas some unfavorable compositions should simply be discarded. In particular, the recommended limitation to Cr/(Cr+Al)mol > 0.1 excludes about two thirds of reported inclusions classified as websteritic. When applied to properly filtered graphite- or diamond-bearing xenoliths and inclusions in peridotitic diamonds, the Cr-in-clinopyroxene barometers yield results consistent with or within ~0.3 GPa of the stability field of the respective carbon phase (Nimis and Taylor 2000; see also Fig. 8a). The lack of adequate filtering in previously published thermobarometric results greatly contributed to the larger overall scatter in clinopyroxene P–T plots relative to those for orthopyroxene–garnet pairs (cf. Stachel and Harris 2008).Griffin and Ryan (1995) devised an empirical barometer based on the Cr content of garnet in equilibrium with orthopyroxene and spinel. This Cr-in-garnet barometer was later refined by Ryan et al. (1996) and further revised, using a different Cr/Ca-in-garnet approach, by Grütter et al. (2006). The last two versions are the most widely used, typically in combination with a Ni-in-garnet thermometer. That of Grütter et al. (2006) gives up to ~10% lower estimates and a better agreement with petrological constraints imposed by the graphite–diamond transition, and is therefore preferable. If the coexistence of garnet with spinel is unknown, as is the case for many garnet inclusions in diamonds, the Cr-in-garnet methods only provide an estimate of minimum pressure.Simakov (2008) proposed a simplified version of the clinopyroxene–garnet barometer for eclogites that solely relies on the Ca-Tschermak content in the clinopyroxene. The accuracy of this monomineral barometer cannot be better than that of the native two-phase barometer and is thus prone to the same problematic issues (see above). Ashchepkov et al. (2017) calibrated or recalibrated a series of empirical monomineral thermobarometers, which should be suitable for peridotitic and eclogitic garnets and clinopyroxenes. When applied to mantle xenoliths and diamond inclusions from individual localities, these methods produce strongly scattered P–T estimates (e.g., Figs. 9–12 in Ashchepkov et al. 2017), which are difficult to compare with the more regular P–T distributions that are typically obtained with other, more widely tested methods. The usefulness of the Ashchepkov et al. (2017) monomineral formulations for diamond thermobarometry is therefore questionable.With increasing pressure beyond ~7 GPa, mantle garnets incorporate increasing proportions of majoritic (i.e., pyroxene-like) component by substitution of Si for octahedrally coordinated Al and Cr. The Na and Ti contents also tend to increase. Collerson et al. (2000) utilized the relationship observed between majorite fractions and pressure in experiments on mafic and ultramafic systems to devise an empirical single-mineral barometer. The method was later revised by Collerson et al. (2010), Wijbrans et al. (2016) and Beyer and Frost (2017). Only Wijbrans et al. (2016) proposed two distinct calibrations, one for peridotitic and another for eclogitic systems. Experimental tests by Beyer and Frost (2017) and Beyer et al. (2018) showed systematic deviations or decreased precision for all but the Beyer and Frost (2017) formulation. This was particularly relevant for eclogitic compositions, which were underrepresented in previous calibration datasets. The Beyer and Frost (2017) barometer was calibrated at pressures of 6 to 20 GPa and temperatures of 900 to 2100 °C. These conditions cover sublithospheric mantle conditions up to ~575 km depth, corresponding to the lower transition zone. Experimental pressures are reproduced with a standard error of estimate of 0.86 GPa. In relative terms, considering the enormous range of pressures covered by the calibration, this uncertainty is not much larger than that of the barometers used for lithospheric diamonds. The effect of temperature on pressure estimates is apparently negligible. Note that application of the majorite barometer to lithospheric garnets equilibrated at P < 6 GPa is not recommended, because under such conditions the barometer becomes extremely sensitive to analytical errors (Beyer and Frost 2017). Tao et al. (2018) showed that high Fe3+ contents in the garnet tend to produce underestimated pressures with previous majorite barometers and provided a further revised calibration of the Collerson et al. (2010) barometer that explicitly takes into account Fe3+. The Tao et al. (2018) calibration may be preferable when Fe3+ data are available and measured Fe3+/ΣFe ratios are greater than 0.2. All these majorite barometers require that garnet occurs in equilibrium with a clinopyroxene phase. This may be an issue, because with increasing pressure pyroxene progressively dissolves into garnet and at pressures above ~15–19 GPa, depending on bulk composition and temperature, mantle rocks no longer contain pyroxene (Irifune et al. 1986; Irifune 1987; Okamoto and Maruyama 2004). Pyroxene-undersaturated condition can be suggested by the presence of CaSiO3 phases along with majoritic garnet as inclusions in diamonds or by high Ca contents in majoritic garnet. Under such conditions, the barometer would give erroneously low pressures (Harte and Cayzer 2007). If there is no independent evidence that the garnet was pyroxene-saturated, any pressure estimates close to that of the clinopyroxene-out reaction should be considered as estimates of the minimum P of entrapment. Thomson et al. (2021) devised a novel type of majorite barometer using machine-learning algorithms. Compared to traditional majorite barometers, this new barometer is apparently insensitive to petrological limitations (most notably, the absence of clinopyroxene) and reproduces experimental majoritic garnet compositions over a wider range of bulk compositions and experimental pressures (6–25 GPa) with much better overall precision (mostly within ±2 GPa). Thomson et al. (2021) caution that the compositions of majoritic garnet inclusions in diamonds lie in a region with relatively fewer experiments and machine-learning regressions may not be reliable in extrapolation. Thus, pressure predictions for inclusions in diamonds may have somewhat larger uncertainties. Still, this is the only available barometric method that allows estimating pressures for majoritic garnet inclusions beyond the clinopyroxene-out reaction. When applied to inclusions in diamonds, differences between pressure estimates obtained with the Thomson et al. (2021) and Beyer and Frost (2017) barometers lie in the range –2.5 to +6 GPa. Whatever the majorite barometric method used, if the garnet reequilibrated with clinopyroxene after entrapment, as is typically the case in unmixed inclusions (see below), underestimation of P will occur.Elastic methods. When an inclusion is entrapped in a diamond, both the inclusion and the diamond are initially at the same pressure. On eruption, the pressure acting on the diamond drops to 1 atm, but a residual pressure, Pinc, of up to several GPa may develop on the inclusion as a consequence of the inclusion and host having different elastic properties. If we know the Pinc and the elastic properties of the inclusion and host, we can back-calculate the conditions under which the two minerals would be at the same pressure. These conditions describe a line in P–T space, known as the entrapment isomeke, and the entrapment conditions (i.e., the diamond formation P–T) will be one of the points along the isomeke. This forms the basis for elastic barometry of diamonds. The principles, problems and on-going developments of this method are described in detail in Angel et al. (2022, this volume).Projection onto a geotherm. When only thermometric estimates are available due to a lack of suitable barometers, a tentative estimate of pressure can be obtained by projecting temperature estimates on a reference geotherm. This will most conveniently be a geotherm derived by fitting P–T data for mantle xenoliths or xenocrysts from the same kimberlitic source. The FITPLOT computational software by Mather et al. (2011) provides a useful aid to model geotherms from xenolith P–T data. There may be significant uncertainties in some of the input parameters that are used for the modeling, specifically the thickness and heat production of the upper and lower crust and the mantle potential temperature, but in most cases these uncertainties will not significantly affect the shape of the geotherm within the lithospheric diamond window, which will be chiefly constrained by the xenolith P–T data. Nonetheless, assuming different mantle potential temperatures will affect the estimated thickness of the lithosphere and the depth extent of the conductive portion of the geotherm. This may have some relevance for some old, high-T diamonds, as mantle potential temperatures may have varied over geological time. If this variation is not considered, thermometric estimates may unduly be projected onto the adiabatic portion of the geotherm and lead to severe overestimation of P. Values for the present-day mantle potential temperature and its change with time can be found in Katsura et al. (2010) and Ganne and Feng (2017). The output data that describe the relevant section of the model geotherm can conveniently be fitted through a polynomial expression to derive P as a function of T. The final P–T estimate will be given by the intersection of this polynomial with the isotherm described by the chosen thermometer. Note that this isotherm will not generally be a flat line in a P–T plot, due to the generally significant P-dependence of the thermometers (Fig. 4).The basic assumption underlying this approach is that the inclusions last equilibrated at P–T conditions lying on the mantle geotherm at the time of eruption. This is generally acceptable for touching inclusion pairs, which can reequilibrate in mostly the same way as touching minerals in xenoliths (but see caveat in section below dedicated to touching vs. non-touching inclusions), whereas it is much less obviously so for non-touching inclusions. Comparisons between P–T estimates for inclusions and xenoliths from the same sources showed that the assumption often holds also for non-touching inclusions (Nimis 2002). Nonetheless, there is also evidence of some diamond inclusions recording conditions significantly hotter or colder than the xenolith geotherm (Griffin et al. 1993; Sobolev and Yefimova 1998; Nimis 2002; Stachel et al. 2003, 2004; Weiss et al. 2018; Nimis et al. 2020). Also, multiple non-touching inclusions within individual diamonds may record a range of conditions, indicating thermal fluctuations during the growth history of the diamond (Griffin et al. 1993). Projection onto a geotherm should therefore be used with caution and relatively large uncertainties should be allowed. Assuming a maximum possible T difference from a xenolith geotherm of ~250 °C, consistent with available estimates for diamonds using single-clinopyroxene (Nimis and Taylor 2000), single-garnet (Canil 1999) and orthopyroxene–garnet (Harley 1984) thermometers, projection on a typical cratonic geotherm may result in a P uncertainty of up to ~1.5 GPa (Fig. 4).Projection onto the adiabatic portion of the mantle geotherm may provide a necessary independent constraint for the application of elastic barometric methods to sublithospheric diamonds (e.g., Anzolini et al. 2019).Mineral stability. Lithospheric mantle rocks provide little opportunity to estimate conditions of diamond formation based only on the stability of mineral assemblages. For example, across the entire diamond window, the only significant mineralogical change is the spinel-to-garnet transition in peridotite. The spinel-to-garnet reaction is not univariant and in depleted cratonic peridotites with strongly elevated Cr/Al, spinel + garnet assemblages may persist over a large range of P–T conditions (e.g., MacGregor 1970; Webb and Wood 1986; Klemme 2004; Ziberna et al. 2013). In general, spinel is stabilized to higher pressures in more refractory, Cr-rich and Al-poor bulk compositions. This allows magnesiochromite to be incorporated as inclusions in diamonds that were formed in very deep, refractory environments (e.g., 6.5 GPa, corresponding to a depth of over 200 km, for a diamond from the Udachnaya kimberlite, Siberia; Nestola et al. 2019a).Mineral stabilities become more interesting in the sublithospheric mantle, where a number of characteristic mineralogical changes occur (Harte 2010; Harte and Hudson 2013). Although these changes do not permit bracketing of P–T conditions with a resolution comparable to that of conventional thermobarometers, in some cases they constrain the formation of a diamond to within specific mantle depth regions. For instance, the coexistence of Fe-bearing periclase with enstatite, interpreted to be inverted bridgmanite (previously known as MgSi-perovskite), has long provided evidence that some diamonds were formed at lower-mantle depths (Scott Smith et al. 1984; Moore et al. 1986). Even in the absence of periclase, inverted bridgmanite from the lower mantle can readily be distinguished from upper-mantle enstatite by its very low Ni contents (Stachel et al. 2000). The finding of a rare inclusion of ringwoodite in a diamond from Juina, Brazil, unequivocally demonstrated its origin from the mantle transition zone (Pearson et al. 2014). Also, the reconstructed compositions of a suite of exsolved inclusions in Juina diamonds nicely matched those of minerals expected to form in a basaltic system in the lower mantle (Walter et al. 2011).However, using incomplete mineral assemblages found as inclusions in diamonds as depth markers is not always free from ambiguity. For instance, inclusions of breyite (previously known as CaSi-walstromite) have long been considered to represent retrogressed CaSiO3-perovskite from depths greater than ~600 km (e.g., Harte et al. 1999; Joswig et al. 1999), but there is compelling evidence that at least some breyite inclusions originated at much shallower depths within the upper mantle (Brenker et al. 2005; Anzolini et al. 2016). Woodland et al. (2020) recently provided experimental support for a possible upper-mantle origin of breyite. Also, Fe-bearing periclase is the most common inclusion of interpreted lower-mantle origin (Kaminsky 2012), but there is evidence that Fe-bearing periclase participates in mineral parageneses straddling the upper mantle-lower mantle boundary (Hutchison et al. 2001) and that it may be a co-product of diamond-forming reactions that may occur in the transition zone and even in the overlying upper mantle (Thomson et al. 2016; Bulatov et al. 2019).Data quality. A sometimes underrated prerequisite for robust thermobarometry is precise and accurate chemical analyses. Inclusions in diamonds are often small and their chemical analysis can be challenging. Moreover, some thermobarometers may be particularly sensitive to analytical uncertainties on elements that occur at critical concentration levels. For instance, Ziberna et al. (2016) explored propagation of analytical errors on P estimates using the single-clinopyroxene barometer of Nimis and Taylor (2000) and found that routine electron microprobe analytical conditions may be insufficient for meaningful barometry of many inclusions in diamonds. Clearly, not only poor counting statistics but also bad sample preparation and improper standardization can be an issue and may lead to spurious results. A survey of published chemical data for over 600 clinopyroxene inclusions in lithospheric diamonds reveals that ~30% of them have oxide total weight percentages < 98.5% or > 101%, or cation sums < 3.98 or > 4.03 atoms per formula unit and, thus, are definitely not of good quality! Note that the above cut-offs generously take into account the possible effects of considering all Fe as Fe2+ in electron microprobe analyses.The next critical step when studying inclusions that were separated from coexisting mineral phases is the definition of their original paragenesis. Even if single-mineral thermobarometers may often come in handy, they invariably assume that the investigated mineral last equilibrated with specific other minerals. As for clinopyroxene and garnet, which are major constituents in several types of mantle rocks, simple compositional screening based on either major or trace element concentrations may help to discriminate between the most typical paragenetic varieties (e.g., Griffin et al. 2002; Grütter et al. 2004; Nimis et al. 2009; Ziberna et al. 2016). Unusual compositions that are not sensitively discriminated by available classification schemes or fall outside the compositional range over which the thermobarometers were calibrated, as well as false positives that unduly survive compositional filtering, may undermine the reliability of thermobarometric data. This problem may be minimized by investigating large datasets, when this is permitted by the number and nature of the available samples.Even when all cautionary steps have been made to select suitable compositional data, thermobarometric estimates are subject to errors. As discussed in the previous section, some of these errors are systematic and lead to predictable inconsistencies between estimates obtained using different methods. This fact should always be considered when comparing data obtained using different thermobarometers or even different formulations of the same thermobarometer. Thermobarometric methods with demonstrated internal consistency must be prioritized if the purpose is to explore the distribution of data for heterogeneous inclusion populations.In this respect, thermobarometry of eclogitic diamonds remains problematic. Considering the compositional limitations required by the ‘best’ available barometric methods (i.e., SiCpx < 1.985 apfu and NaCpx < 0.5 apfu), ~45% of over 200 reported eclogitic clinopyroxene–garnet inclusion pairs are automatically excluded. In addition, the reliability of barometric methods for eclogitic inclusions is questionable (see above).When P estimates are too uncertain or not available, for example, due to a lack of suitable barometers for the specific inclusions investigated, projection of temperatures onto the local xenolith geotherm can provide a last-resort solution to estimate P. The resulting uncertainties can be too large for an accurate characterization of genesis conditions for individual diamonds, but the procedure can have some utility for exploring general trends in large datasets (e.g., Korolev et al. 2018; Nimis et al. 2020). It probably remains the safest available method for barometry of eclogitic inclusions, notwithstanding potential over- or underestimation for diamonds formed at conditions hotter or, respectively, colder than ambient mantle at the time of eruption. A fairly popular alternate has been to calculate T for these diamonds at a fixed P of 5 GPa, approximately corresponding to the average P for diamonds worldwide. This method is not as effective and may overcompress the range of T estimates, due to the generally significant P-dependence of the thermometers (Fig. 4).Touching vs. non-touching. When a pair of touching mineral grains included in a diamond reequilibrate to changing external conditions, their total encapsulated volume remains constrained by the surrounding diamond. This forces the included minerals to follow a distinct pressure path and to achieve final compositions that are different from those of the same minerals outside the diamond (see Angel et al. 2015 and Ferrero and Angel 2018 for more general discussion of this phenomenon). The most common changes in ambient conditions are probably isobaric cooling or heating after diamond growth, in response to secular variations in mantle thermal state or following short-term thermal pulses related to diamond-forming processes. In most cases of diamond-hosted inclusions, cooling will cause a decrease in the inclusion P even if the external P remains constant, leading to underestimation of the actual diamond P using chemical thermobarometers. The extent of this underestimation can be calculated if the thermoelastic properties of the inclusion and host minerals are known (Angel et al. 2015). For various combinations of garnet, pyroxene and olivine inclusions in diamond, the error would amount to ~0.3 GPa for a temperature drop of 100 °C. Heating will have the opposite effect. These errors are of a similar magnitude as those of the orthopyroxene–garnet and single-clinopyroxene barometers and may thus contribute to some of the scatter in P–T plots for touching inclusions. The resulting errors may be lower, however, if plastic deformation in the diamond is able to reduce the under- or overpressures developed on the inclusions.If some thermal reequilibration occurs, the use of single-mineral thermobarometers on touching inclusions may generate further errors when some key minerals are not part of the inclusion assemblage. For instance, touching inclusions of clinopyroxene-orthopyroxene devoid of garnet will reliably record the ambient thermal state at the time of eruption, but will be poor candidates for single-clinopyroxene barometry. In response to a thermal change, the clinopyroxene will easily reequilibrate with orthopyroxene, but not with garnet, and so a false apparent encapsulation P will be calculated at the final equilibrium T (Nimis 2002). Touching inclusions of clinopyroxene ± garnet ± olivine would instead produce estimates close to the diamond formation conditions, because the composition of the clinopyroxene in orthopyroxene-free assemblages would be little affected by a variation in temperature. Polymineralic inclusions consisting of touching clinopyroxene-orthopyroxene–garnet ± olivine will fully reequilibrate and will provide P and T conditions close to those at the time of eruption (except for the small systematic errors described in the previous paragraph).Non-touching inclusions are immune to the above problems, because the inert diamond protects them from post-entrapment reequilibration, and are unequivocally the best candidates to determine the diamond formation P–T, provided reliable monomineral thermobarometers can be used. In all other cases, non-touching inclusion pairs are prone to potential errors, because the individual inclusions may have been incorporated at different times and under different conditions (see above) and may thus not be in mutual equilibrium. The largest T difference between multiple inclusions in individual diamonds was reported for garnets in a diamond from the Mir kimberlite, Siberia (Griffin et al. 1993). The calculated difference varies greatly depending on which calibration of the Ni-in-garnet thermometer is used, i.e., ca. 400 °C with the Ryan et al. (1996) calibration or ca. 230 °C with the Canil (1999) calibration. Disequilibrium amongst non-touching inclusions in diamonds, however, appears to be the exception rather than the rule, as multiple inclusions in individual diamonds most often yield similar P–T estimates once the systematic deviations between different thermobarometers are taken into account (Stachel and Harris 2008; Stachel and Luth 2015).Syngenesis vs. protogenesis. For several decades, thermobarometry of inclusions in diamonds has relied on the assumption of syngenesis, i.e., the inclusions formed or completely recrystallized during the growth of diamond. A corollary of this assumption was that the inclusions record the conditions of diamond formation, excepting for potential post-entrapment modifications in touching inclusions. The main proof for syngenesis was that diamond typically imposed its shape on the inclusions. This widely accepted paradigm was challenged by Nestola et al. (2014), who found evidence of protogenesis (i.e., formation before the diamond) for some olivine inclusions that would have been classified as syngenetic based on morphological criteria. Similar evidence was then found for other olivine, clinopyroxene, garnet and magnesiochromite inclusions (Milani et al. 2016; Nestola et al. 2017, 2019b; Nimis et al. 2019). The possibility that inclusions in diamonds are protogenetic suggests that they might record conditions predating diamond formation. This might be particularly relevant for isolated inclusions, if their host diamonds were formed from hot fluids or melts intruding through relatively cool mantle rocks. Did the preexisting mineral grains have enough time to fully reequilibrate to the changing thermal conditions during diamond growth and before their incorporation was complete?The time required by a mineral to adjust its composition to new physicochemical conditions strongly depends on temperature and grain size, and can be estimated from diffusion kinetics. For instance, a clinopyroxene grain with a diameter of 200 µm would fully reset its Ca/Mg ratio (the main temperature sensor in pyroxene thermometry) in less than 100,000 yr above 1200 °C and in a few m.y. below 900 °C (Fig. 7). Whether these time spans are short or long enough for complete reequilibration depends on the diamond growth rate, which is not precisely known. It is tempting to infer that incomplete reequilibration of protogenetic inclusions, especially the largest and shallowest (i.e., coldest) ones, may account for some of the scatter in P–T plots for non-touching inclusions (Fig. 8). Also, the similarity of P–T conditions occasionally observed for inclusions in diamonds and xenoliths from the same sources may alternatively be ascribed to diamond formation from thermally-equilibrated media (Nimis 2002) or to lack of reequilibration of protogenetic inclusions during short-lived diamond-forming processes (Nestola et al. 2014).Exsolved inclusions. Some inclusions in diamonds show evidence of post-entrapment unmixing. Reported examples are clinopyroxenes with orthopyroxene exsolution lamellae (± coesite), representing unmixed high-T clinopyroxenes, and intergrowths of more or less majoritic garnet and clinopyroxene, representing unmixed high-P majoritic garnets.Exsolved clinopyroxene inclusions are uncommon and have only been reported in diamonds from Namibian placers and from the Voorspoed kimberlite, South Africa (Leost et al. 2003; Viljoen et al. 2018). The coexistence of clinopyroxene and orthopyroxene makes these inclusions ideal for pyroxene thermometry, but the exsolved assemblage does not provide barometric estimates. Since pyroxene thermometers have a small P-dependence, projection onto the local xenolith geotherm provides in these cases reasonable estimates of the depth of origin of the diamonds (Nimis et al. 2020). If the compositions and relative proportions of the exsolved phases are known (e.g., by combining electron microprobe data with 2D or, better, 3D image analyses), the integrated pre-exsolution composition of the clinopyroxene can be reconstructed. The integrated composition can provide a T estimate at the time of diamond growth, assuming equilibrium with a separate orthopyroxene during entrapment, or a minimum T estimate of diamond formation, in the absence of orthopyroxene. In the reported cases, thermometry of the reconstructed clinopyroxenes provided evidence for diamond formation at temperatures near the mantle adiabat and ~200 °C hotter than the ambient mantle at the time of eruption. Attempts to apply the single-clinopyroxene barometer to the reconstructed compositions yielded unreasonably low pressure estimates, suggesting formation in strongly modified mantle environments (Nimis et al. 2020).Exsolution textures are common in majoritic garnet inclusions in sublithospheric diamonds (Moore and Gurney 1989; Wilding 1990; Harte and Cayzer 2007). These textures are interpreted to reflect decompression during slow ascent of the diamonds in convecting mantle or mantle plumes. The presence of exsolutions implies that pressure estimates based on majoritic garnet compositions are minima. Using reconstructed pre-exsolution garnet compositions will yield conditions closer to those of diamond formation. Even so, the resulting estimates may be minima, because the garnet may have originated at depths beyond the clinopyroxene-out reaction (see previous Barometry section).Figure 8 shows a compilation of P–T data for lithospheric diamonds from worldwide sources, obtained using the most robust available thermobarometer combinations. The compilation comprises clinopyroxene inclusions in 100 lherzolitic diamonds and orthopyroxene–garnet inclusion pairs in 77 harzburgitic, 12 lherzolitic and 10 websteritic diamonds. The great majority of the clinopyroxene inclusions were isolated within their host diamonds and should generally record conditions close to those of diamond formation. The orthopyroxene–garnet pairs consist of both touching and non-touching assemblages. The touching pairs may have reequilibrated after their incorporation and should provide conditions close to the final conditions of residence in the mantle. In addition, P–T data were calculated, using the same methods, for diamondiferous xenoliths for which suitable mineral analyses were available.In comparing the two plots for single-clinopyroxene and orthopyroxene–garnet thermobarometry (Fig. 8a,b), it is important to consider the known systematic inconsistencies between the thermobarometric methods used (see above). The orthopyroxene–garnet thermometer tends to overestimate temperature below 1100 °C and to underestimate above 1100 °C. This accounts at least in part for the more restricted temperature range obtained for the orthopyroxene–garnet inclusions. Nonetheless, disregarding a single obvious outlier, all P–T values fall within 0.3 GPa of the diamond stability field, confirming the general reliability of the thermobarometers. Most of the P–T results lie between the 35- and 40-mW/m2 conductive geotherms of Hasterok and Chapman (2011), i.e., at conditions typical for cratonic lithospheres. Most exceptions record higher temperature conditions, which in the case of clinopyroxene may reach the mantle adiabat. All of these exceptions consist of non-touching inclusions. A possible explanation is that several diamonds were formed from thermally non-equilibrated media derived from the sublithospheric convective mantle. Another potential source of mantle temperature inhomogeneity during diamond formation can be incomplete cratonization at the time of diamond growth. For instance, in the Kaapvaal craton a stable geothermal gradient was established by about 2.5 Ga and followed by secular cooling (e.g., Michaut and Jaupart 2007; Brey and Shu 2018) and re-heating (Griffin et al. 2003), but many Kaapvaal diamonds are significantly older than 2.5 Ga (see review in Shirey and Richardson 2011) or may have formed during important perturbations of the local lithosphere (Griffin et al. 2003; Korolev et al. 2018). Therefore, it is not surprising that many of these diamonds may record conditions hotter or colder than those at the time of kimberlite eruption (e.g., Nimis et al. 2020).The touching orthopyroxene–garnet inclusion pairs deserve further discussion. In the available dataset, the majority of them (23 out of 31) refer to diamonds from the De Beers Pool (Kimberley, South Africa). These inclusions were studied by Phillips et al. (2004), who carefully documented differences between P–T estimates recorded by touching vs. non-touching inclusions. These workers found that the non-touching inclusions yielded higher P–T values than the touching inclusions and ascribed this observation to post-entrapment reequilibration of the touching inclusions under decreasing T conditions, possibly accompanied by mantle uplift and decompression. They also correctly noted that post-entrapment elastic modifications were not of sufficient magnitude to reconcile the observed P–T differences between touching and non-touching inclusions. However, Nimis et al. (2020) observed that both touching and non-touching inclusions in the Phillips et al. (2004) dataset record a colder geothermal regime than mantle xenoliths and xenocrysts in the same kimberlites, casting doubts on the secular cooling hypothesis. Whereas the relatively cold signature of the non-touching inclusions might reflect old conditions pre-dating the final thermal equilibration of the lithosphere recorded by the xenoliths, the data for the touching inclusions are difficult to interpret. Weiss et al. (2018) speculated that the low temperatures recorded by the touching inclusions could be related to late infiltration of cold, slab-derived fluids, remnants of which were preserved in some cloudy and coated diamonds from the same locality. Yet it remains unclear why infiltration of these fluids should leave virtually no traces in the sampled mantle xenoliths.Touching orthopyroxene–garnet inclusions from other localities are probably too few to be statistically significant. Nonetheless, in all these cases the touching inclusions yield temperatures significantly lower than non-touching inclusions in other diamonds from about the same depth (Fig. 8b). Additional thermobarometric evidence that diamond formation may be followed by cooling was provided by Stachel and Luth (2015) and Viljoen et al. (2018). However, Stachel and Luth (2015) also described touching inclusions yielding temperatures similar to or even slightly higher (up to 40 °C) than non-touching inclusions in the same diamonds.Diamond’s depth systematics. Irrespective of the nature of the inclusions and the thermobarometric methods used, the frequency distribution of pressures for lithospheric diamond is essentially unimodal, with a mode at ~6 GPa, corresponding to depths around ~190 km (Fig. 8; see also Stachel and Harris 2008 and Stachel 2014 for broadly similar distributions obtained using different thermobarometers and partly different datasets). The slight differences between the pressure distributions obtained for clinopyroxene and orthopyroxene–garnet inclusions may largely be ascribed to recognized inconsistencies between the different thermobarometer combinations used and, perhaps, to sampling bias. Therefore, these differences are probably not significant. The hard constraints imposed by the graphite–diamond and lithosphere-asthenosphere boundaries cannot fully explain the observed depth distributions, which show decreased diamond frequencies approaching these boundaries (Fig. 8). Nimis et al. (2020) obtained very similar depth distributions with depth modes at 180 ± 10 km for diamonds from three individual South African kimberlite sources by using various combinations of thermobarometric methods on different inclusion types. They also found no clear correlation with the depth distributions of mantle xenocrysts from the same sources calculated using the same thermobarometric methods. They concluded that the observed diamond depth distributions are unrelated to the kimberlite sampling efficiency and, thus, have genetic meaning and likely global significance.The global significance of the observed depth distributions is also supported by the abundant TNi-in-garnet data for garnet inclusions (Fig. 9). These data show a mode at ~1200 °C, but no pressure constraints are generally available. Assuming that the garnets last equilibrated at ‘average’ conditions between the 35- and 40- mW/m2 model geotherms, similar to the clinopyroxene and orthopyroxene–garnet inclusions (Fig. 8), the Ni-in-garnet temperature mode would correspond, again, to a pressure mode of ~6 GPa and a depth mode of ~190 km. The gentler increase of frequencies to ~1200 °C and their steeper decrease at higher temperatures (Fig. 9) also mirror the frequency variations with increasing pressure to ~6 GPa and beyond exhibited by the clinopyroxene and orthopyroxene–garnet inclusions (Fig. 8).The diamond depth distributions in the lithosphere may be the result of both constructive and destructive processes (Nimis et al. 2020). In particular, the progressive decrease in diamond concentrations at depths shallower than ~170–190 km may reflect (i) decreasing precipitation rates for diamond from ascending parental media or (ii) progressive decrease in the amount of these deep-sourced media at shallower levels. Probably, both mechanisms concur to limit the diamond endowment in the shallow portion of the diamond window. The rapid decrease of diamonds near the base of the lithosphere may reflect (i) a positive balance between the opposed effects that an increase of P and an increase of T may have on carbon solubility in some mantle fluids/melts along specific P–T paths or (ii) a progressive decrease in diamond endowment due to reactions with carbon-undersaturated asthenospheric fluids/melts. These two alternative hypotheses are difficult to prove or disprove, because our knowledge of the behavior of natural carbon-bearing melts and fluids at high P–T is still limited. The prevalence of melt-metasomatized lithologies near the base of typical lithospheres coincides with diminished diamond frequency and suggests that melt-driven metasomatism may be an important destructive agent for diamond (e.g., McCammon et al. 2001). Remobilization of carbon from these deep lithospheric roots may eventually contribute to build up the diamond concentration at shallower depths (Nimis et al. 2020).The recognition that depths around ~190 km are particularly favorable for diamonds may have interesting implications for their mineralogical exploration. Areas characterized by the occurrence of xenocrystic mantle minerals (e.g., clinopyroxenes or garnets) from these particular depths may be considered as high-priority targets. Nonetheless, since the xenocrysts and diamond depth distributions may be uncorrelated, even limited xenocryst records from the diamond’s most favorable depth range may be indicative of significant diamond potential (Nimis et al. 2020).Figure 10 shows the global distribution of pressure estimates for diamonds containing majoritic garnets. At present, these are the only types of inclusions that provide abundant, quantitative barometric data for sublithospheric diamonds. A difficulty in investigating these data lies in the fact that they may represent minimum P estimates. As explained in the previous section, this limitation is due to the frequent occurrence of exsolutions and, for all but the Thomson et al. (2021) barometer, to the possible absence of accompanying clinopyroxene in the original mineral assemblage from which the majoritic garnet inclusion was separated on entrapment. If one considers only inclusions that are reportedly free of clinopyroxene exsolutions, two distinct unimodal distributions are obtained for peridotitic and eclogitic/pyroxenitic garnets, respectively (Fig. 10). Using the Beyer and Frost (2017) barometer, the peridotitic inclusions are mostly clustered in the 175–225 and 225–275 km depth intervals, near the lithosphere-asthenosphere boundary. Most of these inclusions would thus more appropriately be classified as deep lithospheric. The eclogitic/pyroxenitic inclusions are, on average, much deeper and mostly fall in the depth interval between 300 and 500 km. If one considers also inclusions that do contain exsolved pyroxenes, the peak of eclogitic/pyroxenitic inclusions at ~400 km becomes more prominent. Using the Thomson et al. (2021) barometer, the distribution for the eclogitic/pyroxenitic inclusions is stretched to slightly higher pressures, with a maximum estimated depth of 613 ± 44 km, whereas many of the peridotitic inclusions are shifted to significantly greater depths, falling well into the sublithospheric region (Fig. 10). Some of the deepest (>450 km) majoritic garnets fall in the region where clinopyroxene disappears and thus the Beyer and Frost (2017) barometer may underestimate pressure for these samples. However, the Thomson et al. (2021) machine-learning barometer, which is immune to this effect, does not suggest a much deeper origin for these garnets (Fig. 10). Differences between depth estimates for these very deep samples using the two alternative barometers are between –65 and +81 km (average +17 km).It should be considered that, in some cases, in the absence of accurate imaging data, the occurrence of fine clinopyroxene exsolutions may have been overlooked. The use of improved BSE and EBSD imaging might allow recognition of exsolved clinopyroxene in nominally ‘clinopyroxene-free’ majoritic inclusions (cf. Harte and Cayzer 2007; Zedgenizov et al. 2014). The thirteen Juina majoritic inclusions for which no coexisting clinopyroxene has been reported to date give depth estimates in a restricted range between 353 and 455 km (using the Beyer and Frost 2017 barometer) or 368 to 462 km (using the Thomson et al. 2021 barometer) and are in part responsible for the prominent peak at ~400 km in Figure 10. If these inclusions also contained exsolutions and the available electron microprobe data do not faithfully reflect their original compositions, then the overall peak could be shifted to somewhat greater depths or, at least, the overall depth distribution would be smoother than shown in Figure 10. Thomson et al. (2021) suggested that all majoritic garnet inclusions in South American diamonds probably contain exsolutions. They also noted that in the eight inclusions for which both exsolved and reconstructed bulk compositions were available the amount of exsolution is a simple function of the calculated pressure. Accordingly, they devised an empirical correction that can be applied to all inclusions for which reconstructed bulk compositions are not available. When this correction is applied to South American diamonds, the peak for eclogitic majoritic garnet inclusions is smoothed and shifted to ca. 100 km greater depth. It remains unclear if this correction, which is based on a restricted number of inclusions, has any general validity at either global or local scale. In any case, the high abundance of depth estimates between ~300 and ~500 km, even for inclusions that are certainly clinopyroxene-free or whose original composition was reconstructed, suggests that most of the eclogitic/pyroxenitic majoritic garnets indeed originated from this depth range (Harte 2010), or perhaps from a slightly more extended range of ~300 to ~600 km if the Thomson et al. (2021) correction for South American diamonds is valid. This range overlaps with independent depth estimates obtained from phase stability constraints for other sublithospheric diamonds containing Ca-silicate inclusions (300–360 km; Brenker et al. 2005; Anzolini et al. 2016) and with a minimum depth estimate obtained from elastic barometry for a single ferropericlase inclusion (~450 km; Anzolini et al. 2019). Of further interest, the 300–600 km range corresponds to the depths at which carbonated MORB slabs are most likely to intersect their solidus, providing an ideal environment for focused melting and potential diamond growth (Thomson et al. 2016, 2021). Considering that many periclase-magnesiowüstite inclusions, for which barometric data are not available, may also have formed from subduction-related melts (Thomson et al. 2016; Nimis et al. 2018) the actual proportion of diamonds from the 300–600 km interval may be even larger than shown in Figure 10. The reader is referred to Walter et al. (2022, this volume) for a thorough discussion on these and related aspects.A second favorable region for diamond formation is probably located near the upper mantle–lower mantle boundary and uppermost lower mantle (Harte 2010). Evidence for diamond formation in this region is provided by a number of key mineral associations within individual diamonds, mostly consisting of various combinations of periclase, (Mg,Fe)SiO3, (Mg,Fe)2SiO4, jeffbenite (formerly known as TAPP), breyite, corundum and SiO2, and of a NaAl-pyroxene phase (believed to have formed as highly majoritic garnet) alongside with periclase or jeffbenite (Hutchison et al. 2001; Harte 2010). Although in several diamonds only part of the key mineral assemblages is actually represented, the general consistency between the observed mineral associations led Harte (2010) to suppose that a large number of diamonds could come from the relatively narrow depth range of 600 to 800 km. Most inclusions in diamonds from this depth range are ultramafic, as opposed to the mafic inclusions that dominate in the 300–600 km interval.Mineral inclusions are very rare in diamonds and occur in only about 1% of examined stones, with significant differences between individual localities (Stachel and Harris 2008). Moreover, only a few mantle minerals or mineral assemblages that may occur as inclusions in diamonds are suitable for thermobarometry. Therefore, only a small proportion of natural diamonds may yield robust information on the P–T conditions of their formation, and these diamonds are mostly peridotitic. This proportion may increase in the future, in parallel with the development of new thermobarometric methods or refinement of existing ones. A significant field of potential and desirable improvement concerns the eclogitic inclusions, which represent a major fraction of inclusions in lithospheric diamonds, but for which existing barometric methods still provide inconsistent results. A dedicated test on clinopyroxene–garnet pairs similar to that performed on chromian diopsides by Ziberna et al. (2016) might shed more light on the minimum analytical quality that is required for reliable barometry of these inclusions. Extending the applicability of non-traditional methods, such as elastic barometry, to non-ideal configurations and to a more diverse list of inclusion minerals might also have a dramatic effect on the number of diamonds suitable for thermobarometry, although the need for elastically preserved inclusions generally restricts applications to relatively small inclusions not surrounded by fractures. An interesting advantage of the elastic method is that it can rely on non-destructive techniques and, thus, can leave the samples intact and ready for further investigation with other methods. Such investigations require the use of undamaged samples in which internal strains are well preserved. Legacy destructive techniques used to expose minerals included in diamond, unfortunately, prohibit future examination of these unique samples for purposes of elastic barometry.Reanalysis of some previously investigated inclusions or diamondiferous xenoliths might provide a relatively effortless way to increase the number of diamond P–T data. Four of 195 published chromian clinopyroxene analyses and 6 of 156 published orthopyroxene analyses had to be discarded while preparing this review only because of sub-standard analytical quality. Moreover, twenty-five clinopyroxenes were rejected because their compositions fell just off the major-element compositional fields that are used to classify garnet-associated chromian diopsides (Ziberna et al. 2016). Trace element analysis of these diopsides might have allowed more robust discrimination (e.g., based on Sc/V relationships; Nimis et al. 2009) of the garnet-associated samples and many of them might have turned out to be suitable for thermobarometry. Reintegrating all these samples would increase the number of P–T data for chromian diopside-bearing diamonds by a remarkable 27% and for diamonds overall by 17%.An enlargement of the number of P–T data for diamonds would have significant positive outcomes in terms of sample bias reduction. Sample bias is presently a major problem, since the number of localities for which a statistically significant number of sufficiently reliable P–T estimates are currently available is limited at best to three and all are from the Kaapvaal craton (Nimis et al. 2020). Consequent benefits of sample bias reduction would be a better assessment of diamond depth distributions at individual localities, and a more robust evaluation of diamond P–T systematics and of their significance in terms of diamond forming processes and preservation.If the number of P–T data for diamonds is important, their internal consistency may be even more so. In fact, most of the combinations of thermobarometric methods that may be applied to diamond inclusions do show some degree of inconsistency with one another. These inconsistencies mostly stem from simplified thermodynamic treatment of solid solutions in minerals and from uncertainties in the data used for their calibration and derived either from experiments or from independent xenolith thermobarometry. The internal consistency of thermobarometric methods is particularly critical when comparing data for different inclusion types, as commonly is necessary in diamond studies. This is another field of potential scientific advance that may eventually contribute to an improved understanding of diamond forming processes.This review is the result of the knowledge, expertise and generosity of many people. Fruitful collaboration through the years with many of them, and particularly with W.R. Taylor, H. Grütter, L. Ziberna, F. Nestola and R.J. Angel, has been of decisive importance in producing my personal contributions to diamond thermobarometry. Yet many more people have in fact contributed to this review through their enormous scientific work and by providing the community with invaluable data over about five decades. Most of them appear in the reference list below. I express my sincere gratefulness to all of these people and, in particular, to T. Stachel, who generously shared a diamond inclusion database, which served as a useful basis to build the updated database used in this work. I am also grateful to A. Thomson, for help in majorite barometry calculations, and to H. Grütter, G. Brey and editor T. Stachel for their very helpful and thorough reviews and comments.","PeriodicalId":501196,"journal":{"name":"Reviews in Mineralogy and Geochemistry","volume":"63 8","pages":""},"PeriodicalIF":0.0000,"publicationDate":"2022-07-01","publicationTypes":"Journal Article","fieldsOfStudy":null,"isOpenAccess":false,"openAccessPdf":"","citationCount":"0","resultStr":"{\"title\":\"Pressure and Temperature Data for Diamonds\",\"authors\":\"Paolo Nimis\",\"doi\":\"10.2138/rmg.2022.88.10\",\"DOIUrl\":null,\"url\":null,\"abstract\":\"One of the key scientific questions about diamonds is “how are they formed?” To answer this question, we need to know the diamond-forming reactions and the physicochemical conditions under which these reactions take place. The pressure (P) and temperature (T) of diamond formation are an essential part of this knowledge and their assessment is pivotal to develop predictive scenarios of diamond distribution in the Earth interior. These scenarios may contribute to our understanding of global Earth processes, such as the long-term carbon cycle, and might also eventually improve our capability to select potential targets for diamond exploration (Shirey et al. 2013; Nimis et al. 2020).The evaluation of the P and T of diamond formation can be carried out at two levels of investigation. The first is concerned with formation conditions for individual diamonds or small populations of diamonds from specific sources. This approach has been so far the most widely practiced. The second level considers the statistical distribution of P–T conditions for diamond formation at either local or global scale. This type of investigation is hampered by the difficulty of obtaining large sets of suitable samples from a specific locality or for a statistically significant number of localities, and is therefore unavoidably affected to some degree by sampling bias. Despite inherent limitations, the latter approach is the most appropriate to reveal systematics in diamond P–T distributions and, ultimately, in diamond depth distribution within the Earth.Early reviews of P–T distributions for lithospheric diamonds were made by Nimis (2002), based on thermobarometry of chromian diopside inclusions, and by Stachel and Harris (2008) and Stachel (2014), using a more comprehensive set of thermobarometers. More recently, Nimis et al. (2020) investigated diamond depth distributions for a set of South African kimberlites and provided evidence for systematic trends of likely global significance. The depth distribution for sublithospheric diamonds worldwide was reviewed by Harte (2010). In this contribution, I first describe the methods that can be used to estimate the P–T conditions of diamond formation, highlighting their respective strengths and weaknesses. I then review existing diamond P–T data and their implications for diamond distribution with depth from both a local and a global perspective.Thermobarometry of diamonds can be carried out by estimating P–T conditions of chemical or elastic equilibrium of mineral inclusions contained within them. With some assumptions, the aggregation state of nitrogen substituting for carbon in the diamond lattice can also be used as a thermometer. In some cases, it is possible to derive both P and T estimates for a diamond by combining independent thermobarometric methods. In most instances, however, either P or T estimates can be directly retrieved with sufficient confidence. Below is a list of currently available methods for diamond thermobarometry. Those based on chemical equilibria are specific for different minerals or mineral pairs. Since many inclusions in diamonds occur as isolated grains, detached from their original mineral assemblage, particular emphasis is given in this review to methods that use single minerals to retrieve P and/or T. The following descriptions emphasize thermobarometry applications rather than principles, and the reader is referred to the original publications for details of the specific thermobarometers.Pyroxenes. Many ultramafic mantle rocks contain both clinopyroxene and orthopyroxene. This mineral pair forms one of the most widely used geological thermometers, which is based on the net-transfer reaction of the enstatite component between the two pyroxenes MgSiO3(Opx) ↔ MgSiO3(Cpx). The very small P-dependency of this reaction limits propagation of errors when T is calculated by iteration in combination with an independent barometer. Several versions of this thermometer have been proposed in the literature. Tests based on experimental results and pyroxene compositions in natural xenoliths suggest that the calibration of Taylor (1998) is the most reliable for mantle peridotites and pyroxenites in a T range of 700 to at least 1400 °C (Nimis and Grütter 2010), which covers the entire range of the lithospheric diamond window. The much more popular Brey and Köhler (1990) calibration gives similar results for clinopyroxenes with relatively low Na contents around 0.05 atoms per 6-oxygen formula unit (apfu), but was proven to overestimate with increasing NaCpx (e.g., +150 °C at NaCpx = 0.25 apfu; Nimis and Grütter 2010). For inclusions in diamonds and diamond-bearing peridotitic xenoliths (average NaCpx = 0.12 apfu, median = 0.09 apfu, standard deviation = 0.08 apfu), the Brey and Köhler (1990) calibration will yield estimates that are on average ~50 °C higher that those using the Taylor (1998) calibration. Larger discrepancies may be expected for unusually Na-rich clinopyroxenes.Since clinopyroxene–orthopyroxene inclusion pairs are very rare in diamonds, it is generally more practical to use the alternative single-clinopyroxene version of Nimis and Taylor (2000). The two-pyroxene Taylor (1998) and single-clinopyroxene Nimis and Taylor (2000) methods provide almost indistinguishable results when applied to mantle ultramafic rocks that contain both pyroxenes (Nimis and Grütter 2010). Although the Nimis and Taylor (2000) thermometer uses only one pyroxene to calculate T, its application expressly requires that both pyroxenes were part of the mineral assemblage and were in chemical equilibrium. Therefore, this method is only suitable for diamonds belonging to the lherzolitic or websteritic suites and containing clinopyroxene inclusions, either isolated or associated with other minerals. If orthopyroxene was not part of the original mineral assemblage, as is the case for wehrlitic inclusions, single-clinopyroxene thermometry would only provide a minimum T estimate. An anomalously low T estimate (e.g., one that lies much below the local geotherm) may be a sign that the clinopyroxene is wehrlitic. Ziberna et al. (2016) suggested that a Ca/(Ca + Mg) molar ratio of > 0.5 should also be considered as suspicious, as only ~1% of orthopyroxene-saturated mantle clinopyroxenes lie above this value. Nonetheless, discrimination of wehrlitic clinopyroxenes is not generally possible based merely on compositional criteria.Simakov (2008) proposed a different, more complex calibration of the single-clinopyroxene thermometer, which definitely improves performance above 1500 °C, though not over the T < 1400 °C range typical of ‘lithospheric’ temperatures (note that the TNimis and Taylor 2000 estimates reported in Fig. 10 of Simakov 2008 are incorrect and unduly suggest overestimation below 1300 °C).The Ca-in-Opx thermometer of Brey and Köhler (1990) is the single-orthopyroxene version of the pyroxene thermometer and can provide estimates that are complementary to, and independent from those obtained with the single-clinopyroxene thermometer. The two single-pyroxene methods were shown to be mutually consistent, provided a correction is applied below 900 °C (Nimis and Grütter 2010). However, since inclusions of likely harzburgitic (i.e., clinopyroxene-free) affinity are relatively common in diamonds (Stachel and Harris 2008), the possibility that TCa-in-Opx estimates for orthopyroxene inclusions are only minimum T values is much higher than for the single-clinopyroxene thermometer. Therefore, its usefulness is rather limited in diamond studies.Garnet. The Ni content of garnet in equilibrium with olivine is very sensitive to T and apparently independent of P (Ryan et al. 1996; Canil 1999). Since the Ni content of olivine shows small variations in both mantle xenoliths (O’Reilly et al. 1997) and inclusions in diamonds (Griffin et al. 1992; Sobolev et al. 2008), T can be retrieved from Nigarnet alone, by assuming an appropriate Ni content for coexisting forsteritic olivine. A useful reference values for Niolivine is the mean value for mantle olivines (mean ± standard deviation = 2900 ± 360 ppm; Ryan et al. 1996), which is also close to the average value for olivine inclusions in diamonds worldwide (Stachel and Harris 2008; Sobolev et al. 2009). If available, the mean Niolivine value for olivine inclusions in diamonds from the same locality may be used (e.g., 3150 ± 200 ppm for 51 inclusions from the Kalahari craton, Griffin et al. 1992; 2700 ppm for 88 inclusions from Arkhangel’sk province, Malkovets et al. 2011). Choosing one or the other value changes the final T estimate by at most a few tens of degrees (Fig. 1). The concentrations of Nigarnet can be measured with a laser-ablation inductively-coupled plasma mass spectrometer (LA-ICP-MS), an ion microprobe (SIMS) or a proton microprobe (PIXE), which typically ensure a precision within a few ppm. The assumption of equilibrium with olivine may require compositional filtering (e.g., Grütter et al. 2004) to exclude any non-peridotitic garnet.The Ni-in-garnet thermometer has seen wide application in studies of Cr-pyrope garnets included in diamonds (e.g., Griffin et al. 1992, 1993; Davies et al. 2004a; Viljoen et al. 2014; De Hoog et al. 2019). Its calibration, however, is somewhat controversial. Ryan et al. (1996) calibrated it against mantle xenoliths, using T values derived from a combination of the olivine–garnet Fe–Mg-exchange thermometer (O’Neill and Wood 1979) and the MacGregor (1974) and Brey and Köhler (1990) orthopyroxene–garnet barometers. Canil (1999) calibrated it against experiments at T ≥ 1200 °C. The two calibrations give identical results at ~1100 °C, but progressively diverge at lower and higher T (Fig. 1). As a result, estimates obtained through the Canil (1999) formulation will typically span over narrower T intervals. Considering the most typical range for Ni in garnet inclusions, the maximum difference TCanil–TRyan is ca. +100 °C at 20 ppm Ni and ca. –250 °C at 180 ppm Ni. It has been variously claimed that either calibration, or even their ‘average’, agree best with other independent thermometers when applied to mantle xenoliths (e.g., Ryan et al. 1996; Canil 1999; De Hoog et al. 2019; Czas et al. 2020; Nimis et al. 2020), but a definitive assessment using internally consistent thermobarometers as reference is still lacking. The latest attempt to refine the Ni-in-garnet thermometer was made by Sudholz et al. (2021a), who recalibrated it against new experiments in a relatively narrow T range (1100–1325 °C) and introduced correction terms for Ca and Cr contents in the garnet. When tested against independent estimates for xenoliths using the Nimis and Taylor (2000) enstatite-in-clinopyroxene thermometer, the Sudholz et al. (2021a) calibration shows improved overall accuracy above 1100 °C relative to the Canil (1999) experimental calibration, but poorer overall precision and slightly stronger progressive overestimation at lower T (see Fig. 7 in Sudholz et al. 2021a).The Mn content of garnet in equilibrium with mantle olivine is sensitive to T (Smith et al. 1991) and can be used as a single-mineral thermometer, assuming a constant Mn content in olivine. This thermometer relies on electron microprobe data and was proposed as a substitute for the Ni-in-garnet thermometer when trace element data are not available (Grütter et al. 1999; Creighton 2009). The declared precision is rather poor (mostly ±150 °C) and, in the absence of independent olivine data, severe underestimation may occur at high T. Therefore, this method may have some use in surveys of large populations of garnet xenocrysts recovered during diamond exploration (e.g., Grütter and Tuer 2009), but its utility for diamond thermobarometry is limited.Ashchepkov et al. (2010) calibrated two monomineral thermometers for garnets in equilibrium with clinopyroxene or olivine, respectively. These methods are simplified versions of Fe–Mg-exchange thermometers for clinopyroxene–garnet or olivine-garnet pairs, in which the clinopyroxene and olivine compositions are modeled from the composition of the garnet. Their accuracy cannot be better than that of the original two-mineral formulations, which are problematic themselves (see below). Also, P must be independently known to calculate T. In a test on mantle xenoliths from the Udachnaya kimberlite, these thermometers reproduced temperatures calculated with the pyroxene thermometer of Brey and Köhler (1990), at pressures calculated with the orthopyroxene–garnet barometer, to within ca. ±100 °C. Additional uncertainties due to problematic estimation of P in the absence of compositional data for coexisting orthopyroxene (see below) probably make these thermometers of little interest for diamond thermobarometry, especially when investigating small populations of diamonds.Olivine. The Al content in peridotitic olivine is sensitive to T and can be used as a thermometer. The first Al-in-olivine thermometer for peridotitic olivine at conditions relevant to diamond was empirically calibrated on mantle xenoliths by De Hoog et al. (2010), using P–T estimates derived by a combination of the two-pyroxene thermometer and orthopyroxene–garnet barometer of Brey and Köhler (1990). The method was recalibrated by Bussweiler et al. (2017) by analysis of olivine from high-P–T experimental charges. The Al content in cratonic mantle olivine ranges from ~1 ppm to ~250 ppm and is conveniently analyzed by techniques such as LA-ICP-MS or SIMS, which ensure a precision within a few ppm. Electron microprobe analysis (EMPA) using very high probe currents (e.g., ≥ 200 nA) and long count times (≥ 100 s) may be a valid alternative (Batanova et al. 2018; D’Souza et al. 2020).Al-in-olivine temperatures for spinel peridotites or strongly metasomatized garnet peridotites may be significantly underestimated or, respectively, overestimated. Compositional screening using an Al vs V diagram may help to discriminate these ‘unsafe’ olivines (Bussweiler et al. 2017). Screening is especially important if olivine is not physically associated with garnet, as occurs with many inclusions in diamonds. The Al-in-olivine thermometer gives systematically higher estimates (~50 °C, slightly increasing below 900 °C) than the single-clinopyroxene thermometer (Nimis and Taylor 2000) for natural mantle peridotites, probably reflecting slight suppression of Al incorporation in the experimental olivines due to Na loss (Bussweiler et al. 2017). The discrepancy is sufficiently small to ensure successful application of the method, but should be considered when comparing estimates obtained using different thermometers. The Al-in-olivine thermometer has not yet been widely used in diamond studies (see Korolev et al. 2018 for an example of application), owing to its recent development and the necessity of non-routine analysis of Al. Considering the great abundance of olivine among inclusions in diamonds (Stachel and Harris 2008), its application is likely to increase significantly in the future.Spinel. The Zn content of spinel in equilibrium with mantle olivine is sensitive to T. Since the Zn content of mantle olivine is nearly constant (mean ± standard deviation = 52 ± 14 ppm), the Zn content of spinel can directly be used as a thermometer (Ryan et al. 1996). The concentrations of Zn in spinel can be measured by SIMS or PIXE, which ensure a precision within a few ppm. The current version of the Zn-in-spinel thermometer was calibrated over the range 680–1180 °C against T for coexisting garnets using the Ryan et al. (1996) version of the Ni-in-garnet thermometer. Therefore, it is subject to at least the same uncertainties as the Ni-in-garnet thermometer (see above).Fe–Mg exchange thermometers. These popular thermometers are based on the distribution of Fe2+ and Mg between two mineral phases and can be used to derive T for inclusions in which garnet is associated with olivine, orthopyroxene or clinopyroxene. Tests using mineral compositions from mantle xenoliths and experiments showed that these thermometers suffer from either low precision or systematic errors or both (see review of Nimis and Grütter 2010). In peridotitic systems, the olivine-garnet and the clinopyroxene–garnet thermometers show the lowest precision, with possible discrepancies exceeding 200 °C. The Harley (1984) orthopyroxene–garnet formulation shows the highest precision, but systematically overestimates at T < 1100 °C and underestimates at T > 1100 °C (by up to ~150 °C, on average, for mantle xenoliths). All these discrepancies may reflect a modest sensitivity to T of the exchange reaction (particularly for olivine-garnet), oversimplification in the solid solution models used (especially for garnet and clinopyroxene), and Fe3+/ΣFe ratios in the minerals different from those in the calibration samples. Matjuschkin et al. (2014) experimentally demonstrated the large effect of neglected Fe3+ on orthopyroxene–garnet and olivine-garnet thermometers. Their experiments in the Na-free CaO–FeO–Fe2O3–MgO–Al2O3–SiO2 system at 5 GPa and 1000–1400 °C suggest that temperature estimates may be significantly improved if calculations are made using Fe2+ instead of total Fe in the garnet. Mössbauer data for orthopyroxene and garnet in peridotitic xenoliths, however, suggest that the partitioning of Fe3+ between these two minerals is affected by pressure and Na content in the orthopyroxene (Nimis et al. 2015). Therefore, a test over a range of relevant pressures in a Na-bearing system would be desirable. Nimis and Grütter (2010) empirically corrected the Harley (1984) thermometer using two-pyroxene thermometry of well equilibrated mantle xenoliths as calibration. The corrected version generally yields lower T estimates at low T and higher T estimates at high T relative to the uncorrected thermometer and may be more accurate for orthopyroxene–garnet pairs equilibrated under ‘average’ mantle redox conditions. On the other hand, it may produce large errors (> to ≫100 °C) if redox conditions are far from ‘average’, especially for strongly oxidized conditions at very high pressure and when both P and T are calculated by iteration in combination with an orthopyroxene–garnet barometer (Nimis et al. 2015). Considering that at deep lithospheric levels diamond is stable over a relatively wide range of redox conditions (∆logfO2 FMQ ≈ –5 to –2), this problem can be serious. Also, application to some xenolith suites tends to produce frequent high-P–T outliers (Grütter, pers. comm.). In spite of its recognized systematic discrepancies, the original Harley (1984) version may still be preferable for diamond thermobarometry due to its lower P dependence, which reduces error propagation during iterative P–T calculations. However, the systematic deviations of the Harley (1984) thermometer at T far from 1100 °C, especially if all Fe is treated as Fe2+, should be considered in data evaluation.The clinopyroxene–garnet thermometer represents the only viable method to determine T from major element compositions of eclogitic inclusions in diamonds. There are many versions of this thermometer in the literature. The most recent calibration (Nakamura 2009) is based on an expanded experimental database covering mafic and ultramafic compositions across the P and T ranges relevant to diamond (1.5–7.5 GPa, 800–1820 °C) and incorporates updated solid solution models for both clinopyroxene and garnet. Older, more simplified, but in some cases still very popular versions of this thermometer do not reproduce the same experiments equally well and often show systematic deviations with changing T and composition. The nominal uncertainty of the Nakamura (2009) thermometer (±74 °C at the 1-sigma level) is still about double that of the best performing thermometers for peridotitic inclusions. None of the available versions of the clinopyroxene–garnet thermometer include corrections for the content of jadeite in clinopyroxene, therefore application to clinopyroxenes with Na > 0.5 apfu is not recommended. This excludes 15% of reported compositions for inclusions in eclogitic diamonds. As regards the Fe3+ problem, Purwin et al. (2013) and Matjuschkin et al. (2014) measured Fe3+/ΣFe ratios of coexisting clinopyroxenes and garnets in a CaO–FeO–Fe2O3–MgO–Al2O3–SiO2 system at 2.5–5.0 GPa and 800–1400 °C and found no systematic deviations in the results of clinopyroxene–garnet Fe–Mg-exchange thermometry using the Krogh (1988) thermometer when all Fe was treated as Fe2+. However, since the system investigated was Na-free, it did not allow for the incorporation of Fe3+ in clinopyroxene as aegirine component. This mechanism was shown to be relevant in clinopyroxenes from garnet peridotites (Woodland 2009) and may be even more so in eclogitic clinopyroxenes. The effect of Na on Fe3+ uptake in clinopyroxene probably contributes to the lower precision of the clinopyroxene–garnet thermometer in applications to natural samples.Trace element-based clinopyroxene–garnet thermometers. The partitioning of REE between clinopyroxene and garnet is sensitive to T and has been used as a base for a clinopyroxene–garnet thermometer (Witt-Eickschen and O’Neill 2005; Sun and Liang 2015; Pickles et al. 2016). A possible advantage of REE thermometry over classical Fe–Mg-exchange thermometry is that REE diffusivity is small compared to that of Fe and Mg. Therefore, REE thermometry is predictably more robust against reequilibration of touching inclusions in diamonds during kimberlite transport or subsequent cooling. The version by Pickles et al. (2016) is particularly interesting, as it is independent of other thermometers and reproduces experimental temperatures moderately well (mean deviation of ±76 °C) with a small P dependence (~50 °C/1 GPa). This method is also independent of variations in Fe3+/ΣFe ratio and may thus pose a promising alternative to the traditional clinopyroxene–garnet Fe–Mg exchange thermometer. A practical disadvantage is the requirement of high-quality REE data, which cannot be obtained with routine electron microprobe analysis. To the author’s knowledge, there are no T data for diamonds using this method.Nitrogen-aggregation thermochronometry. Nitrogen is the most common impurity in natural diamonds. With time, nitrogen atoms in the diamond crystal lattice tend to aggregate, proceeding from initial C-centers (single nitrogen) in Type Ib diamond, to A-centers (nitrogen pairs) in Type IaA diamond and, finally, to B-centers (nitrogen in tetrahedral arrangement) in Type IaB diamond. The first transformation is completed in relatively short geological times (for a N content of 1000 ppm: ~100 yr at 1350 °C, and ~16 m.y at 950 °C; Taylor et al. 1996). Therefore, most natural diamonds are of the intermediate IaAB variety, the proportion of B-centers being a function of nitrogen content, temperature and time (Taylor et al. 1990; Mendelssohn and Milledge 1995). If the ages of the diamond and of its host kimberlite are known, one can calculate the integrated average temperature under which the diamond has resided in the mantle since its formation, i.e., the mantle residence temperature, Tres. Radiometric dating of inclusions within a particular diamond is rarely available. More commonly, the age of the diamond is assumed on the basis of radiometric dating of other diamonds from the same source. However, nitrogen aggregation is much more sensitive to temperature than to time and can be used as an effective thermometer, assuming the diamond has resided in the mantle at constant T since formation (Taylor et al. 1990, 1996). A diagram of nitrogen content vs the relative percentage of B-centers permits appreciation of age uncertainties on thermometry, since isotherms can be drawn for different mantle residence times (Fig. 2). Chronometry errors of one b.y. propagate thermometry errors of only a few tens of degrees on calculated Tres, unless the diamond is unusually young.Nitrogen content and aggregation can be measured by Fourier-Transform Infrared (FTIR) spectroscopy, using an infrared microscope in transmission mode. This method logically provides nitrogen content and aggregation data averaged over the volume of diamond analyzed. Working on thinned (ideally, a few hundred µm) samples at high spatial resolution may help to detect zoning and to obtain more robust estimates from distinct diamond growth zones. The reader is referred to Kohn et al. (2016) for practical guidelines on spatially resolved FTIR thermochronometry and for interpretation of data for zoned diamonds. Detection limits for nitrogen content in diamonds are typically on the order of 5–10 ppm for high-quality spectra. Nitrogen contents below detection limit (Type-II diamonds) are relatively rare in lithospheric diamonds, but are more common in sublithospheric diamonds and are clearly unsuitable for nitrogen-aggregation thermochronometry. It must be clear that, contrary to thermodynamic equilibria used in inclusion thermobarometry, nitrogen aggregation is a kinetic process and so does not provide a snapshot of temperature conditions at a certain time. For instance, a short period of elevated temperature, e.g., due to circulation of hot fluids in the mantle during or after diamond growth, may have a large effect on nitrogen aggregation and could lead to ambiguous interpretations of calculated Tres. In general, unless the diamond is very young relative to its kimberlite host, heating on a time scale of ~1 m.y. will have small effects on the calculated Tres, which will approach ambient mantle conditions rather than conditions during diamond formation (Fig. 3a). If the diamond experienced secular cooling over a b.y. time scale subsequent to formation, the calculated Tres may both significantly underestimate the formation T and overestimate the final ambient T (Fig. 3b).Orthopyroxene–garnet. The Al content of orthopyroxene in equilibrium with garnet is strongly sensitive to P and forms the basis of the most widely used barometer for garnet-bearing peridotites. In this barometer, P mostly depends on the Al content in the orthopyroxene, but the compositions of both minerals must be considered to obtain robust estimates. Of the many calibrations of this barometer, that of Brey and Köhler (1990) is by far the most popular, but the older Nickel and Green (1985) calibration reproduces experimental pressures for most peridotitic systems to somewhat better precision (Nimis and Grütter 2010). Still, the Nickel and Green (1985) version was formulated under simple model-system assumptions and requires correction of Al-component activity to avoid P underestimation for orthopyroxenes with non-negligible contents of jadeite component. Carswell and Gibb (1987) and Carswell (1991) proposed two simple correction schemes, based on different assumptions on how Ti is coupled with Na and Al in orthopyroxene. The two corrections only apply to orthopyroxenes with Na > (Cr + Fe3+ + 2·Ti) or, respectively, Na > (Cr + Fe3+ + Ti) atoms per formula unit, Fe3+ usually being neglected in practical applications. That of Carswell and Gibb (1987), which was also adopted by Brey and Köhler (1990) for their barometer, produces slightly smaller P variations, but there is as yet no solid experimental basis to prefer one correction over the other. The T-dependence of the orthopyroxene–garnet barometer is reasonably small (~0.2–0.3 GPa / 50 °C), but still sufficiently large to require an accurate independent estimate of T. This fact should be considered when this barometer is used in combination with the Harley (1984) orthopyroxene–garnet thermometer, which is known to overestimate at low T and underestimate at high T (see above): the artificial compression of T estimates produced by the Harley (1984) thermometer will necessarily result in non-natural compression also of P estimates. In general, errors in input T will displace the calculated P–T estimates roughly along a typical cratonic geotherm, thus limiting the ‘apparent’ scatter in P–T plots for samples equilibrated on the geotherm (Fig. 4). This may be an advantage if the aim is to define the mantle geotherm, but may give the misperception of a high overall precision by masking the true dispersion of the data.Clinopyroxene–garnet. Barometry of bimineralic eclogites and of eclogitic clinopyroxene–garnet inclusion pairs in diamonds has long been a problem for mantle scientists. There have been several attempts to develop a suitable barometer based on the P-sensitive equilibrium grossular + pyrope ↔ diopside + Ca-Tschermak. According to a recent test by Beyer et al. (2018), the Beyer et al. (2015) formulation reproduces pressures for experiments in natural systems significantly better than the earlier formulations of Simakov and Taylor (2000) and Simakov (2008). Nonetheless, in the Beyer et al. (2018) test on natural xenoliths from the Jericho kimberlite, Slave craton, P estimates for eclogites were systematically lower (by ~1 GPa) than orthopyroxene–garnet P estimates for peridotites recording similar T. As a result, the Jericho eclogites oddly appeared to fall on a hotter geotherm. The T-dependence of the clinopyroxene–garnet barometer is ca. 0.25 GPa / 50 °C, i.e., similar to that of the orthopyroxene–garnet barometer. Therefore, also in this case, errors in input T would displace P–T estimates roughly along the geotherm and cannot explain the observed discrepancy between peridotite-system and eclogite-system thermobarometry results.A further test of clinopyroxene–garnet thermobarometry is illustrated in Figure 5, for touching or non-touching inclusions in diamonds and diamond- or graphite-bearing xenoliths. The P–T estimates reflect iterations of the Beyer et al. (2015) barometer and the Nakamura (2009) Fe–Mg exchange thermometer. Since the uncertainty of the barometer increases rapidly with decreasing concentrations of [4]Al in clinopyroxene, application to clinopyroxenes with Si contents >1.985 atoms per 6-oxygen formula is not recommended by Beyer et al. (2015). These compositions are relatively common for inclusions in diamonds (35% of 224 inclusions) and diamondiferous xenoliths (42% of 233 xenoliths), thus numerous samples had to be excluded. Even if further severe filters were applied to discard potentially low-quality chemical analyses, the results are disconcerting: although graphite-bearing xenoliths plot in the graphite stability field, most diamond-bearing xenoliths and inclusions in diamonds yield too low pressures well outside the diamond stability field (Fig. 5). Iteration of earlier formulations of the same barometer and thermometer does not improve the results (cf. Fig. 10 in Shirey et al. 2013). The reason for this discrepancy is unclear. Simakov and Taylor (2000) cautioned that a barometer based on the Ca-Tschermak content in clinopyroxene may lead to incorrect estimates when applied to kyanite- or SiO2-oversaturated assemblages, but the Beyer et al. (2015) barometer shows no systematic deviations in calculated pressures for experiments in these systems. The large chemical complexity of eclogitic garnets and, especially, eclogitic clinopyroxenes might in part explain the poor performance of the barometer in applications to natural systems, in which variable amounts of Na, Fe3+, Cr, K and vacancy-bearing (i.e., Ca-Eskola) components occur. A further possible explanation is the high sensitivity of the barometer to even small errors in the chemical analyses of clinopyroxene. Beyer et al. (2015) provided a formula to predict relative errors on calculated pressures due to propagation of uncertainties on SiO2 contents, assuming an uncertainty of 1% relative in electron microprobe analysis: %err = 1.94 × 10–8e10.18398[Si apfu]. These model errors do not consider the additional errors that derive from iterative calculation of P and T (Fig. 5). A 1% relative increase in SiO2 contents may move many inaccurate P–T estimates into the diamond field, suggesting that application of the clinopyroxene–garnet barometer may require very high-quality, well-standardized electron microprobe analyses to yield meaningful results. Many of the analyses reported in the literature are possibly of routine quality and do not meet the necessary standards for robust clinopyroxene–garnet barometry. Still, analytical quality alone is unlikely to account for all of the observed discrepancies (Fig. 5).Sun and Liang (2015) proposed that REE partitioning between clinopyroxene and garnet could be used not only as a thermometer (see above) but also as a barometer. Unfortunately, the discrepancies between calculated PREE and run pressures for a set of independent validation experiments appear to be very large (up to ~3 GPa). A test conducted on twelve quartz eclogites, two graphite eclogites and nine diamond eclogites passed the appropriate stability-field constraints for all but two of the nine diamond eclogites; these fell at 0.2 and 0.5 GPa lower P than the diamond-graphite phase boundary. The proposed REE thermobarometry approach is promising, but further tests are necessary before this method can be used with confidence in diamond studies.If trace element data are available, the partitioning of Li between clinopyroxene and garnet can also be used as a barometer (Hanrahan et al. 2009). Pressures for fifteen calibration experiments on a Mg-rich eclogitic composition at 4–13 GPa and 1100–1500 °C are reproduced to a very reasonable ±0.2 GPa (1 sigma) and the T dependence of the barometer is small (~0.2 GPa / 50 °C). Preliminary tests yielded conditions compatible with diamond for three diamond-bearing xenoliths and for non-touching inclusion pairs in four diamonds, whereas two other diamonds yielded conditions up to 2 GPa lower than the diamond stability field. Hanrahan et al. (2009) suggested that the two spurious P estimates could reflect disequilibrium between the non-touching inclusions. However, the use of non-touching inclusions is essential for this barometer, as the high diffusivity of Li may lead to reequilibration of small touching inclusions during and after transport to surface. Similar to the REE barometer, further tests are needed before this method can be recommended for diamond studies.Monomineral barometers.Nimis and Taylor (2000) developed an empirical Cr-inclinopyroxene barometer for chromian diopsides in equilibrium with garnet. In combination with the Nimis and Taylor (2000) single-clinopyroxene thermometer, this barometer uniquely permits derivation of both P and T conditions based on the composition of a single mineral inclusion. The T-dependence of the Cr-in-clinopyroxene barometer (0.15–0.25 GPa / 50 °C for diamond inclusion compositions) is generally lower than that of the orthopyroxene–garnet barometer (~0.2–0.3 GPa / 50 °C) and the P-dependence of the complementary single-clinopyroxene thermometer is very small (Fig. 4). Therefore, if P and T are calculated by iteration, errors in input T or P will displace a clinopyroxene P–T estimate at an angle relative to a cratonic geotherm and will contribute to ‘apparent’ scatter in P–T results for samples equilibrated on a steady-state geotherm. The Nimis and Taylor (2000) Cr-in-clinopyroxene barometer tends to progressively underestimate above 4.5 GPa (by up to ~1 GPa at 7 GPa; Nimis 2002; Ziberna et al. 2016). This is a known artifact and the discrepancy should be considered when comparing Cr-in-clinopyroxene estimates with those obtained using other methods. Nimis et al. (2020) proposed an empirical correction to minimize the differences with P estimates obtained using the orthopyroxene–garnet barometer at high P, at the cost of somewhat lower precision. These authors cautioned that their correction represented a temporary measure to reduce inconsistencies between independent barometric estimates. More recently, Sudholz et al. (2021b) provided an experimental recalibration of the Cr-in-clinopyroxene barometer that yields results very similar to the empirically corrected calibration for clinopyroxenes associated with diamonds or hosted in well equilibrated peridotitic xenoliths at P > 5 GPa (Fig. 6a). At lower P the two revised calibrations diverge slightly, but the Nimis et al. (2020) calibration maintains somewhat better overall consistency with orthopyroxene–garnet barometry (Fig. 6b,c). For proper application of the Cr-in-clinopyroxene barometer, the analyses must undergo compositional filtering in order to select clinopyroxenes in equilibrium with garnet and eliminate certain compositions for which analytical uncertainties propagate excessive uncertainties on the calculated P. Ziberna et al. (2016) provided a useful cookbook to perform the necessary compositional screening, which slightly improved an earlier one proposed by Grütter (2009). According to this cookbook, some compositions require that electron microprobe analyses are carried out using beam currents and counting times higher than routine to minimize analytical uncertainties, whereas some unfavorable compositions should simply be discarded. In particular, the recommended limitation to Cr/(Cr+Al)mol > 0.1 excludes about two thirds of reported inclusions classified as websteritic. When applied to properly filtered graphite- or diamond-bearing xenoliths and inclusions in peridotitic diamonds, the Cr-in-clinopyroxene barometers yield results consistent with or within ~0.3 GPa of the stability field of the respective carbon phase (Nimis and Taylor 2000; see also Fig. 8a). The lack of adequate filtering in previously published thermobarometric results greatly contributed to the larger overall scatter in clinopyroxene P–T plots relative to those for orthopyroxene–garnet pairs (cf. Stachel and Harris 2008).Griffin and Ryan (1995) devised an empirical barometer based on the Cr content of garnet in equilibrium with orthopyroxene and spinel. This Cr-in-garnet barometer was later refined by Ryan et al. (1996) and further revised, using a different Cr/Ca-in-garnet approach, by Grütter et al. (2006). The last two versions are the most widely used, typically in combination with a Ni-in-garnet thermometer. That of Grütter et al. (2006) gives up to ~10% lower estimates and a better agreement with petrological constraints imposed by the graphite–diamond transition, and is therefore preferable. If the coexistence of garnet with spinel is unknown, as is the case for many garnet inclusions in diamonds, the Cr-in-garnet methods only provide an estimate of minimum pressure.Simakov (2008) proposed a simplified version of the clinopyroxene–garnet barometer for eclogites that solely relies on the Ca-Tschermak content in the clinopyroxene. The accuracy of this monomineral barometer cannot be better than that of the native two-phase barometer and is thus prone to the same problematic issues (see above). Ashchepkov et al. (2017) calibrated or recalibrated a series of empirical monomineral thermobarometers, which should be suitable for peridotitic and eclogitic garnets and clinopyroxenes. When applied to mantle xenoliths and diamond inclusions from individual localities, these methods produce strongly scattered P–T estimates (e.g., Figs. 9–12 in Ashchepkov et al. 2017), which are difficult to compare with the more regular P–T distributions that are typically obtained with other, more widely tested methods. The usefulness of the Ashchepkov et al. (2017) monomineral formulations for diamond thermobarometry is therefore questionable.With increasing pressure beyond ~7 GPa, mantle garnets incorporate increasing proportions of majoritic (i.e., pyroxene-like) component by substitution of Si for octahedrally coordinated Al and Cr. The Na and Ti contents also tend to increase. Collerson et al. (2000) utilized the relationship observed between majorite fractions and pressure in experiments on mafic and ultramafic systems to devise an empirical single-mineral barometer. The method was later revised by Collerson et al. (2010), Wijbrans et al. (2016) and Beyer and Frost (2017). Only Wijbrans et al. (2016) proposed two distinct calibrations, one for peridotitic and another for eclogitic systems. Experimental tests by Beyer and Frost (2017) and Beyer et al. (2018) showed systematic deviations or decreased precision for all but the Beyer and Frost (2017) formulation. This was particularly relevant for eclogitic compositions, which were underrepresented in previous calibration datasets. The Beyer and Frost (2017) barometer was calibrated at pressures of 6 to 20 GPa and temperatures of 900 to 2100 °C. These conditions cover sublithospheric mantle conditions up to ~575 km depth, corresponding to the lower transition zone. Experimental pressures are reproduced with a standard error of estimate of 0.86 GPa. In relative terms, considering the enormous range of pressures covered by the calibration, this uncertainty is not much larger than that of the barometers used for lithospheric diamonds. The effect of temperature on pressure estimates is apparently negligible. Note that application of the majorite barometer to lithospheric garnets equilibrated at P < 6 GPa is not recommended, because under such conditions the barometer becomes extremely sensitive to analytical errors (Beyer and Frost 2017). Tao et al. (2018) showed that high Fe3+ contents in the garnet tend to produce underestimated pressures with previous majorite barometers and provided a further revised calibration of the Collerson et al. (2010) barometer that explicitly takes into account Fe3+. The Tao et al. (2018) calibration may be preferable when Fe3+ data are available and measured Fe3+/ΣFe ratios are greater than 0.2. All these majorite barometers require that garnet occurs in equilibrium with a clinopyroxene phase. This may be an issue, because with increasing pressure pyroxene progressively dissolves into garnet and at pressures above ~15–19 GPa, depending on bulk composition and temperature, mantle rocks no longer contain pyroxene (Irifune et al. 1986; Irifune 1987; Okamoto and Maruyama 2004). Pyroxene-undersaturated condition can be suggested by the presence of CaSiO3 phases along with majoritic garnet as inclusions in diamonds or by high Ca contents in majoritic garnet. Under such conditions, the barometer would give erroneously low pressures (Harte and Cayzer 2007). If there is no independent evidence that the garnet was pyroxene-saturated, any pressure estimates close to that of the clinopyroxene-out reaction should be considered as estimates of the minimum P of entrapment. Thomson et al. (2021) devised a novel type of majorite barometer using machine-learning algorithms. Compared to traditional majorite barometers, this new barometer is apparently insensitive to petrological limitations (most notably, the absence of clinopyroxene) and reproduces experimental majoritic garnet compositions over a wider range of bulk compositions and experimental pressures (6–25 GPa) with much better overall precision (mostly within ±2 GPa). Thomson et al. (2021) caution that the compositions of majoritic garnet inclusions in diamonds lie in a region with relatively fewer experiments and machine-learning regressions may not be reliable in extrapolation. Thus, pressure predictions for inclusions in diamonds may have somewhat larger uncertainties. Still, this is the only available barometric method that allows estimating pressures for majoritic garnet inclusions beyond the clinopyroxene-out reaction. When applied to inclusions in diamonds, differences between pressure estimates obtained with the Thomson et al. (2021) and Beyer and Frost (2017) barometers lie in the range –2.5 to +6 GPa. Whatever the majorite barometric method used, if the garnet reequilibrated with clinopyroxene after entrapment, as is typically the case in unmixed inclusions (see below), underestimation of P will occur.Elastic methods. When an inclusion is entrapped in a diamond, both the inclusion and the diamond are initially at the same pressure. On eruption, the pressure acting on the diamond drops to 1 atm, but a residual pressure, Pinc, of up to several GPa may develop on the inclusion as a consequence of the inclusion and host having different elastic properties. If we know the Pinc and the elastic properties of the inclusion and host, we can back-calculate the conditions under which the two minerals would be at the same pressure. These conditions describe a line in P–T space, known as the entrapment isomeke, and the entrapment conditions (i.e., the diamond formation P–T) will be one of the points along the isomeke. This forms the basis for elastic barometry of diamonds. The principles, problems and on-going developments of this method are described in detail in Angel et al. (2022, this volume).Projection onto a geotherm. When only thermometric estimates are available due to a lack of suitable barometers, a tentative estimate of pressure can be obtained by projecting temperature estimates on a reference geotherm. This will most conveniently be a geotherm derived by fitting P–T data for mantle xenoliths or xenocrysts from the same kimberlitic source. The FITPLOT computational software by Mather et al. (2011) provides a useful aid to model geotherms from xenolith P–T data. There may be significant uncertainties in some of the input parameters that are used for the modeling, specifically the thickness and heat production of the upper and lower crust and the mantle potential temperature, but in most cases these uncertainties will not significantly affect the shape of the geotherm within the lithospheric diamond window, which will be chiefly constrained by the xenolith P–T data. Nonetheless, assuming different mantle potential temperatures will affect the estimated thickness of the lithosphere and the depth extent of the conductive portion of the geotherm. This may have some relevance for some old, high-T diamonds, as mantle potential temperatures may have varied over geological time. If this variation is not considered, thermometric estimates may unduly be projected onto the adiabatic portion of the geotherm and lead to severe overestimation of P. Values for the present-day mantle potential temperature and its change with time can be found in Katsura et al. (2010) and Ganne and Feng (2017). The output data that describe the relevant section of the model geotherm can conveniently be fitted through a polynomial expression to derive P as a function of T. The final P–T estimate will be given by the intersection of this polynomial with the isotherm described by the chosen thermometer. Note that this isotherm will not generally be a flat line in a P–T plot, due to the generally significant P-dependence of the thermometers (Fig. 4).The basic assumption underlying this approach is that the inclusions last equilibrated at P–T conditions lying on the mantle geotherm at the time of eruption. This is generally acceptable for touching inclusion pairs, which can reequilibrate in mostly the same way as touching minerals in xenoliths (but see caveat in section below dedicated to touching vs. non-touching inclusions), whereas it is much less obviously so for non-touching inclusions. Comparisons between P–T estimates for inclusions and xenoliths from the same sources showed that the assumption often holds also for non-touching inclusions (Nimis 2002). Nonetheless, there is also evidence of some diamond inclusions recording conditions significantly hotter or colder than the xenolith geotherm (Griffin et al. 1993; Sobolev and Yefimova 1998; Nimis 2002; Stachel et al. 2003, 2004; Weiss et al. 2018; Nimis et al. 2020). Also, multiple non-touching inclusions within individual diamonds may record a range of conditions, indicating thermal fluctuations during the growth history of the diamond (Griffin et al. 1993). Projection onto a geotherm should therefore be used with caution and relatively large uncertainties should be allowed. Assuming a maximum possible T difference from a xenolith geotherm of ~250 °C, consistent with available estimates for diamonds using single-clinopyroxene (Nimis and Taylor 2000), single-garnet (Canil 1999) and orthopyroxene–garnet (Harley 1984) thermometers, projection on a typical cratonic geotherm may result in a P uncertainty of up to ~1.5 GPa (Fig. 4).Projection onto the adiabatic portion of the mantle geotherm may provide a necessary independent constraint for the application of elastic barometric methods to sublithospheric diamonds (e.g., Anzolini et al. 2019).Mineral stability. Lithospheric mantle rocks provide little opportunity to estimate conditions of diamond formation based only on the stability of mineral assemblages. For example, across the entire diamond window, the only significant mineralogical change is the spinel-to-garnet transition in peridotite. The spinel-to-garnet reaction is not univariant and in depleted cratonic peridotites with strongly elevated Cr/Al, spinel + garnet assemblages may persist over a large range of P–T conditions (e.g., MacGregor 1970; Webb and Wood 1986; Klemme 2004; Ziberna et al. 2013). In general, spinel is stabilized to higher pressures in more refractory, Cr-rich and Al-poor bulk compositions. This allows magnesiochromite to be incorporated as inclusions in diamonds that were formed in very deep, refractory environments (e.g., 6.5 GPa, corresponding to a depth of over 200 km, for a diamond from the Udachnaya kimberlite, Siberia; Nestola et al. 2019a).Mineral stabilities become more interesting in the sublithospheric mantle, where a number of characteristic mineralogical changes occur (Harte 2010; Harte and Hudson 2013). Although these changes do not permit bracketing of P–T conditions with a resolution comparable to that of conventional thermobarometers, in some cases they constrain the formation of a diamond to within specific mantle depth regions. For instance, the coexistence of Fe-bearing periclase with enstatite, interpreted to be inverted bridgmanite (previously known as MgSi-perovskite), has long provided evidence that some diamonds were formed at lower-mantle depths (Scott Smith et al. 1984; Moore et al. 1986). Even in the absence of periclase, inverted bridgmanite from the lower mantle can readily be distinguished from upper-mantle enstatite by its very low Ni contents (Stachel et al. 2000). The finding of a rare inclusion of ringwoodite in a diamond from Juina, Brazil, unequivocally demonstrated its origin from the mantle transition zone (Pearson et al. 2014). Also, the reconstructed compositions of a suite of exsolved inclusions in Juina diamonds nicely matched those of minerals expected to form in a basaltic system in the lower mantle (Walter et al. 2011).However, using incomplete mineral assemblages found as inclusions in diamonds as depth markers is not always free from ambiguity. For instance, inclusions of breyite (previously known as CaSi-walstromite) have long been considered to represent retrogressed CaSiO3-perovskite from depths greater than ~600 km (e.g., Harte et al. 1999; Joswig et al. 1999), but there is compelling evidence that at least some breyite inclusions originated at much shallower depths within the upper mantle (Brenker et al. 2005; Anzolini et al. 2016). Woodland et al. (2020) recently provided experimental support for a possible upper-mantle origin of breyite. Also, Fe-bearing periclase is the most common inclusion of interpreted lower-mantle origin (Kaminsky 2012), but there is evidence that Fe-bearing periclase participates in mineral parageneses straddling the upper mantle-lower mantle boundary (Hutchison et al. 2001) and that it may be a co-product of diamond-forming reactions that may occur in the transition zone and even in the overlying upper mantle (Thomson et al. 2016; Bulatov et al. 2019).Data quality. A sometimes underrated prerequisite for robust thermobarometry is precise and accurate chemical analyses. Inclusions in diamonds are often small and their chemical analysis can be challenging. Moreover, some thermobarometers may be particularly sensitive to analytical uncertainties on elements that occur at critical concentration levels. For instance, Ziberna et al. (2016) explored propagation of analytical errors on P estimates using the single-clinopyroxene barometer of Nimis and Taylor (2000) and found that routine electron microprobe analytical conditions may be insufficient for meaningful barometry of many inclusions in diamonds. Clearly, not only poor counting statistics but also bad sample preparation and improper standardization can be an issue and may lead to spurious results. A survey of published chemical data for over 600 clinopyroxene inclusions in lithospheric diamonds reveals that ~30% of them have oxide total weight percentages < 98.5% or > 101%, or cation sums < 3.98 or > 4.03 atoms per formula unit and, thus, are definitely not of good quality! Note that the above cut-offs generously take into account the possible effects of considering all Fe as Fe2+ in electron microprobe analyses.The next critical step when studying inclusions that were separated from coexisting mineral phases is the definition of their original paragenesis. Even if single-mineral thermobarometers may often come in handy, they invariably assume that the investigated mineral last equilibrated with specific other minerals. As for clinopyroxene and garnet, which are major constituents in several types of mantle rocks, simple compositional screening based on either major or trace element concentrations may help to discriminate between the most typical paragenetic varieties (e.g., Griffin et al. 2002; Grütter et al. 2004; Nimis et al. 2009; Ziberna et al. 2016). Unusual compositions that are not sensitively discriminated by available classification schemes or fall outside the compositional range over which the thermobarometers were calibrated, as well as false positives that unduly survive compositional filtering, may undermine the reliability of thermobarometric data. This problem may be minimized by investigating large datasets, when this is permitted by the number and nature of the available samples.Even when all cautionary steps have been made to select suitable compositional data, thermobarometric estimates are subject to errors. As discussed in the previous section, some of these errors are systematic and lead to predictable inconsistencies between estimates obtained using different methods. This fact should always be considered when comparing data obtained using different thermobarometers or even different formulations of the same thermobarometer. Thermobarometric methods with demonstrated internal consistency must be prioritized if the purpose is to explore the distribution of data for heterogeneous inclusion populations.In this respect, thermobarometry of eclogitic diamonds remains problematic. Considering the compositional limitations required by the ‘best’ available barometric methods (i.e., SiCpx < 1.985 apfu and NaCpx < 0.5 apfu), ~45% of over 200 reported eclogitic clinopyroxene–garnet inclusion pairs are automatically excluded. In addition, the reliability of barometric methods for eclogitic inclusions is questionable (see above).When P estimates are too uncertain or not available, for example, due to a lack of suitable barometers for the specific inclusions investigated, projection of temperatures onto the local xenolith geotherm can provide a last-resort solution to estimate P. The resulting uncertainties can be too large for an accurate characterization of genesis conditions for individual diamonds, but the procedure can have some utility for exploring general trends in large datasets (e.g., Korolev et al. 2018; Nimis et al. 2020). It probably remains the safest available method for barometry of eclogitic inclusions, notwithstanding potential over- or underestimation for diamonds formed at conditions hotter or, respectively, colder than ambient mantle at the time of eruption. A fairly popular alternate has been to calculate T for these diamonds at a fixed P of 5 GPa, approximately corresponding to the average P for diamonds worldwide. This method is not as effective and may overcompress the range of T estimates, due to the generally significant P-dependence of the thermometers (Fig. 4).Touching vs. non-touching. When a pair of touching mineral grains included in a diamond reequilibrate to changing external conditions, their total encapsulated volume remains constrained by the surrounding diamond. This forces the included minerals to follow a distinct pressure path and to achieve final compositions that are different from those of the same minerals outside the diamond (see Angel et al. 2015 and Ferrero and Angel 2018 for more general discussion of this phenomenon). The most common changes in ambient conditions are probably isobaric cooling or heating after diamond growth, in response to secular variations in mantle thermal state or following short-term thermal pulses related to diamond-forming processes. In most cases of diamond-hosted inclusions, cooling will cause a decrease in the inclusion P even if the external P remains constant, leading to underestimation of the actual diamond P using chemical thermobarometers. The extent of this underestimation can be calculated if the thermoelastic properties of the inclusion and host minerals are known (Angel et al. 2015). For various combinations of garnet, pyroxene and olivine inclusions in diamond, the error would amount to ~0.3 GPa for a temperature drop of 100 °C. Heating will have the opposite effect. These errors are of a similar magnitude as those of the orthopyroxene–garnet and single-clinopyroxene barometers and may thus contribute to some of the scatter in P–T plots for touching inclusions. The resulting errors may be lower, however, if plastic deformation in the diamond is able to reduce the under- or overpressures developed on the inclusions.If some thermal reequilibration occurs, the use of single-mineral thermobarometers on touching inclusions may generate further errors when some key minerals are not part of the inclusion assemblage. For instance, touching inclusions of clinopyroxene-orthopyroxene devoid of garnet will reliably record the ambient thermal state at the time of eruption, but will be poor candidates for single-clinopyroxene barometry. In response to a thermal change, the clinopyroxene will easily reequilibrate with orthopyroxene, but not with garnet, and so a false apparent encapsulation P will be calculated at the final equilibrium T (Nimis 2002). Touching inclusions of clinopyroxene ± garnet ± olivine would instead produce estimates close to the diamond formation conditions, because the composition of the clinopyroxene in orthopyroxene-free assemblages would be little affected by a variation in temperature. Polymineralic inclusions consisting of touching clinopyroxene-orthopyroxene–garnet ± olivine will fully reequilibrate and will provide P and T conditions close to those at the time of eruption (except for the small systematic errors described in the previous paragraph).Non-touching inclusions are immune to the above problems, because the inert diamond protects them from post-entrapment reequilibration, and are unequivocally the best candidates to determine the diamond formation P–T, provided reliable monomineral thermobarometers can be used. In all other cases, non-touching inclusion pairs are prone to potential errors, because the individual inclusions may have been incorporated at different times and under different conditions (see above) and may thus not be in mutual equilibrium. The largest T difference between multiple inclusions in individual diamonds was reported for garnets in a diamond from the Mir kimberlite, Siberia (Griffin et al. 1993). The calculated difference varies greatly depending on which calibration of the Ni-in-garnet thermometer is used, i.e., ca. 400 °C with the Ryan et al. (1996) calibration or ca. 230 °C with the Canil (1999) calibration. Disequilibrium amongst non-touching inclusions in diamonds, however, appears to be the exception rather than the rule, as multiple inclusions in individual diamonds most often yield similar P–T estimates once the systematic deviations between different thermobarometers are taken into account (Stachel and Harris 2008; Stachel and Luth 2015).Syngenesis vs. protogenesis. For several decades, thermobarometry of inclusions in diamonds has relied on the assumption of syngenesis, i.e., the inclusions formed or completely recrystallized during the growth of diamond. A corollary of this assumption was that the inclusions record the conditions of diamond formation, excepting for potential post-entrapment modifications in touching inclusions. The main proof for syngenesis was that diamond typically imposed its shape on the inclusions. This widely accepted paradigm was challenged by Nestola et al. (2014), who found evidence of protogenesis (i.e., formation before the diamond) for some olivine inclusions that would have been classified as syngenetic based on morphological criteria. Similar evidence was then found for other olivine, clinopyroxene, garnet and magnesiochromite inclusions (Milani et al. 2016; Nestola et al. 2017, 2019b; Nimis et al. 2019). The possibility that inclusions in diamonds are protogenetic suggests that they might record conditions predating diamond formation. This might be particularly relevant for isolated inclusions, if their host diamonds were formed from hot fluids or melts intruding through relatively cool mantle rocks. Did the preexisting mineral grains have enough time to fully reequilibrate to the changing thermal conditions during diamond growth and before their incorporation was complete?The time required by a mineral to adjust its composition to new physicochemical conditions strongly depends on temperature and grain size, and can be estimated from diffusion kinetics. For instance, a clinopyroxene grain with a diameter of 200 µm would fully reset its Ca/Mg ratio (the main temperature sensor in pyroxene thermometry) in less than 100,000 yr above 1200 °C and in a few m.y. below 900 °C (Fig. 7). Whether these time spans are short or long enough for complete reequilibration depends on the diamond growth rate, which is not precisely known. It is tempting to infer that incomplete reequilibration of protogenetic inclusions, especially the largest and shallowest (i.e., coldest) ones, may account for some of the scatter in P–T plots for non-touching inclusions (Fig. 8). Also, the similarity of P–T conditions occasionally observed for inclusions in diamonds and xenoliths from the same sources may alternatively be ascribed to diamond formation from thermally-equilibrated media (Nimis 2002) or to lack of reequilibration of protogenetic inclusions during short-lived diamond-forming processes (Nestola et al. 2014).Exsolved inclusions. Some inclusions in diamonds show evidence of post-entrapment unmixing. Reported examples are clinopyroxenes with orthopyroxene exsolution lamellae (± coesite), representing unmixed high-T clinopyroxenes, and intergrowths of more or less majoritic garnet and clinopyroxene, representing unmixed high-P majoritic garnets.Exsolved clinopyroxene inclusions are uncommon and have only been reported in diamonds from Namibian placers and from the Voorspoed kimberlite, South Africa (Leost et al. 2003; Viljoen et al. 2018). The coexistence of clinopyroxene and orthopyroxene makes these inclusions ideal for pyroxene thermometry, but the exsolved assemblage does not provide barometric estimates. Since pyroxene thermometers have a small P-dependence, projection onto the local xenolith geotherm provides in these cases reasonable estimates of the depth of origin of the diamonds (Nimis et al. 2020). If the compositions and relative proportions of the exsolved phases are known (e.g., by combining electron microprobe data with 2D or, better, 3D image analyses), the integrated pre-exsolution composition of the clinopyroxene can be reconstructed. The integrated composition can provide a T estimate at the time of diamond growth, assuming equilibrium with a separate orthopyroxene during entrapment, or a minimum T estimate of diamond formation, in the absence of orthopyroxene. In the reported cases, thermometry of the reconstructed clinopyroxenes provided evidence for diamond formation at temperatures near the mantle adiabat and ~200 °C hotter than the ambient mantle at the time of eruption. Attempts to apply the single-clinopyroxene barometer to the reconstructed compositions yielded unreasonably low pressure estimates, suggesting formation in strongly modified mantle environments (Nimis et al. 2020).Exsolution textures are common in majoritic garnet inclusions in sublithospheric diamonds (Moore and Gurney 1989; Wilding 1990; Harte and Cayzer 2007). These textures are interpreted to reflect decompression during slow ascent of the diamonds in convecting mantle or mantle plumes. The presence of exsolutions implies that pressure estimates based on majoritic garnet compositions are minima. Using reconstructed pre-exsolution garnet compositions will yield conditions closer to those of diamond formation. Even so, the resulting estimates may be minima, because the garnet may have originated at depths beyond the clinopyroxene-out reaction (see previous Barometry section).Figure 8 shows a compilation of P–T data for lithospheric diamonds from worldwide sources, obtained using the most robust available thermobarometer combinations. The compilation comprises clinopyroxene inclusions in 100 lherzolitic diamonds and orthopyroxene–garnet inclusion pairs in 77 harzburgitic, 12 lherzolitic and 10 websteritic diamonds. The great majority of the clinopyroxene inclusions were isolated within their host diamonds and should generally record conditions close to those of diamond formation. The orthopyroxene–garnet pairs consist of both touching and non-touching assemblages. The touching pairs may have reequilibrated after their incorporation and should provide conditions close to the final conditions of residence in the mantle. In addition, P–T data were calculated, using the same methods, for diamondiferous xenoliths for which suitable mineral analyses were available.In comparing the two plots for single-clinopyroxene and orthopyroxene–garnet thermobarometry (Fig. 8a,b), it is important to consider the known systematic inconsistencies between the thermobarometric methods used (see above). The orthopyroxene–garnet thermometer tends to overestimate temperature below 1100 °C and to underestimate above 1100 °C. This accounts at least in part for the more restricted temperature range obtained for the orthopyroxene–garnet inclusions. Nonetheless, disregarding a single obvious outlier, all P–T values fall within 0.3 GPa of the diamond stability field, confirming the general reliability of the thermobarometers. Most of the P–T results lie between the 35- and 40-mW/m2 conductive geotherms of Hasterok and Chapman (2011), i.e., at conditions typical for cratonic lithospheres. Most exceptions record higher temperature conditions, which in the case of clinopyroxene may reach the mantle adiabat. All of these exceptions consist of non-touching inclusions. A possible explanation is that several diamonds were formed from thermally non-equilibrated media derived from the sublithospheric convective mantle. Another potential source of mantle temperature inhomogeneity during diamond formation can be incomplete cratonization at the time of diamond growth. For instance, in the Kaapvaal craton a stable geothermal gradient was established by about 2.5 Ga and followed by secular cooling (e.g., Michaut and Jaupart 2007; Brey and Shu 2018) and re-heating (Griffin et al. 2003), but many Kaapvaal diamonds are significantly older than 2.5 Ga (see review in Shirey and Richardson 2011) or may have formed during important perturbations of the local lithosphere (Griffin et al. 2003; Korolev et al. 2018). Therefore, it is not surprising that many of these diamonds may record conditions hotter or colder than those at the time of kimberlite eruption (e.g., Nimis et al. 2020).The touching orthopyroxene–garnet inclusion pairs deserve further discussion. In the available dataset, the majority of them (23 out of 31) refer to diamonds from the De Beers Pool (Kimberley, South Africa). These inclusions were studied by Phillips et al. (2004), who carefully documented differences between P–T estimates recorded by touching vs. non-touching inclusions. These workers found that the non-touching inclusions yielded higher P–T values than the touching inclusions and ascribed this observation to post-entrapment reequilibration of the touching inclusions under decreasing T conditions, possibly accompanied by mantle uplift and decompression. They also correctly noted that post-entrapment elastic modifications were not of sufficient magnitude to reconcile the observed P–T differences between touching and non-touching inclusions. However, Nimis et al. (2020) observed that both touching and non-touching inclusions in the Phillips et al. (2004) dataset record a colder geothermal regime than mantle xenoliths and xenocrysts in the same kimberlites, casting doubts on the secular cooling hypothesis. Whereas the relatively cold signature of the non-touching inclusions might reflect old conditions pre-dating the final thermal equilibration of the lithosphere recorded by the xenoliths, the data for the touching inclusions are difficult to interpret. Weiss et al. (2018) speculated that the low temperatures recorded by the touching inclusions could be related to late infiltration of cold, slab-derived fluids, remnants of which were preserved in some cloudy and coated diamonds from the same locality. Yet it remains unclear why infiltration of these fluids should leave virtually no traces in the sampled mantle xenoliths.Touching orthopyroxene–garnet inclusions from other localities are probably too few to be statistically significant. Nonetheless, in all these cases the touching inclusions yield temperatures significantly lower than non-touching inclusions in other diamonds from about the same depth (Fig. 8b). Additional thermobarometric evidence that diamond formation may be followed by cooling was provided by Stachel and Luth (2015) and Viljoen et al. (2018). However, Stachel and Luth (2015) also described touching inclusions yielding temperatures similar to or even slightly higher (up to 40 °C) than non-touching inclusions in the same diamonds.Diamond’s depth systematics. Irrespective of the nature of the inclusions and the thermobarometric methods used, the frequency distribution of pressures for lithospheric diamond is essentially unimodal, with a mode at ~6 GPa, corresponding to depths around ~190 km (Fig. 8; see also Stachel and Harris 2008 and Stachel 2014 for broadly similar distributions obtained using different thermobarometers and partly different datasets). The slight differences between the pressure distributions obtained for clinopyroxene and orthopyroxene–garnet inclusions may largely be ascribed to recognized inconsistencies between the different thermobarometer combinations used and, perhaps, to sampling bias. Therefore, these differences are probably not significant. The hard constraints imposed by the graphite–diamond and lithosphere-asthenosphere boundaries cannot fully explain the observed depth distributions, which show decreased diamond frequencies approaching these boundaries (Fig. 8). Nimis et al. (2020) obtained very similar depth distributions with depth modes at 180 ± 10 km for diamonds from three individual South African kimberlite sources by using various combinations of thermobarometric methods on different inclusion types. They also found no clear correlation with the depth distributions of mantle xenocrysts from the same sources calculated using the same thermobarometric methods. They concluded that the observed diamond depth distributions are unrelated to the kimberlite sampling efficiency and, thus, have genetic meaning and likely global significance.The global significance of the observed depth distributions is also supported by the abundant TNi-in-garnet data for garnet inclusions (Fig. 9). These data show a mode at ~1200 °C, but no pressure constraints are generally available. Assuming that the garnets last equilibrated at ‘average’ conditions between the 35- and 40- mW/m2 model geotherms, similar to the clinopyroxene and orthopyroxene–garnet inclusions (Fig. 8), the Ni-in-garnet temperature mode would correspond, again, to a pressure mode of ~6 GPa and a depth mode of ~190 km. The gentler increase of frequencies to ~1200 °C and their steeper decrease at higher temperatures (Fig. 9) also mirror the frequency variations with increasing pressure to ~6 GPa and beyond exhibited by the clinopyroxene and orthopyroxene–garnet inclusions (Fig. 8).The diamond depth distributions in the lithosphere may be the result of both constructive and destructive processes (Nimis et al. 2020). In particular, the progressive decrease in diamond concentrations at depths shallower than ~170–190 km may reflect (i) decreasing precipitation rates for diamond from ascending parental media or (ii) progressive decrease in the amount of these deep-sourced media at shallower levels. Probably, both mechanisms concur to limit the diamond endowment in the shallow portion of the diamond window. The rapid decrease of diamonds near the base of the lithosphere may reflect (i) a positive balance between the opposed effects that an increase of P and an increase of T may have on carbon solubility in some mantle fluids/melts along specific P–T paths or (ii) a progressive decrease in diamond endowment due to reactions with carbon-undersaturated asthenospheric fluids/melts. These two alternative hypotheses are difficult to prove or disprove, because our knowledge of the behavior of natural carbon-bearing melts and fluids at high P–T is still limited. The prevalence of melt-metasomatized lithologies near the base of typical lithospheres coincides with diminished diamond frequency and suggests that melt-driven metasomatism may be an important destructive agent for diamond (e.g., McCammon et al. 2001). Remobilization of carbon from these deep lithospheric roots may eventually contribute to build up the diamond concentration at shallower depths (Nimis et al. 2020).The recognition that depths around ~190 km are particularly favorable for diamonds may have interesting implications for their mineralogical exploration. Areas characterized by the occurrence of xenocrystic mantle minerals (e.g., clinopyroxenes or garnets) from these particular depths may be considered as high-priority targets. Nonetheless, since the xenocrysts and diamond depth distributions may be uncorrelated, even limited xenocryst records from the diamond’s most favorable depth range may be indicative of significant diamond potential (Nimis et al. 2020).Figure 10 shows the global distribution of pressure estimates for diamonds containing majoritic garnets. At present, these are the only types of inclusions that provide abundant, quantitative barometric data for sublithospheric diamonds. A difficulty in investigating these data lies in the fact that they may represent minimum P estimates. As explained in the previous section, this limitation is due to the frequent occurrence of exsolutions and, for all but the Thomson et al. (2021) barometer, to the possible absence of accompanying clinopyroxene in the original mineral assemblage from which the majoritic garnet inclusion was separated on entrapment. If one considers only inclusions that are reportedly free of clinopyroxene exsolutions, two distinct unimodal distributions are obtained for peridotitic and eclogitic/pyroxenitic garnets, respectively (Fig. 10). Using the Beyer and Frost (2017) barometer, the peridotitic inclusions are mostly clustered in the 175–225 and 225–275 km depth intervals, near the lithosphere-asthenosphere boundary. Most of these inclusions would thus more appropriately be classified as deep lithospheric. The eclogitic/pyroxenitic inclusions are, on average, much deeper and mostly fall in the depth interval between 300 and 500 km. If one considers also inclusions that do contain exsolved pyroxenes, the peak of eclogitic/pyroxenitic inclusions at ~400 km becomes more prominent. Using the Thomson et al. (2021) barometer, the distribution for the eclogitic/pyroxenitic inclusions is stretched to slightly higher pressures, with a maximum estimated depth of 613 ± 44 km, whereas many of the peridotitic inclusions are shifted to significantly greater depths, falling well into the sublithospheric region (Fig. 10). Some of the deepest (>450 km) majoritic garnets fall in the region where clinopyroxene disappears and thus the Beyer and Frost (2017) barometer may underestimate pressure for these samples. However, the Thomson et al. (2021) machine-learning barometer, which is immune to this effect, does not suggest a much deeper origin for these garnets (Fig. 10). Differences between depth estimates for these very deep samples using the two alternative barometers are between –65 and +81 km (average +17 km).It should be considered that, in some cases, in the absence of accurate imaging data, the occurrence of fine clinopyroxene exsolutions may have been overlooked. The use of improved BSE and EBSD imaging might allow recognition of exsolved clinopyroxene in nominally ‘clinopyroxene-free’ majoritic inclusions (cf. Harte and Cayzer 2007; Zedgenizov et al. 2014). The thirteen Juina majoritic inclusions for which no coexisting clinopyroxene has been reported to date give depth estimates in a restricted range between 353 and 455 km (using the Beyer and Frost 2017 barometer) or 368 to 462 km (using the Thomson et al. 2021 barometer) and are in part responsible for the prominent peak at ~400 km in Figure 10. If these inclusions also contained exsolutions and the available electron microprobe data do not faithfully reflect their original compositions, then the overall peak could be shifted to somewhat greater depths or, at least, the overall depth distribution would be smoother than shown in Figure 10. Thomson et al. (2021) suggested that all majoritic garnet inclusions in South American diamonds probably contain exsolutions. They also noted that in the eight inclusions for which both exsolved and reconstructed bulk compositions were available the amount of exsolution is a simple function of the calculated pressure. Accordingly, they devised an empirical correction that can be applied to all inclusions for which reconstructed bulk compositions are not available. When this correction is applied to South American diamonds, the peak for eclogitic majoritic garnet inclusions is smoothed and shifted to ca. 100 km greater depth. It remains unclear if this correction, which is based on a restricted number of inclusions, has any general validity at either global or local scale. In any case, the high abundance of depth estimates between ~300 and ~500 km, even for inclusions that are certainly clinopyroxene-free or whose original composition was reconstructed, suggests that most of the eclogitic/pyroxenitic majoritic garnets indeed originated from this depth range (Harte 2010), or perhaps from a slightly more extended range of ~300 to ~600 km if the Thomson et al. (2021) correction for South American diamonds is valid. This range overlaps with independent depth estimates obtained from phase stability constraints for other sublithospheric diamonds containing Ca-silicate inclusions (300–360 km; Brenker et al. 2005; Anzolini et al. 2016) and with a minimum depth estimate obtained from elastic barometry for a single ferropericlase inclusion (~450 km; Anzolini et al. 2019). Of further interest, the 300–600 km range corresponds to the depths at which carbonated MORB slabs are most likely to intersect their solidus, providing an ideal environment for focused melting and potential diamond growth (Thomson et al. 2016, 2021). Considering that many periclase-magnesiowüstite inclusions, for which barometric data are not available, may also have formed from subduction-related melts (Thomson et al. 2016; Nimis et al. 2018) the actual proportion of diamonds from the 300–600 km interval may be even larger than shown in Figure 10. The reader is referred to Walter et al. (2022, this volume) for a thorough discussion on these and related aspects.A second favorable region for diamond formation is probably located near the upper mantle–lower mantle boundary and uppermost lower mantle (Harte 2010). Evidence for diamond formation in this region is provided by a number of key mineral associations within individual diamonds, mostly consisting of various combinations of periclase, (Mg,Fe)SiO3, (Mg,Fe)2SiO4, jeffbenite (formerly known as TAPP), breyite, corundum and SiO2, and of a NaAl-pyroxene phase (believed to have formed as highly majoritic garnet) alongside with periclase or jeffbenite (Hutchison et al. 2001; Harte 2010). Although in several diamonds only part of the key mineral assemblages is actually represented, the general consistency between the observed mineral associations led Harte (2010) to suppose that a large number of diamonds could come from the relatively narrow depth range of 600 to 800 km. Most inclusions in diamonds from this depth range are ultramafic, as opposed to the mafic inclusions that dominate in the 300–600 km interval.Mineral inclusions are very rare in diamonds and occur in only about 1% of examined stones, with significant differences between individual localities (Stachel and Harris 2008). Moreover, only a few mantle minerals or mineral assemblages that may occur as inclusions in diamonds are suitable for thermobarometry. Therefore, only a small proportion of natural diamonds may yield robust information on the P–T conditions of their formation, and these diamonds are mostly peridotitic. This proportion may increase in the future, in parallel with the development of new thermobarometric methods or refinement of existing ones. A significant field of potential and desirable improvement concerns the eclogitic inclusions, which represent a major fraction of inclusions in lithospheric diamonds, but for which existing barometric methods still provide inconsistent results. A dedicated test on clinopyroxene–garnet pairs similar to that performed on chromian diopsides by Ziberna et al. (2016) might shed more light on the minimum analytical quality that is required for reliable barometry of these inclusions. Extending the applicability of non-traditional methods, such as elastic barometry, to non-ideal configurations and to a more diverse list of inclusion minerals might also have a dramatic effect on the number of diamonds suitable for thermobarometry, although the need for elastically preserved inclusions generally restricts applications to relatively small inclusions not surrounded by fractures. An interesting advantage of the elastic method is that it can rely on non-destructive techniques and, thus, can leave the samples intact and ready for further investigation with other methods. Such investigations require the use of undamaged samples in which internal strains are well preserved. Legacy destructive techniques used to expose minerals included in diamond, unfortunately, prohibit future examination of these unique samples for purposes of elastic barometry.Reanalysis of some previously investigated inclusions or diamondiferous xenoliths might provide a relatively effortless way to increase the number of diamond P–T data. Four of 195 published chromian clinopyroxene analyses and 6 of 156 published orthopyroxene analyses had to be discarded while preparing this review only because of sub-standard analytical quality. Moreover, twenty-five clinopyroxenes were rejected because their compositions fell just off the major-element compositional fields that are used to classify garnet-associated chromian diopsides (Ziberna et al. 2016). Trace element analysis of these diopsides might have allowed more robust discrimination (e.g., based on Sc/V relationships; Nimis et al. 2009) of the garnet-associated samples and many of them might have turned out to be suitable for thermobarometry. Reintegrating all these samples would increase the number of P–T data for chromian diopside-bearing diamonds by a remarkable 27% and for diamonds overall by 17%.An enlargement of the number of P–T data for diamonds would have significant positive outcomes in terms of sample bias reduction. Sample bias is presently a major problem, since the number of localities for which a statistically significant number of sufficiently reliable P–T estimates are currently available is limited at best to three and all are from the Kaapvaal craton (Nimis et al. 2020). Consequent benefits of sample bias reduction would be a better assessment of diamond depth distributions at individual localities, and a more robust evaluation of diamond P–T systematics and of their significance in terms of diamond forming processes and preservation.If the number of P–T data for diamonds is important, their internal consistency may be even more so. In fact, most of the combinations of thermobarometric methods that may be applied to diamond inclusions do show some degree of inconsistency with one another. These inconsistencies mostly stem from simplified thermodynamic treatment of solid solutions in minerals and from uncertainties in the data used for their calibration and derived either from experiments or from independent xenolith thermobarometry. The internal consistency of thermobarometric methods is particularly critical when comparing data for different inclusion types, as commonly is necessary in diamond studies. This is another field of potential scientific advance that may eventually contribute to an improved understanding of diamond forming processes.This review is the result of the knowledge, expertise and generosity of many people. Fruitful collaboration through the years with many of them, and particularly with W.R. Taylor, H. Grütter, L. Ziberna, F. Nestola and R.J. Angel, has been of decisive importance in producing my personal contributions to diamond thermobarometry. Yet many more people have in fact contributed to this review through their enormous scientific work and by providing the community with invaluable data over about five decades. Most of them appear in the reference list below. I express my sincere gratefulness to all of these people and, in particular, to T. Stachel, who generously shared a diamond inclusion database, which served as a useful basis to build the updated database used in this work. I am also grateful to A. Thomson, for help in majorite barometry calculations, and to H. Grütter, G. Brey and editor T. Stachel for their very helpful and thorough reviews and comments.\",\"PeriodicalId\":501196,\"journal\":{\"name\":\"Reviews in Mineralogy and Geochemistry\",\"volume\":\"63 8\",\"pages\":\"\"},\"PeriodicalIF\":0.0000,\"publicationDate\":\"2022-07-01\",\"publicationTypes\":\"Journal Article\",\"fieldsOfStudy\":null,\"isOpenAccess\":false,\"openAccessPdf\":\"\",\"citationCount\":\"0\",\"resultStr\":null,\"platform\":\"Semanticscholar\",\"paperid\":null,\"PeriodicalName\":\"Reviews in Mineralogy and Geochemistry\",\"FirstCategoryId\":\"1085\",\"ListUrlMain\":\"https://doi.org/10.2138/rmg.2022.88.10\",\"RegionNum\":0,\"RegionCategory\":null,\"ArticlePicture\":[],\"TitleCN\":null,\"AbstractTextCN\":null,\"PMCID\":null,\"EPubDate\":\"\",\"PubModel\":\"\",\"JCR\":\"\",\"JCRName\":\"\",\"Score\":null,\"Total\":0}","platform":"Semanticscholar","paperid":null,"PeriodicalName":"Reviews in Mineralogy and Geochemistry","FirstCategoryId":"1085","ListUrlMain":"https://doi.org/10.2138/rmg.2022.88.10","RegionNum":0,"RegionCategory":null,"ArticlePicture":[],"TitleCN":null,"AbstractTextCN":null,"PMCID":null,"EPubDate":"","PubModel":"","JCR":"","JCRName":"","Score":null,"Total":0}
引用次数: 0

摘要

这构成了钻石弹性气压计的基础。这种方法的原理、问题和持续发展在Angel等人(2022年,本卷)中有详细描述。投影到地热上。当由于缺乏合适的气压计而只能获得温度估计时,可以通过在参考地热上投射温度估计来获得压力的暂定估计。这将最方便地通过拟合来自同一金伯利岩源的地幔捕虏体或异晶的P-T数据得到地热。Mather等人(2011)的FITPLOT计算软件为从捕虏体P-T数据中模拟地热提供了有用的帮助。一些用于建模的输入参数可能存在显著的不确定性,特别是上下地壳的厚度和产热以及地幔的位温,但在大多数情况下,这些不确定性不会显著影响岩石圈钻石窗内的地热形状,这将主要受到捕虏体P-T数据的限制。然而,假设不同的地幔位温会影响岩石圈厚度的估计和地热传导部分的深度范围。这可能与一些古老的高t钻石有关,因为地幔的潜在温度可能随地质时间而变化。如果不考虑这种变化,测温估计可能会不适当地投影到地热的绝热部分,并导致对现今地幔势温及其随时间变化的p值的严重高估,可以在Katsura et al.(2010)和Ganne and Feng(2017)中找到。描述模型地热相关部分的输出数据可以方便地通过多项式表达式进行拟合,从而推导出P作为t的函数。最终的P - t估计值将由该多项式与所选温度计所描述的等温线相交得到。注意,这条等温线在P-T图中通常不会是一条平坦的线,因为温度计通常具有显著的p依赖性(图4)。这种方法的基本假设是,包裹体在喷发时位于地幔地热的P-T条件下最后达到平衡。对于接触包裹体对,这通常是可以接受的,它可以以与接触捕虏体中的矿物相同的方式重新平衡(但请参阅下面专门讨论接触与非接触包裹体的注意事项),而对于非接触包裹体,这种情况就不那么明显了。对来自同一来源的包裹体和捕虏体的P-T估计的比较表明,该假设通常也适用于非接触包裹体(Nimis 2002)。尽管如此,也有证据表明一些钻石包裹体记录的条件比捕虏体地热明显更热或更冷(Griffin et al. 1993;Sobolev and Yefimova 1998;Nimis 2002;Stachel et al. 2003, 2004;Weiss et al. 2018;Nimis et al. 2020)。此外,单个钻石内的多个非接触内含物可能记录了一系列条件,表明钻石生长历史中的热波动(Griffin et al. 1993)。因此,对地热的投影应谨慎使用,并应允许相对较大的不确定性。假设捕虏岩地热的最大可能温差为~250°C,这与使用单斜辉石(Nimis and Taylor 2000)、单石榴石(Canil 1999)和正辉石-石榴石(Harley 1984)温度计对钻石的现有估计一致。对典型克拉通地热的投影可能导致P不确定性高达~1.5 GPa(图4)。对地幔地热绝热部分的投影可能为弹性气压测量方法在岩石圈下钻石的应用提供必要的独立约束(例如,Anzolini et al. 2019)。矿物的稳定。岩石圈地幔仅根据矿物组合的稳定性来估计金刚石形成条件的机会很少。例如,在整个钻石窗口中,唯一显著的矿物学变化是橄榄岩中尖晶石向石榴石的转变。尖晶石-石榴石的反应不是一成不变的,在Cr/Al强烈升高的枯竭克拉通橄榄岩中,尖晶石+石榴石组合可能在很大范围的P-T条件下持续存在(例如,MacGregor 1970;Webb and Wood 1986;Klemme 2004;Ziberna et al. 2013)。一般来说,尖晶石在更难熔、富cr和贫al的大块成分中稳定在更高的压力下。这使得镁铬铁矿可以作为包裹体掺入在非常深的难熔环境中形成的钻石中(例如,6.5 GPa,对应深度超过200公里,来自西伯利亚Udachnaya金伯利岩的钻石;Nestola et al. 2019a)。 矿物稳定性在岩石圈下地幔中变得更加有趣,在那里发生了许多特征矿物学变化(Harte 2010;Harte and Hudson 2013)。尽管这些变化不允许将P-T条件与传统的温压计的分辨率相媲美,但在某些情况下,它们将钻石的形成限制在特定的地幔深度区域内。例如,含铁方长石与顽辉石共存,被解释为倒桥辉石(以前称为mgsi -钙钛矿),长期以来为一些钻石形成于下地幔深处提供了证据(Scott Smith et al. 1984;Moore et al. 1986)。即使在没有方长石的情况下,下地幔的倒菱辉石也可以很容易地与上地幔的顽辉石区分开来,因为它们的镍含量非常低(Stachel et al. 2000)。在巴西Juina的一颗钻石中发现了罕见的环伍德石包裹体,明确地证明了它起源于地幔过渡带(Pearson et al. 2014)。此外,重建的Juina钻石中一套脱溶包裹体的成分与预计在下地幔玄武岩系统中形成的矿物非常匹配(Walter et al. 2011)。然而,使用钻石包裹体中发现的不完整矿物组合作为深度标记并不总是没有歧义。例如,白晶石包裹体(以前称为CaSi-walstromite)长期以来被认为代表深度大于~600 km的后退casio3钙钛矿(例如,Harte et al. 1999;Joswig et al. 1999),但有令人信服的证据表明,至少有一些白晶石包裹体起源于上地幔较浅的深度(Brenker et al. 2005;Anzolini et al. 2016)。Woodland et al.(2020)最近为白晶石可能的上地幔起源提供了实验支持。此外,含铁方长石是解释的下地幔起源中最常见的包裹体(Kaminsky 2012),但有证据表明,含铁方长石参与了跨越上地幔-下地幔边界的矿物共生(Hutchison et al. 2001),它可能是在过渡带甚至上覆上地幔中可能发生的钻石形成反应的副产物(Thomson et al. 2016;Bulatov et al. 2019)。数据质量。精确和准确的化学分析是可靠的热气压测定法的一个有时被低估的先决条件。钻石中的内含物通常很小,它们的化学分析可能具有挑战性。此外,有些温度计可能对临界浓度水平下元素的分析不确定度特别敏感。例如,Ziberna等人(2016)使用Nimis和Taylor(2000)的单斜云石气压计探索了分析误差对P估计的传播,并发现常规的电子探针分析条件可能不足以对钻石中的许多内含物进行有意义的气压测定。显然,不仅计数统计数据不准确,而且样品制备不当和标准化不当也可能是一个问题,并可能导致错误的结果。一项对岩石圈钻石中600多个斜辉石包裹体的公开化学数据的调查表明,其中约30%的氧化物总重量百分比< 98.5%或> 101%,或阳离子总和< 3.98或> 4.03原子/公式单位,因此,质量肯定不是很好!请注意,上述截止值充分考虑了在电子探针分析中将所有铁都视为Fe2+可能产生的影响。当研究从共存的矿物相中分离出来的包裹体时,下一个关键步骤是定义它们的原始共生。即使单矿物温压计经常会派上用场,它们也总是假定所研究的矿物最后与特定的其他矿物达到平衡。至于斜辉石和石榴石,它们是几种类型地幔岩石的主要成分,基于主要或微量元素浓度的简单成分筛选可能有助于区分最典型的共生品种(例如,Griffin et al. 2002;gr<s:1>等人,2004;Nimis等人,2009;Ziberna et al. 2016)。不寻常的成分不能被现有的分类方案敏感地区分出来,或者超出了温度计校准的成分范围,以及在成分过滤中不适当地幸存下来的假阳性,可能会破坏热气压计数据的可靠性。当可用样本的数量和性质允许时,通过调查大型数据集可以最大限度地减少这个问题。即使采取了所有谨慎的步骤来选择合适的成分数据,测温估计也会有误差。正如前一节所讨论的,其中一些错误是系统的,并导致使用不同方法获得的估计之间可预测的不一致。 在所有其他情况下,非接触包涵体对容易出现潜在的错误,因为单个包涵体可能在不同的时间和不同的条件下被纳入(见上文),因此可能不处于相互平衡状态。据报道,在西伯利亚米尔金伯利岩的一颗钻石中的石榴石中,单个钻石中多个包裹体之间的T差最大(Griffin et al. 1993)。根据所使用的石榴石镍温度计的校准方式,计算出的差异差异很大,即Ryan等人(1996年)的校准约为400°C, Canil(1999年)的校准约为230°C。然而,钻石中非接触内含物之间的不平衡似乎是例外,而不是规则,因为一旦考虑到不同温度计之间的系统偏差,单个钻石中的多个内含物通常会产生相似的P-T估计值(Stachel和Harris 2008;Stachel and Luth 2015)。同生与原生。几十年来,金刚石中包裹体的热气压测量依赖于共生假设,即包裹体在金刚石生长过程中形成或完全再结晶。这一假设的一个推论是,包裹体记录了钻石形成的条件,除了接触包裹体时潜在的包裹后修饰。同生说的主要证据是,钻石通常把自己的形状强加在内含物上。Nestola等人(2014)对这种被广泛接受的范式提出了挑战,他们发现了一些橄榄石包裹体的原生形成(即在钻石形成之前形成)的证据,这些包裹体根据形态标准被归类为同生。随后在其他橄榄石、斜辉石、石榴石和镁铬铁矿包裹体中发现了类似的证据(Milani et al. 2016;Nestola et al. 2017,2019b;Nimis et al. 2019)。钻石中的内含物可能是原生成因的,这表明它们可能记录了钻石形成之前的条件。这可能与孤立的包裹体特别相关,如果它们的宿主钻石是由热流体或熔体侵入相对较冷的地幔岩石形成的。先前存在的矿物颗粒在金刚石生长过程中,在它们合并完成之前,是否有足够的时间完全重新平衡以适应不断变化的热条件?矿物调整其成分以适应新的物理化学条件所需的时间在很大程度上取决于温度和粒度,并且可以通过扩散动力学来估计。例如,直径为200微米的斜辉石颗粒将在1200℃以上不到10万年的时间和900℃以下几微米的时间内完全重置其Ca/Mg比率(辉石测温中的主要温度传感器)(图7)。这些时间跨度是否足够短或足够长,以完全重新平衡取决于金刚石的生长速度,这是未知的。很容易推断,原生包裹体的不完全再平衡,特别是最大和最浅(即最冷)的包裹体,可能解释了非接触包裹体在P-T图中的一些散射(图8)。在相同来源的钻石包裹体和捕虏体中偶尔观察到的P-T条件的相似性,可能归因于钻石是由热平衡介质形成的(Nimis 2002),也可能归因于在短暂的钻石形成过程中原始成因包裹体缺乏再平衡(Nestola et al. 2014)。Exsolved夹杂物。钻石中的一些内含物显示出包裹后分解的证据。报道的例子有斜辉石与正辉石脱溶片层(±coesite),代表未混合的高t斜辉石,或多或少多数石榴石与斜辉石共生,代表未混合的高p多数石榴石。溶解的斜辉石包裹体并不常见,仅在纳米比亚砂矿和南非Voorspoed金伯利岩的钻石中报道过(leest et al. 2003;Viljoen et al. 2018)。斜辉石和正辉石的共存使这些包裹体成为辉石测温的理想选择,但暴露的组合不能提供气压估计。由于辉石温度计具有较小的p依赖性,因此在这些情况下,对当地捕体地热的投影可以提供对钻石起源深度的合理估计(Nimis et al. 2020)。如果已知出溶相的组成和相对比例(例如,通过将电子探针数据与2D或更好的3D图像分析相结合),则可以重建斜辉石的整体出溶前组成。综合成分可以提供钻石生长时的T估计,假设在捕获期间与单独的正辉石平衡,或者在没有正辉石的情况下,提供钻石形成的最小T估计。 在报告的案例中,重建斜辉石质的测温提供了钻石形成温度接近地幔绝热层的证据,并且在喷发时比周围地幔温度高~200°C。试图将单斜辉石气压计应用于重建的成分,得到了不合理的低压估计,表明地层是在强烈改变的地幔环境中形成的(Nimis et al. 2020)。出溶结构在岩石圈下钻石的多数石榴石包裹体中很常见(Moore and Gurney 1989;野生植物1990;Harte and Cayzer 2007)。这些结构被解释为反映了对流地幔或地幔柱中钻石缓慢上升时的减压。出溶体的存在意味着基于多数石榴石成分的压力估计是最小的。利用重建的预溶石榴石组成将得到更接近金刚石地层的条件。即便如此,由此得出的估计也可能是最小的,因为石榴石可能起源于斜辉石出反应以外的深度(见前面的气压测定部分)。图8显示了来自世界各地的岩石圈钻石的P-T数据汇编,使用最可靠的温度计组合获得。该汇编包括100颗黑曜石型钻石中的斜辉石包裹体和77颗黑曜石型、12颗黑曜石型和10颗韦氏石型钻石中的正辉石-石榴石包裹体对。绝大多数斜辉石包裹体是在它们的寄主钻石中分离出来的,通常应该记录的条件接近于钻石的形成条件。正辉石榴石对由接触组合和非接触组合组成。接触对可能在合并后重新平衡,并应提供接近地幔中最终居住条件的条件。此外,用同样的方法计算了可用于适当矿物分析的金刚石不同捕虏体的P-T数据。在比较单斜辉石和正斜辉石-石榴石的两个测温图时(图8a,b),重要的是要考虑所使用的测温方法之间已知的系统不一致性(见上文)。正辉石榴石温度计倾向于高估1100°C以下的温度,低估1100°C以上的温度。这至少在一定程度上解释了为什么正辉石榴石包裹体的温度范围受到限制。尽管如此,除去一个明显的异常值,所有的P-T值都落在金刚石稳定性场的0.3 GPa以内,证实了温压表的总体可靠性。大部分P-T结果位于Hasterok和Chapman(2011)的35- 40 mw /m2导电地温之间,即在典型的克拉通岩石圈条件下。大多数例外记录了较高的温度条件,在斜辉石的情况下,可能达到地幔绝热层。所有这些例外都是由非接触夹杂物组成的。一种可能的解释是,一些钻石是由来自岩石圈下对流地幔的热不平衡介质形成的。金刚石形成过程中地幔温度不均匀性的另一个潜在来源可能是金刚石生长过程中的不完全克拉通化。例如,在Kaapvaal克拉通,一个稳定的地热梯度在2.5 Ga左右建立,随后是长期冷却(例如,Michaut和Jaupart 2007;Brey and Shu 2018)和重新加热(Griffin et al. 2003),但许多Kaapvaal钻石的年龄明显超过2.5 Ga(见Shirey and Richardson 2011的综述),或者可能是在当地岩石圈的重要扰动期间形成的(Griffin et al. 2003;Korolev et al. 2018)。因此,这些钻石中的许多可能记录的条件比金伯利岩喷发时的条件更热或更冷也就不足为奇了(例如,Nimis et al. 2020)。接触的正辉石榴石包裹体对值得进一步讨论。在现有的数据集中,它们中的大多数(31个中的23个)都是来自De Beers Pool(南非金伯利)的钻石。Phillips等人(2004)对这些内含物进行了研究,他们仔细记录了接触与非接触内含物记录的P-T估价值之间的差异。这些工作者发现,非接触包裹体产生的P-T值高于接触包裹体,并将这一观察结果归因于接触包裹体在降低T条件下的圈闭后再平衡,可能伴随着地幔的隆起和减压。他们还正确地注意到,包裹后的弹性修饰没有足够的幅度来调和观察到的接触和非接触夹杂物之间的P-T差异。然而,Nimis等人(2020)观察到菲利普斯等人的触摸和非触摸夹杂物。 (2004)的数据集记录了在同一金伯利岩中比地幔捕虏体和异晶更冷的地热状态,对长期冷却假说提出了质疑。而非接触包裹体相对较冷的特征可能反映了捕虏体记录的岩石圈最终热平衡之前的旧条件,而接触包裹体的数据很难解释。Weiss等人(2018)推测,接触包裹体记录的低温可能与较晚的冷的板状流体渗透有关,这些液体的残留物被保存在来自同一地点的一些云状和涂层钻石中。然而,目前尚不清楚为什么这些流体的渗透在取样的地幔捕虏体中几乎没有留下任何痕迹。从其他地方接触到的正辉石榴石包裹体可能太少,没有统计学意义。尽管如此,在所有这些情况下,接触包裹体产生的温度明显低于来自相同深度的其他钻石中的非接触包裹体(图8b)。Stachel和Luth(2015)和Viljoen等人(2018)提供了更多的热气压证据,表明钻石形成后可能会发生冷却。然而,Stachel和Luth(2015)也描述了在相同的钻石中,接触夹杂物产生的温度与非接触夹杂物相似甚至略高(高达40°C)。钻石深度系统学。无论包裹体的性质和使用的热气压测量方法如何,岩石圈钻石的压力频率分布基本上是单峰的,模态为~6 GPa,对应于~190 km左右的深度(图8;另见Stachel和Harris 2008年和Stachel 2014年使用不同的温度计和部分不同的数据集获得的大致相似的分布)。斜辉石和正辉石石榴石包裹体的压力分布之间的细微差异,可能主要归因于所使用的不同温度计组合之间的不一致,也可能是采样偏差。因此,这些差异可能并不显著。石墨-金刚石和岩石圈-软流圈边界所施加的硬约束不能完全解释观测到的深度分布,其显示接近这些边界的钻石频率减少(图8)。Nimis等人(2020)通过对不同包裹体类型使用不同的热气压测量方法组合,从三个南非金伯利岩源获得了非常相似的深度分布,深度模式为180±10 km。他们还发现,使用相同的热气压计方法计算的来自相同来源的地幔异种晶体的深度分布没有明显的相关性。他们的结论是,观察到的钻石深度分布与金伯利岩取样效率无关,因此具有遗传意义,并可能具有全球意义。丰富的石榴石包裹体ni -in-garnet数据也支持了观测深度分布的全球意义(图9)。这些数据显示了~1200°C的模式,但通常没有压力约束。假设石榴石最后在35和40 mW/m2模式地热之间的“平均”条件下达到平衡,类似于斜辉石和正辉石石榴石包裹体(图8),则石榴石中镍的温度模式将再次对应于~6 GPa的压力模式和~190 km的深度模式。频率在~1200°C时的缓慢增加和在较高温度下的急剧下降(图9)也反映了斜辉石和正辉石-石榴石包裹体在~6 GPa及更高压力下的频率变化(图8)。岩石圈中的钻石深度分布可能是建设性和破坏性过程的结果(Nimis et al. 2020)。特别是,在小于~170 ~ 190 km的深度,金刚石浓度的逐渐减少可能反映了(i)上升母介质对金刚石的降水速率的减少或(ii)这些深源介质在较浅水平的数量的逐渐减少。很可能,这两种机制共同限制了钻石窗口浅层部分的钻石禀赋。岩石圈底部附近钻石的快速减少可能反映了(i)沿特定的P - T路径,P的增加和T的增加可能对某些地幔流体/熔体中的碳溶解度产生的相反影响之间的正平衡,或(ii)由于与碳不饱和软流圈流体/熔体的反应,钻石禀量逐渐减少。这两种假设很难证明或反驳,因为我们对高P-T下天然含碳熔体和流体的行为的了解仍然有限。 在典型岩石圈底部附近,熔融交代岩性的普遍存在与金刚石频率的减少相吻合,这表明熔融交代可能是金刚石的重要破坏因素(例如,McCammon et al. 2001)。这些岩石圈深层根部的碳再活化可能最终有助于在较浅深度建立钻石浓度(Nimis et al. 2020)。认识到~190公里左右的深度对钻石特别有利,这可能对它们的矿物学勘探具有有趣的意义。从这些特定的深度以异晶地幔矿物(如斜辉石或石榴石)的出现为特征的地区可被视为高度优先的目标。尽管如此,由于异种晶体与钻石深度分布可能不相关,因此即使是来自钻石最有利深度范围的有限异种晶体记录也可能表明具有重要的钻石潜力(Nimis et al. 2020)。图10显示了含有多数石榴石的钻石的压力估计的全球分布。目前,这些包裹体是唯一能提供岩石圈下钻石丰富的定量气压数据的类型。调查这些数据的一个困难在于它们可能代表最小P估计。正如前一节所解释的,这种限制是由于经常出现脱溶,除了Thomson等人(2021)的晴雨表外,在原始矿物组合中可能没有伴随的斜辉石,而大部分石榴石包裹体是在包裹中分离出来的。如果只考虑据报道不含斜辉石的包裹体,则橄榄岩和辉生岩/辉生岩石榴石分别得到两种不同的单峰分布(图10)。利用Beyer和Frost(2017)气压计,橄榄岩包裹体主要聚集在175 ~ 225 km和225 ~ 275 km深度区间,靠近岩石圈-软流圈边界。因此,这些包裹体中的大多数可以更恰当地归类为深部岩石圈。辉长岩/辉长岩包裹体的平均深度要深得多,大部分落在300 ~ 500 km的深度区间。如果还考虑含有外溶辉石岩的包裹体,则在~400 km处的榴辉石岩/辉石岩包裹体峰值变得更加突出。使用Thomson等人(2021)的气压计,榴辉岩/辉长岩包裹体的分布被拉伸到稍高的压力,最大估计深度为613±44 km,而许多橄榄岩包裹体被转移到更大的深度,很好地落入岩石圈下区域(图10)。一些最深(>450公里)的多数石榴石落在斜辉石消失的地区,因此Beyer和Frost(2017)的气压计可能低估了这些样品的压力。然而,Thomson等人(2021)的机器学习晴雨表不受这种影响,并没有表明这些石榴石的深层起源(图10)。使用两种气压计对这些极深样品的深度估计之间的差异在-65和+81公里之间(平均+17公里)。应该考虑到,在某些情况下,由于缺乏准确的成像数据,可能会忽略细斜辉石渗出的发生。使用改进的BSE和EBSD成像可以在名义上“无斜辉”的多数包裹体中识别出游离的斜辉石(参见Harte和Cayzer 2007;Zedgenizov et al. 2014)。迄今为止,未发现共存斜辉石的13个Juina多数包裹体给出了在353至455公里(使用Beyer和Frost 2017年晴雨表)或368至462公里(使用Thomson等人2021年晴雨表)之间的有限范围内的深度估计,并在一定程度上解释了图10中约400公里处的突出峰值。如果这些内含物还含有脱溶物,并且现有的电子探针数据不能忠实地反映它们的原始成分,那么总体峰值可能会移到更大的深度,或者至少总体深度分布会比图10所示的更平滑。Thomson等人(2021)认为,南美钻石中的多数石榴石包裹体可能都含有外溶物。他们还注意到,在8个既可溶又可重构的包裹体成分中,溶出量是计算压力的简单函数。因此,他们设计了一种经验校正,可以应用于所有的内含物,其中重建的大块成分是不可用的。当这一修正应用于南美钻石时,榴辉岩多数石榴石包裹体的峰值被平滑并移动到大约100公里的深度。 目前尚不清楚这种基于有限数量的纳入的修正是否在全球或地方范围内具有普遍有效性。无论如何,在~300至~500 km之间的深度估计的高丰度,即使对于肯定不含斜辉石质或其原始成分被重建的包裹体,也表明大多数榴辉石质/辉石质多数石榴石确实起源于此深度范围(Harte 2010),或者如果Thomson等人(2021)对南美钻石的校正有效,则可能起源于~300至~600 km的稍微扩大的范围。这一范围与其他岩石圈下含硅酸钙包裹体(300-360 km;Brenker et al. 2005;Anzolini et al. 2016),并通过弹性气压法获得单个铁长石包裹体的最小深度估计(~450 km;Anzolini et al. 2019)。进一步有趣的是,300-600公里范围对应于碳化MORB板最有可能与固相相交的深度,为集中熔化和潜在的钻石生长提供了理想的环境(Thomson et al. 2016, 2021)。考虑到许多没有气压数据的镁镁镁石包裹体也可能是由俯冲相关的熔体形成的(Thomson et al. 2016;Nimis et al. 2018) 300-600公里区间的钻石实际比例可能比图10所示的还要大。读者可参考Walter等人(2022年,本卷)对这些和相关方面进行彻底的讨论。第二个有利于钻石形成的区域可能位于上地幔-下地幔边界和上下地幔附近(Harte 2010)。该地区钻石形成的证据来自于单个钻石内的一些关键矿物组合,主要由方长石、(Mg,Fe)SiO3、(Mg,Fe)2SiO4、白方辉石(以前称为TAPP)、白方辉石、刚玉和SiO2的各种组合组成,以及铝辉石相(据信形成了高度多数的石榴石)与方长石或白方辉石(Hutchison et al. 2001;哈特2010)。虽然在一些钻石中,只有部分关键矿物组合得到了体现,但观察到的矿物组合之间的总体一致性使Harte(2010)假设大量钻石可能来自相对较窄的600至800公里的深度范围。在这个深度范围内的钻石中,大多数包裹体都是超镁铁性的,而在300-600公里的范围内则主要是基性包裹体。矿物包裹体在钻石中非常罕见,仅在被检查的钻石中出现约1%,在个别地点之间存在显着差异(Stachel和Harris 2008)。此外,只有少数可能以包裹体形式出现在钻石中的地幔矿物或矿物组合适合于热气压测量。因此,只有一小部分天然钻石可以提供关于其形成的P-T条件的可靠信息,这些钻石大多是橄榄岩。随着新的热气压测量方法的发展或现有方法的改进,这一比例将来可能会增加。一个重要的潜在和理想的改进领域涉及榴辉包裹体,它代表岩石圈钻石包裹体的主要部分,但现有的气压测量方法仍然提供不一致的结果。与Ziberna等人(2016)对铬透辉石进行的测试类似,对斜辉石-石榴石对进行专门测试,可能会更清楚地了解这些包裹体的可靠大气压测定所需的最低分析质量。将非传统方法(如弹性气压测量)的适用性扩展到非理想结构和更多样化的包裹体矿物列表,也可能对适合热气压测量的钻石数量产生巨大影响,尽管对弹性保存包裹体的需求通常限制了应用于相对较小的未被裂缝包围的包裹体。弹性方法的一个有趣的优点是,它可以依赖于非破坏性技术,因此,可以使样品完好无损,并准备好与其他方法进一步调查。这种调查需要使用内部应变保存良好的未损坏样品。不幸的是,传统的破坏技术用于暴露钻石中包含的矿物质,禁止未来对这些独特的样品进行弹性气压测定。重新分析一些先前研究的包裹体或含金刚石捕虏体可能提供一种相对轻松的方法来增加金刚石P-T数据的数量。由于分析质量不达标,195篇已发表的斜斜辉分析中有4篇和156篇已发表的正斜辉分析中有6篇在准备本综述时不得不放弃。 此外,25颗斜辉石被拒绝,因为它们的成分刚好落在用于分类石榴石相关的铬透辉石的主元素成分场之外(Ziberna et al. 2016)。这些透辉石的微量元素分析可能允许更强大的区分(例如,基于Sc/V关系;Nimis et al. 2009)的石榴石相关样品,其中许多可能已被证明适合热气压测量。重新整合所有这些样品将使含铬透辉钻石的P-T数据数量显著增加27%,总体钻石的P-T数据数量增加17%。扩大钻石的P-T数据数量将在减少样本偏差方面产生显著的积极结果。样本偏差目前是一个主要问题,因为目前可用的具有统计意义的足够可靠的P-T估计的地点最多只有三个,而且全部来自Kaapvaal克拉通(Nimis et al. 2020)。减少样本偏差的好处是更好地评估单个地点的钻石深度分布,更可靠地评估钻石P-T系统及其在钻石形成过程和保存方面的重要性。如果钻石的P-T数据的数量很重要,那么它们的内部一致性可能更重要。事实上,大多数用于钻石内含物的热气压测量方法的组合确实显示出某种程度的不一致。这些不一致主要源于对矿物中固溶体的简化热力学处理,以及用于校准的数据的不确定性,这些数据要么来自实验,要么来自独立的捕虏体热气压测定。当比较不同包裹体类型的数据时,热气压测量方法的内部一致性尤为关键,这在金刚石研究中通常是必要的。这是另一个潜在的科学进步领域,可能最终有助于提高对金刚石形成过程的理解。这次审查是许多人的知识、专业知识和慷慨的结果。多年来,我与他们中的许多人,特别是与W.R. Taylor、H. gr<e:1>特、L. Ziberna、F. Nestola和R.J. Angel卓有成效的合作,对我个人对钻石温度测量学的贡献具有决定性的重要性。然而,事实上,有更多的人通过他们大量的科学工作,并在大约50年的时间里为科学界提供了宝贵的数据,为这一综述做出了贡献。他们中的大多数出现在下面的参考列表中。我对所有这些人表示衷心的感谢,特别是T. Stachel,他慷慨地分享了一个钻石内含物数据库,这是建立本工作中使用的更新数据库的有用基础。我还要感谢A. Thomson,感谢他在大多数气压计计算方面提供的帮助,感谢H. gr<s:1> tter、G. Brey和编辑T. Stachel,感谢他们非常有益和彻底的审查和评论。 由于斜辉石-正辉石包裹体对在钻石中非常罕见,因此通常使用Nimis和Taylor(2000)的替代单斜辉石版本更为实用。双辉石Taylor(1998)和单斜辉石Nimis和Taylor(2000)方法在应用于含有两种辉石的地幔超基性岩石时提供了几乎无法区分的结果(Nimis和gr<e:1> tter 2010)。虽然Nimis和Taylor(2000)的温度计只使用一种辉石来计算T,但其应用明确要求两种辉石都是矿物组合的一部分,并且处于化学平衡状态。因此,这种方法只适用于属于辉长岩组或韦氏岩组且含有斜辉石包裹体的钻石,无论是孤立的还是与其他矿物伴生的。如果正斜辉石不是原始矿物组合的一部分,如石英包裹体的情况,则单斜辉石测温只能提供最小的T估计。异常低的T值(例如,远低于当地地热)可能是斜辉石质的标志。Ziberna等人(2016)认为Ca/(Ca + Mg)的摩尔比大于0.5也应该被认为是可疑的,因为只有~1%的饱和正斜辉石高于这个值。尽管如此,仅根据成分标准对斜辉石类进行鉴别通常是不可能的。Simakov(2008)提出了一种不同的、更复杂的单斜氧甲烷温度计校准方法,它肯定能提高1500°C以上的性能,但不能超过典型的“岩石圈”温度T < 1400°C范围(请注意,Simakov 2008的图10中报告的TNimis和Taylor 2000估计是不正确的,并且过度建议高估1300°C以下的温度)。Brey和Köhler(1990)的Ca-in-Opx温度计是辉石温度计的单斜辉石版本,可以提供与单斜辉石温度计相补充且独立的估计值。如果在900°C以下进行校正,这两种单辉石方法被证明是相互一致的(Nimis和gr<s:1> tter, 2010)。然而,由于可能具有斜辉石(即不含斜辉石)亲和力的包裹体在钻石中相对常见(Stachel和Harris 2008),因此对正辉石包裹体的TCa-in-Opx估计仅为最小T值的可能性远远高于单斜辉石温度计。因此,它在钻石研究中的作用相当有限。石榴石与橄榄石平衡时的Ni含量对T非常敏感,显然与P无关(Ryan et al. 1996;Canil 1999)。由于橄榄石的镍含量在地幔捕虏体(O 'Reilly et al. 1997)和钻石包裹体(Griffin et al. 1992)中都显示出很小的变化;Sobolev et al. 2008), T可以单独从Nigarnet中检索,通过假设共存的森林橄榄石具有适当的Ni含量。一个有用的参考值是地幔橄榄石的平均值(平均值±标准差= 2900±360 ppm;Ryan et al. 1996),这也接近全球钻石中橄榄石包裹体的平均值(Stachel and Harris 2008;Sobolev et al. 2009)。如果可能,可以使用同一地点钻石中橄榄石包裹体的平均铌橄榄石值(例如,卡拉哈里克拉通的51个包裹体的铌橄榄石平均值为3150±200 ppm, Griffin et al. 1992;来自阿尔汉格尔斯克省的88个夹杂物为2700 ppm, Malkovets等人,2011年)。选择其中一个或另一个值最多会使最终的T估计值改变几十度(图1)。Nigarnet的浓度可以用激光烧蚀电感耦合等离子体质谱计(LA-ICP-MS)、离子微探针(SIMS)或质子微探针(PIXE)测量,通常可以确保精度在几ppm以内。与橄榄石平衡的假设可能需要成分过滤(例如,gr<s:1>等人,2004年),以排除任何非橄榄岩石榴石。镍石榴石温度计已广泛应用于钻石中含铬-硫铁矿石榴石的研究(例如,Griffin等人,1992,1993;Davies et al. 2004a;Viljoen et al. 2014;De Hoog et al. 2019)。然而,它的校准多少有些争议。Ryan等人(1996)利用橄榄石-石榴石fe - mg交换温度计(O 'Neill and Wood 1979)和MacGregor(1974)、Brey和Köhler(1990)正辉石-石榴石大气压计组合得出的T值,根据地幔捕虏体对其进行了校准。Canil(1999)根据温度≥1200°C的实验校准了它。两种校准在~1100°C时给出相同的结果,但在较低和较高的T下逐渐偏离(图1)。因此,通过Canil(1999)公式获得的估计值通常会跨越较窄的T区间。考虑到石榴石包裹体中最典型的Ni含量范围,TCanil-TRyan的最大差异约为。 +100°C在20 ppm镍和约-250°C在180 ppm镍。有各种各样的说法认为,无论是校准,甚至是它们的“平均值”,在应用于地幔捕虏体时,都与其他独立的温度计最一致(例如,Ryan等人,1996;Canil 1999;De Hoog et al. 2019;Czas et al. 2020;Nimis et al. 2020),但仍缺乏使用内部一致温度计作为参考的明确评估。Sudholz等人(2021a)对镍石榴石温度计进行了最新的改进尝试,他们根据新的实验在相对较窄的温度范围(1100-1325°C)对其进行了重新校准,并引入了石榴石中Ca和Cr含量的校正项。当使用Nimis和Taylor(2000)斜斜长钙铁石温度计对捕虏体进行独立估计时,Sudholz等人(2021a)校准显示,相对于Canil(1999)实验校准,在1100°C以上的总体精度有所提高,但总体精度较差,并且在较低温度下略强的渐进式高估(见Sudholz等人2021a中的图7)。石榴石与地幔橄榄石平衡时的Mn含量对T敏感(Smith et al. 1991),假设橄榄石中Mn含量恒定,可作为单矿物温度计。该温度计依赖于电子探针数据,在没有微量元素数据时被提出作为石榴石镍温度计的替代品(gr<s:1> tter等人,1999;克莱顿2009)。声明的精度相当差(大多数为±150°C),并且由于缺乏独立的橄榄石数据,可能会在高t处发生严重的低估。因此,该方法可能在钻石勘探期间回收的大量石榴石异晶的调查中有一定的用途(例如,gr<s:1> tter和Tuer 2009),但其在钻石热压测量中的效用有限。Ashchepkov等人(2010)分别校准了与斜辉石或橄榄石平衡的石榴石的两个单矿物温度计。这些方法是斜辉石-石榴石或橄榄石-石榴石对的fe - mg交换温度计的简化版本,其中斜辉石和橄榄石的组成是由石榴石的组成模拟的。它们的准确性不可能比原来的两种矿物配方更好,后者本身也有问题(见下文)。此外,要计算t, P必须是独立已知的。在对来自Udachnaya金伯利岩的地幔包体的测试中,这些温度计再现了Brey和Köhler(1990)的辉石温度计计算出的温度,在用正辉石-石榴石气压计计算出的压力下,温度在±100°C左右。由于缺乏共存的正辉石的成分数据(见下文),对P值的估计存在问题(见下文),这可能使这些温度计对金刚石热气压测定不感兴趣,特别是在调查小种群的钻石时。橄榄石中Al含量对T敏感,可作为测温指标。De Hoog等人(2010)在地幔捕虏体上对与钻石相关条件下的橄榄岩橄榄石的第一个al -in-橄榄石温度计进行了经验校准,使用了由Brey和Köhler(1990)的双辉石温度计和正辉石-石榴石气压计组合得出的P-T估计值。Bussweiler等人(2017)通过分析高p - t实验电荷中的橄榄石,对该方法进行了重新校准。克拉通地幔橄榄石中Al的含量范围为~ 1ppm ~ ~ 250ppm,通过LA-ICP-MS或SIMS等技术可以方便地进行分析,确保精度在几ppm以内。使用非常高的探针电流(例如,≥200 nA)和长计数时间(≥100 s)的电子微探针分析(EMPA)可能是一种有效的替代方法(Batanova等人,2018;D 'Souza et al. 2020)。尖晶石橄榄岩或强交代石榴石橄榄岩的铝-橄榄石温度可能分别被严重低估或高估。使用ai vs V图进行成分筛选可能有助于区分这些“不安全”橄榄石(Bussweiler et Al . 2017)。如果橄榄石与石榴石没有物理上的联系,就像钻石中的许多包裹体一样,筛选尤其重要。对于天然地幔橄榄岩,与单斜辉石温度计(Nimis and Taylor 2000)相比,Al-in-橄榄石温度计给出了系统性更高的估计值(~50°C,在900°C以下略有增加),这可能反映了由于Na损失,实验橄榄石中的Al掺入受到轻微抑制(Bussweiler et Al . 2017)。差异很小,足以确保方法的成功应用,但在比较使用不同温度计获得的估计值时应予以考虑。al -in-橄榄石温度计尚未广泛应用于钻石研究(参见Korolev等人2018年的应用示例),这是由于其最近的发展以及对al进行非常规分析的必要性。 考虑到钻石包裹体中含有大量的橄榄石(Stachel and Harris 2008),它的应用在未来可能会显著增加。尖晶石与地幔橄榄石平衡态Zn含量对t敏感,由于地幔橄榄石Zn含量几乎是恒定的(平均值±标准差= 52±14 ppm),所以尖晶石Zn含量可以直接作为温度计使用(Ryan et al. 1996)。尖晶石中锌的浓度可以通过SIMS或PIXE进行测量,其精度在几ppm以内。当前版本的尖晶石锌温度计使用Ryan等人(1996)版本的石榴石镍温度计在680-1180°C范围内对共存的石榴石进行校准。因此,它至少具有与石榴石镍温度计相同的不确定度(见上文)。铁-镁交换温度计。这些流行的温度计是基于两种矿物相之间的Fe2+和Mg的分布,可以用来计算石榴石与橄榄石、正辉石或斜辉石相结合的包裹体的T。利用地幔捕体矿物成分进行的测试和实验表明,这些温度计要么精度低,要么存在系统误差,或者两者兼而有之(见Nimis和gr<s:1> tter 2010年综述)。在橄榄岩系统中,橄榄石-石榴石和斜辉石-石榴石温度计的精度最低,误差可能超过200°C。Harley(1984)的正辉石-石榴石配方显示出最高的精度,但系统地高估了温度< 1100°C,低估了温度> 1100°C(地幔捕虏体平均高达150°C)。所有这些差异可能反映了交换反应对T的适度敏感性(特别是橄榄石-石榴石),所使用的固溶体模型的过度简化(特别是石榴石和斜辉石),以及矿物中的Fe3+/ΣFe比率与校准样品中的不同。Matjuschkin等人(2014)通过实验证明了被忽略的Fe3+对正辉石-石榴石和橄榄石-石榴石温度计的巨大影响。他们在5gpa和1000-1400℃的无na CaO-FeO-Fe2O3-MgO-Al2O3-SiO2体系中进行的实验表明,如果使用Fe2+代替石榴石中的总铁进行计算,温度估计可能会显着提高。然而,Mössbauer橄榄岩捕虏体中正辉石和石榴石的数据表明,这两种矿物之间的Fe3+分配受到正辉石中压力和Na含量的影响(Nimis et al. 2015)。因此,在na轴承系统中进行相关压力范围内的测试是可取的。Nimis和gr<e:1>特(2010)利用平衡良好的地幔捕虏体的双辉石测温作为校准,对哈雷(1984)的温度计进行了经验校正。相对于未校正的温度计,校正后的版本通常在低温度下产生较低的T估计,在高温度下产生较高的T估计,并且对于在“平均”地幔氧化还原条件下平衡的正辉石-石榴石对可能更准确。另一方面,如果氧化还原条件远非“平均”,特别是在高压下的强氧化条件下,当P和T都是通过与正铁榴石-石榴石气压计结合迭代计算时,它可能会产生很大的误差(>到> 100°C) (Nimis et al. 2015)。考虑到在岩石圈深处,金刚石在相对广泛的氧化还原条件下是稳定的(∆logfO2 FMQ≈-5至-2),这个问题可能很严重。此外,应用于某些捕虏岩套往往会产生频繁的高p - t异常值(gritter, pers)。通讯)。尽管存在公认的系统差异,但由于其较低的P依赖性,原始哈雷(1984)版本可能仍然更适合金刚石热气压测量,这减少了迭代P - t计算过程中的误差传播。然而,Harley(1984)温度计在远离1100°C的温度下的系统偏差,特别是当所有Fe都被视为Fe2+时,在数据评估中应该考虑。斜辉石-石榴石温度计是唯一可行的方法,可以从钻石中榴辉包裹体的主要元素组成中测定T。这种温度计在文献中有许多版本。最近的校准(Nakamura 2009)是基于一个扩展的实验数据库,涵盖了与钻石(1.5-7.5 GPa, 800-1820°C)相关的P和T范围内的基性和超基性成分,并结合了斜辉石和石榴石的最新固溶体模型。旧的,更简化的,但在某些情况下,这种温度计的版本仍然很流行,不能同样很好地重现同样的实验,并且经常显示出系统的偏差随着T和成分的变化。Nakamura(2009)温度计的标称不确定度(1西格玛水平±74°C)仍然是橄榄岩包裹体最佳表现温度计的两倍左右。 低于检测限的氮含量(ii型钻石)在岩石圈钻石中相对罕见,但在岩石圈下钻石中更为常见,显然不适合用于氮聚集热时测定。必须明确的是,与包体热气压计中使用的热力学平衡相反,氮聚集是一个动力学过程,因此不能提供特定时间温度条件的快照。例如,在金刚石生长期间或之后,由于地幔中热流体的循环而导致的短时间温度升高,可能对氮聚集产生重大影响,并可能导致对计算出的Tres的模糊解释。一般来说,除非钻石相对于它的金伯利岩宿主非常年轻,否则在~1亿年的时间尺度上加热对计算出的地温影响很小,它将接近环境地幔条件,而不是钻石形成时的条件(图3a)。如果钻石在形成后经历了一段时间的长期冷却,那么计算出的Tres可能会大大低估地层温度,而高估最终环境温度(图3b)。正辉石中与石榴石平衡态的Al含量对磷元素有很强的敏感性,是含石榴石橄榄岩最广泛使用的气压计的基础。在这个晴雨表中,P主要取决于正辉石中的Al含量,但必须考虑两种矿物的组成才能获得可靠的估计。在该气压计的许多校准中,Brey和Köhler(1990)的校准是迄今为止最受欢迎的,但较老的Nickel和Green(1985)校准以更好的精度再现了大多数橄榄岩系统的实验压力(Nimis和gr<s:1> tter 2010)。尽管如此,Nickel and Green(1985)的版本是在简单的模型系统假设下制定的,并且需要对al组分活性进行校正,以避免对翡翠成分含量不可忽略的正辉石中P的低估。Carswell和Gibb(1987)和Carswell(1991)提出了两种简单的校正方案,基于对钛如何与正硅氧烷中的Na和Al耦合的不同假设。这两种修正只适用于Na > (Cr + Fe3+ + 2·Ti)或Na > (Cr + Fe3+ + Ti)原子/公式单位的正映石,而Fe3+在实际应用中通常被忽略。Carswell和Gibb(1987)的方法,Brey和Köhler(1990)也采用了该方法作为他们的气压计,产生的P变化略小,但目前还没有坚实的实验基础来选择一种校正而不是另一种。正钇石榴石气压计的T依赖性相当小(~ 0.2-0.3 GPa / 50°C),但仍然足够大,需要对T进行准确的独立估计。当该气压计与哈雷(1984)正钇石榴石温度计结合使用时,应考虑到这一事实,该温度计已知在低T时高估,在高T时低估(见上文)。由哈雷(1984)温度计产生的T估计值的人为压缩必然会导致P估计值的非自然压缩。一般来说,输入T的误差将大致取代沿典型克拉通地热计算的P-T估计,从而限制了在地热上平衡的样品的P-T图中的“明显”散射(图4)。如果目的是定义地幔地热,这可能是一个优势,但可能会因为掩盖了数据的真实分散而导致对高整体精度的误解。双矿物榴辉岩和钻石中榴辉石-石榴石包裹体对的气压测定一直是困扰地幔科学家的难题。已经有几次尝试开发一种适合的晴雨表,该晴雨表是基于对p敏感的平衡gross sular + pyrope +透辉石Ca-Tschermak。根据Beyer等人(2018)最近的一项测试,Beyer等人(2015)的公式比Simakov和Taylor(2000)以及Simakov(2008)的早期公式更能再现自然系统中实验的压力。尽管如此,在Beyer等人(2018)对来自Slave克拉通杰里科金伯利岩的天然包体进行的测试中,榴辉岩的P估计比记录相似t的橄榄岩的正辉石石榴石P估计系统性地低(约1 GPa)。因此,杰里科榴辉岩奇怪地似乎落在更热的地热上。斜斜辉石-石榴石气压计的t依赖性约为0.25 GPa / 50℃,与正斜辉石-石榴石气压计相似。因此,同样在这种情况下,输入T的误差将大致取代沿地热方向的P-T估计值,并且不能解释橄榄岩系统和榴辉岩系统测温结果之间的观测差异。图5显示了斜辉石-石榴石热气压计的进一步测试,用于检测钻石和含金刚石或含石墨捕虏体中的接触或非接触包裹体。 P-T估计反映了Beyer等人(2015)气压计和Nakamura (2009) Fe-Mg交换温度计的迭代。由于气压计的不确定度随着斜辉石中[4]Al浓度的降低而迅速增加,因此Beyer等(2015)不建议将其应用于Si含量>1.985个6氧分子式的斜辉石。这些成分在金刚石包裹体(224个包裹体中的35%)和金刚石异捕虏体(233个捕虏体中的42%)中相对常见,因此必须排除大量样品。即使进一步使用严格的过滤器来排除潜在的低质量化学分析,结果也令人不安:尽管含石墨捕虏体位于石墨稳定性区域,但钻石中的大多数含金刚石捕虏体和包裹体产生的压力过低,远远超出了钻石稳定性区域(图5)。重复使用相同的气压计和温度计的早期配方并没有改善结果(参见Shirey et al. 2013的图10)。造成这种差异的原因尚不清楚。Simakov和Taylor(2000)警告说,基于斜斜石中Ca-Tschermak含量的气压计在应用于蓝晶石或sio2过饱和组合时可能导致不正确的估计,但Beyer等人(2015)的气压计显示,在这些系统中进行实验的计算压力没有系统偏差。榴辉柘榴石,特别是榴辉石的巨大化学复杂性可能在一定程度上解释了气压表在自然系统中的应用表现不佳,在自然系统中,Na、Fe3+、Cr、K和含空位(即Ca-Eskola)成分的含量是可变的。另一个可能的解释是,在斜辉石的化学分析中,气压计对即使很小的误差也具有很高的灵敏度。Beyer等人(2015)提供了一个公式来预测由于SiO2含量的不确定度传播而导致的计算压力的相对误差,假设电子探针分析的相对不确定度为1%:%err = 1.94 × 10-8e10.18398 [Si apfu]。这些模型误差没有考虑P和T的迭代计算带来的额外误差(图5)。SiO2含量相对增加1%可能会将许多不准确的P - T估计转移到钻石领域,这表明斜辉石-石榴石气压计的应用可能需要非常高质量、标准化的电子显微分析来产生有意义的结果。文献中报道的许多分析可能是常规质量,不符合可靠斜辉石-石榴石气压测定的必要标准。尽管如此,分析质量本身不太可能解释所有观察到的差异(图5)。sun和Liang(2015)提出斜辉石和石榴石之间的REE划分不仅可以用作温度计(见上文),还可以用作气压计。不幸的是,在一组独立验证实验中,计算出的PREE和运行压力之间的差异似乎非常大(高达~ 3gpa)。对12块石英榴辉岩、2块石墨榴辉岩和9块金刚石榴辉岩进行的测试,除了2块金刚石榴辉岩外,其余均通过了适当的稳定性场约束;它们在P比金刚石-石墨相边界低0.2和0.5 GPa时下降。提出的REE热气压测量方法很有前景,但在该方法可以可靠地用于钻石研究之前,还需要进一步的测试。如果有微量元素数据,斜辉石和石榴石之间的锂划分也可以用作晴雨表(Hanrahan et al. 2009)。在4-13 GPa和1100-1500°C条件下,对富镁生态成分进行了15次校准实验,压力重现到非常合理的±0.2 GPa (1 sigma),气压计的T依赖性很小(~0.2 GPa / 50°C)。初步测试得出了三个含金刚石捕虏体和四个金刚石中的非接触包裹体对与金刚石相容的条件,而另外两个金刚石的条件比金刚石稳定场低2 GPa。Hanrahan等人(2009)认为,这两个虚假的P值可能反映了非接触内含物之间的不平衡。然而,非接触夹杂物的使用对于该晴雨表至关重要,因为锂的高扩散率可能导致小的接触夹杂物在运输到表面期间和之后重新平衡。与稀土元素晴雨表类似,在将该方法推荐用于钻石研究之前,还需要进行进一步的测试。单矿物指标。Nimis和Taylor(2000)开发了一种经验性的cr -斜辉石气压计,用于与石榴石平衡的铬透辉石。结合Nimis和Taylor(2000)单斜辉石温度计,该气压计独特地允许根据单一矿物包裹体的组成推导P和T条件。cr -in- clinoproxene气压计的t依赖性为0.15-0。 25 GPa / 50°C(金刚石包体成分)一般低于正斜辉石-石榴石晴雨表(~ 0.2-0.3 GPa / 50°C),互补单斜辉石温度计的P依赖性很小(图4)。因此,如果通过迭代计算P和T,输入T或P的误差将以相对于克拉通地热的角度取代斜辉石P - T估计,并将导致在稳态地热上平衡的样品的P - T结果的“明显”散射。Nimis和Taylor(2000)的斜斜氧辉铬气压计倾向于逐渐低估4.5 GPa以上的温度(在7 GPa时高达1 GPa;Nimis 2002;Ziberna et al. 2016)。这是一个已知的人工产物,在比较斜辉石中铬的估计值与使用其他方法获得的估计值时应考虑到这种差异。Nimis等人(2020)提出了一种经验校正,以尽量减少与高P下使用正辉石榴石气压计获得的P估计值的差异,但代价是精度有所降低。这些作者警告说,他们的修正是一种临时措施,以减少独立气压估计之间的不一致。最近,Sudholz等人(2021b)对斜斜辉石中cr -in-斜斜辉石气压计进行了实验重新校准,其结果与P > 5 GPa时与钻石相关或位于平衡良好的橄榄岩捕虏体中的斜斜辉石的经验校正校准非常相似(图6a)。在较低的P下,两种修订后的校准略有差异,但Nimis等人(2020)的校准与正辉石榴石气压计保持了更好的总体一致性(图6b,c)。为了正确应用cr -in-斜辉石晴雨表,分析必须经过成分过滤,以选择与石榴石平衡的斜辉石,并消除某些成分,这些成分的分析不确定度在计算的P. Ziberna等人(2016)提供了一本有用的食谱来执行必要的成分筛选,该食谱略微改进了gr<s:1> tter(2009)提出的早期配方。根据这本食谱,一些组合物需要使用比常规更高的束流和计数次数进行电子探针分析,以尽量减少分析不确定性,而一些不利的组合物应该简单地丢弃。特别是,推荐的Cr/(Cr+Al)mol > 0.1的限制排除了约三分之二的被分类为韦氏体的报道内含物。当应用于经过适当过滤的含石墨或含金刚石的包体和包裹体时,斜辉铬气压计的结果与相应碳相的稳定场一致或在~0.3 GPa范围内(Nimis and Taylor 2000;另见图8a)。先前发表的热气压测量结果缺乏足够的过滤,这在很大程度上导致斜辉石P-T图中相对于正辉石-石榴石对的总体散射更大(参见Stachel和Harris, 2008)。Griffin和Ryan(1995)设计了一种基于石榴石与正辉石和尖晶石平衡态的Cr含量的经验晴雨表。该Cr-in-石榴石气压计后来由Ryan等人(1996)改进,并由gr<e:1>等人(2006)使用不同的Cr/ ca -in-石榴石方法进一步修订。最后两个版本是最广泛使用的,通常与镍石榴石温度计结合使用。gr<s:1>等人(2006)给出的估计值降低了约10%,并且与石墨-金刚石转变所施加的岩石学限制条件更吻合,因此更可取。如果石榴石与尖晶石的共存是未知的,就像钻石中许多石榴石包裹体的情况一样,那么石榴石中的cr方法只能提供最小压力的估计。Simakov(2008)提出了一种简化版的斜辉石-石榴石晴雨表,该晴雨表仅依赖斜辉石中的Ca-Tschermak含量。这种单相气压计的精度不可能比原生的两相气压计好,因此容易出现同样的问题(见上文)。Ashchepkov等人(2017)校准或重新校准了一系列经验单矿物温度计,这些温度计应该适用于橄榄岩和榴辉岩石榴石和斜辉石。当应用于单个地点的地幔捕虏体和钻石包裹体时,这些方法产生了强烈分散的P-T估计(例如,Ashchepkov等人2017年的图9-12),很难与其他更广泛测试的方法通常获得的更规则的P-T分布进行比较。因此,Ashchepkov等人(2017)的金刚石测温单元素公式的实用性值得怀疑。在~7 GPa以上,随着压力的增加,地幔石榴石中多数(即: 八面体配位的Al和Cr被Si取代,Na和Ti含量也有增加的趋势。Collerson等人(2000)利用在镁铁和超镁铁系统实验中观察到的主要矿物组分与压力之间的关系,设计了一种经验的单矿物气压计。该方法后来由Collerson等人(2010),Wijbrans等人(2016)和Beyer和Frost(2017)修订。只有Wijbrans等人(2016)提出了两种不同的校准,一种用于橄榄岩,另一种用于生态系统。Beyer and Frost(2017)和Beyer et al.(2018)的实验测试显示,除了Beyer and Frost(2017)公式外,所有公式都存在系统性偏差或精度降低。这与以前的校准数据集中未充分代表的生态成分特别相关。Beyer and Frost(2017)气压计的校准压力为6至20 GPa,温度为900至2100°C。这些条件覆盖了~575 km深度的岩石圈下地幔条件,对应于下过渡带。再现实验压力的标准误差估计为0.86 GPa。相对而言,考虑到校准所涵盖的巨大压力范围,这种不确定度并不比岩石圈钻石所用气压计的不确定度大多少。温度对压力估计的影响显然可以忽略不计。请注意,不建议将主岩气压计应用于P < 6 GPa平衡的岩石圈石榴石,因为在这种条件下,气压计对分析误差非常敏感(Beyer和Frost 2017)。Tao等人(2018)表明,石榴石中Fe3+的高含量往往会在以前的大多数岩石气压计中产生低估的压力,并提供了Collerson等人(2010)气压计的进一步修订校准,明确考虑了Fe3+。Tao等人(2018)在Fe3+数据可用且测量到的Fe3+/ΣFe比率大于0.2时可能更可取。所有这些主辉石晴雨表都要求石榴石与斜辉石相处于平衡状态。这可能是一个问题,因为随着压力的增加,辉石逐渐溶解成石榴石,在高于~ 15-19 GPa的压力下,根据体积成分和温度,地幔岩石不再含有辉石(Irifune et al. 1986;入船1987;冈本和丸山2004)。金刚石包裹体中含有CaSiO3相和多数石榴石,或多数石榴石中Ca含量高,表明辉石欠饱和状态。在这种情况下,气压计会给出错误的低压(Harte and Cayzer 2007)。如果没有独立的证据表明石榴石是辉石饱和的,任何接近斜辉石出反应的压力估计都应被认为是最小俘获P的估计。Thomson等人(2021)使用机器学习算法设计了一种新型的主要气压计。与传统的多数榴石气压计相比,这种新型气压计明显不受岩石学限制(最明显的是缺少斜辉石),并且在更大的体积组成和实验压力(6-25 GPa)范围内再现实验多数榴石组成,整体精度更高(大多数在±2 GPa范围内)。Thomson等人(2021)警告说,钻石中大多数石榴石包裹体的组成位于实验相对较少的区域,机器学习回归在外推中可能不可靠。因此,钻石内含物的压力预测可能有更大的不确定性。尽管如此,这是唯一可用的气压测量方法,可以估算出大多数石榴石包裹体在斜辉石-出反应之外的压力。当应用于钻石中的内含物时,Thomson等人(2021年)和Beyer和Frost(2017年)气压计获得的压力估计值之间的差异在-2.5至+6 GPa之间。无论采用何种主辉石气压法,如果在夹带后石榴石与斜辉石重新平衡,如未混合包裹体中的典型情况(见下文),则会发生P的低估。弹性的方法。当包裹体被包裹在钻石中时,包裹体和钻石最初处于相同的压力下。在喷发时,作用在金刚石上的压力降至1atm,但由于包裹体和宿主具有不同的弹性特性,包裹体上可能会产生高达数GPa的残余压力。如果我们知道包裹体和矿体的平度和弹性特性,我们就可以反算出两种矿物处于相同压力下的条件。这些条件描述了P-T空间中的一条线,称为圈闭等值线,圈闭条件(即钻石地层P-T)将是等值线上的一个点。
本文章由计算机程序翻译,如有差异,请以英文原文为准。
Pressure and Temperature Data for Diamonds
One of the key scientific questions about diamonds is “how are they formed?” To answer this question, we need to know the diamond-forming reactions and the physicochemical conditions under which these reactions take place. The pressure (P) and temperature (T) of diamond formation are an essential part of this knowledge and their assessment is pivotal to develop predictive scenarios of diamond distribution in the Earth interior. These scenarios may contribute to our understanding of global Earth processes, such as the long-term carbon cycle, and might also eventually improve our capability to select potential targets for diamond exploration (Shirey et al. 2013; Nimis et al. 2020).The evaluation of the P and T of diamond formation can be carried out at two levels of investigation. The first is concerned with formation conditions for individual diamonds or small populations of diamonds from specific sources. This approach has been so far the most widely practiced. The second level considers the statistical distribution of P–T conditions for diamond formation at either local or global scale. This type of investigation is hampered by the difficulty of obtaining large sets of suitable samples from a specific locality or for a statistically significant number of localities, and is therefore unavoidably affected to some degree by sampling bias. Despite inherent limitations, the latter approach is the most appropriate to reveal systematics in diamond P–T distributions and, ultimately, in diamond depth distribution within the Earth.Early reviews of P–T distributions for lithospheric diamonds were made by Nimis (2002), based on thermobarometry of chromian diopside inclusions, and by Stachel and Harris (2008) and Stachel (2014), using a more comprehensive set of thermobarometers. More recently, Nimis et al. (2020) investigated diamond depth distributions for a set of South African kimberlites and provided evidence for systematic trends of likely global significance. The depth distribution for sublithospheric diamonds worldwide was reviewed by Harte (2010). In this contribution, I first describe the methods that can be used to estimate the P–T conditions of diamond formation, highlighting their respective strengths and weaknesses. I then review existing diamond P–T data and their implications for diamond distribution with depth from both a local and a global perspective.Thermobarometry of diamonds can be carried out by estimating P–T conditions of chemical or elastic equilibrium of mineral inclusions contained within them. With some assumptions, the aggregation state of nitrogen substituting for carbon in the diamond lattice can also be used as a thermometer. In some cases, it is possible to derive both P and T estimates for a diamond by combining independent thermobarometric methods. In most instances, however, either P or T estimates can be directly retrieved with sufficient confidence. Below is a list of currently available methods for diamond thermobarometry. Those based on chemical equilibria are specific for different minerals or mineral pairs. Since many inclusions in diamonds occur as isolated grains, detached from their original mineral assemblage, particular emphasis is given in this review to methods that use single minerals to retrieve P and/or T. The following descriptions emphasize thermobarometry applications rather than principles, and the reader is referred to the original publications for details of the specific thermobarometers.Pyroxenes. Many ultramafic mantle rocks contain both clinopyroxene and orthopyroxene. This mineral pair forms one of the most widely used geological thermometers, which is based on the net-transfer reaction of the enstatite component between the two pyroxenes MgSiO3(Opx) ↔ MgSiO3(Cpx). The very small P-dependency of this reaction limits propagation of errors when T is calculated by iteration in combination with an independent barometer. Several versions of this thermometer have been proposed in the literature. Tests based on experimental results and pyroxene compositions in natural xenoliths suggest that the calibration of Taylor (1998) is the most reliable for mantle peridotites and pyroxenites in a T range of 700 to at least 1400 °C (Nimis and Grütter 2010), which covers the entire range of the lithospheric diamond window. The much more popular Brey and Köhler (1990) calibration gives similar results for clinopyroxenes with relatively low Na contents around 0.05 atoms per 6-oxygen formula unit (apfu), but was proven to overestimate with increasing NaCpx (e.g., +150 °C at NaCpx = 0.25 apfu; Nimis and Grütter 2010). For inclusions in diamonds and diamond-bearing peridotitic xenoliths (average NaCpx = 0.12 apfu, median = 0.09 apfu, standard deviation = 0.08 apfu), the Brey and Köhler (1990) calibration will yield estimates that are on average ~50 °C higher that those using the Taylor (1998) calibration. Larger discrepancies may be expected for unusually Na-rich clinopyroxenes.Since clinopyroxene–orthopyroxene inclusion pairs are very rare in diamonds, it is generally more practical to use the alternative single-clinopyroxene version of Nimis and Taylor (2000). The two-pyroxene Taylor (1998) and single-clinopyroxene Nimis and Taylor (2000) methods provide almost indistinguishable results when applied to mantle ultramafic rocks that contain both pyroxenes (Nimis and Grütter 2010). Although the Nimis and Taylor (2000) thermometer uses only one pyroxene to calculate T, its application expressly requires that both pyroxenes were part of the mineral assemblage and were in chemical equilibrium. Therefore, this method is only suitable for diamonds belonging to the lherzolitic or websteritic suites and containing clinopyroxene inclusions, either isolated or associated with other minerals. If orthopyroxene was not part of the original mineral assemblage, as is the case for wehrlitic inclusions, single-clinopyroxene thermometry would only provide a minimum T estimate. An anomalously low T estimate (e.g., one that lies much below the local geotherm) may be a sign that the clinopyroxene is wehrlitic. Ziberna et al. (2016) suggested that a Ca/(Ca + Mg) molar ratio of > 0.5 should also be considered as suspicious, as only ~1% of orthopyroxene-saturated mantle clinopyroxenes lie above this value. Nonetheless, discrimination of wehrlitic clinopyroxenes is not generally possible based merely on compositional criteria.Simakov (2008) proposed a different, more complex calibration of the single-clinopyroxene thermometer, which definitely improves performance above 1500 °C, though not over the T < 1400 °C range typical of ‘lithospheric’ temperatures (note that the TNimis and Taylor 2000 estimates reported in Fig. 10 of Simakov 2008 are incorrect and unduly suggest overestimation below 1300 °C).The Ca-in-Opx thermometer of Brey and Köhler (1990) is the single-orthopyroxene version of the pyroxene thermometer and can provide estimates that are complementary to, and independent from those obtained with the single-clinopyroxene thermometer. The two single-pyroxene methods were shown to be mutually consistent, provided a correction is applied below 900 °C (Nimis and Grütter 2010). However, since inclusions of likely harzburgitic (i.e., clinopyroxene-free) affinity are relatively common in diamonds (Stachel and Harris 2008), the possibility that TCa-in-Opx estimates for orthopyroxene inclusions are only minimum T values is much higher than for the single-clinopyroxene thermometer. Therefore, its usefulness is rather limited in diamond studies.Garnet. The Ni content of garnet in equilibrium with olivine is very sensitive to T and apparently independent of P (Ryan et al. 1996; Canil 1999). Since the Ni content of olivine shows small variations in both mantle xenoliths (O’Reilly et al. 1997) and inclusions in diamonds (Griffin et al. 1992; Sobolev et al. 2008), T can be retrieved from Nigarnet alone, by assuming an appropriate Ni content for coexisting forsteritic olivine. A useful reference values for Niolivine is the mean value for mantle olivines (mean ± standard deviation = 2900 ± 360 ppm; Ryan et al. 1996), which is also close to the average value for olivine inclusions in diamonds worldwide (Stachel and Harris 2008; Sobolev et al. 2009). If available, the mean Niolivine value for olivine inclusions in diamonds from the same locality may be used (e.g., 3150 ± 200 ppm for 51 inclusions from the Kalahari craton, Griffin et al. 1992; 2700 ppm for 88 inclusions from Arkhangel’sk province, Malkovets et al. 2011). Choosing one or the other value changes the final T estimate by at most a few tens of degrees (Fig. 1). The concentrations of Nigarnet can be measured with a laser-ablation inductively-coupled plasma mass spectrometer (LA-ICP-MS), an ion microprobe (SIMS) or a proton microprobe (PIXE), which typically ensure a precision within a few ppm. The assumption of equilibrium with olivine may require compositional filtering (e.g., Grütter et al. 2004) to exclude any non-peridotitic garnet.The Ni-in-garnet thermometer has seen wide application in studies of Cr-pyrope garnets included in diamonds (e.g., Griffin et al. 1992, 1993; Davies et al. 2004a; Viljoen et al. 2014; De Hoog et al. 2019). Its calibration, however, is somewhat controversial. Ryan et al. (1996) calibrated it against mantle xenoliths, using T values derived from a combination of the olivine–garnet Fe–Mg-exchange thermometer (O’Neill and Wood 1979) and the MacGregor (1974) and Brey and Köhler (1990) orthopyroxene–garnet barometers. Canil (1999) calibrated it against experiments at T ≥ 1200 °C. The two calibrations give identical results at ~1100 °C, but progressively diverge at lower and higher T (Fig. 1). As a result, estimates obtained through the Canil (1999) formulation will typically span over narrower T intervals. Considering the most typical range for Ni in garnet inclusions, the maximum difference TCanil–TRyan is ca. +100 °C at 20 ppm Ni and ca. –250 °C at 180 ppm Ni. It has been variously claimed that either calibration, or even their ‘average’, agree best with other independent thermometers when applied to mantle xenoliths (e.g., Ryan et al. 1996; Canil 1999; De Hoog et al. 2019; Czas et al. 2020; Nimis et al. 2020), but a definitive assessment using internally consistent thermobarometers as reference is still lacking. The latest attempt to refine the Ni-in-garnet thermometer was made by Sudholz et al. (2021a), who recalibrated it against new experiments in a relatively narrow T range (1100–1325 °C) and introduced correction terms for Ca and Cr contents in the garnet. When tested against independent estimates for xenoliths using the Nimis and Taylor (2000) enstatite-in-clinopyroxene thermometer, the Sudholz et al. (2021a) calibration shows improved overall accuracy above 1100 °C relative to the Canil (1999) experimental calibration, but poorer overall precision and slightly stronger progressive overestimation at lower T (see Fig. 7 in Sudholz et al. 2021a).The Mn content of garnet in equilibrium with mantle olivine is sensitive to T (Smith et al. 1991) and can be used as a single-mineral thermometer, assuming a constant Mn content in olivine. This thermometer relies on electron microprobe data and was proposed as a substitute for the Ni-in-garnet thermometer when trace element data are not available (Grütter et al. 1999; Creighton 2009). The declared precision is rather poor (mostly ±150 °C) and, in the absence of independent olivine data, severe underestimation may occur at high T. Therefore, this method may have some use in surveys of large populations of garnet xenocrysts recovered during diamond exploration (e.g., Grütter and Tuer 2009), but its utility for diamond thermobarometry is limited.Ashchepkov et al. (2010) calibrated two monomineral thermometers for garnets in equilibrium with clinopyroxene or olivine, respectively. These methods are simplified versions of Fe–Mg-exchange thermometers for clinopyroxene–garnet or olivine-garnet pairs, in which the clinopyroxene and olivine compositions are modeled from the composition of the garnet. Their accuracy cannot be better than that of the original two-mineral formulations, which are problematic themselves (see below). Also, P must be independently known to calculate T. In a test on mantle xenoliths from the Udachnaya kimberlite, these thermometers reproduced temperatures calculated with the pyroxene thermometer of Brey and Köhler (1990), at pressures calculated with the orthopyroxene–garnet barometer, to within ca. ±100 °C. Additional uncertainties due to problematic estimation of P in the absence of compositional data for coexisting orthopyroxene (see below) probably make these thermometers of little interest for diamond thermobarometry, especially when investigating small populations of diamonds.Olivine. The Al content in peridotitic olivine is sensitive to T and can be used as a thermometer. The first Al-in-olivine thermometer for peridotitic olivine at conditions relevant to diamond was empirically calibrated on mantle xenoliths by De Hoog et al. (2010), using P–T estimates derived by a combination of the two-pyroxene thermometer and orthopyroxene–garnet barometer of Brey and Köhler (1990). The method was recalibrated by Bussweiler et al. (2017) by analysis of olivine from high-P–T experimental charges. The Al content in cratonic mantle olivine ranges from ~1 ppm to ~250 ppm and is conveniently analyzed by techniques such as LA-ICP-MS or SIMS, which ensure a precision within a few ppm. Electron microprobe analysis (EMPA) using very high probe currents (e.g., ≥ 200 nA) and long count times (≥ 100 s) may be a valid alternative (Batanova et al. 2018; D’Souza et al. 2020).Al-in-olivine temperatures for spinel peridotites or strongly metasomatized garnet peridotites may be significantly underestimated or, respectively, overestimated. Compositional screening using an Al vs V diagram may help to discriminate these ‘unsafe’ olivines (Bussweiler et al. 2017). Screening is especially important if olivine is not physically associated with garnet, as occurs with many inclusions in diamonds. The Al-in-olivine thermometer gives systematically higher estimates (~50 °C, slightly increasing below 900 °C) than the single-clinopyroxene thermometer (Nimis and Taylor 2000) for natural mantle peridotites, probably reflecting slight suppression of Al incorporation in the experimental olivines due to Na loss (Bussweiler et al. 2017). The discrepancy is sufficiently small to ensure successful application of the method, but should be considered when comparing estimates obtained using different thermometers. The Al-in-olivine thermometer has not yet been widely used in diamond studies (see Korolev et al. 2018 for an example of application), owing to its recent development and the necessity of non-routine analysis of Al. Considering the great abundance of olivine among inclusions in diamonds (Stachel and Harris 2008), its application is likely to increase significantly in the future.Spinel. The Zn content of spinel in equilibrium with mantle olivine is sensitive to T. Since the Zn content of mantle olivine is nearly constant (mean ± standard deviation = 52 ± 14 ppm), the Zn content of spinel can directly be used as a thermometer (Ryan et al. 1996). The concentrations of Zn in spinel can be measured by SIMS or PIXE, which ensure a precision within a few ppm. The current version of the Zn-in-spinel thermometer was calibrated over the range 680–1180 °C against T for coexisting garnets using the Ryan et al. (1996) version of the Ni-in-garnet thermometer. Therefore, it is subject to at least the same uncertainties as the Ni-in-garnet thermometer (see above).Fe–Mg exchange thermometers. These popular thermometers are based on the distribution of Fe2+ and Mg between two mineral phases and can be used to derive T for inclusions in which garnet is associated with olivine, orthopyroxene or clinopyroxene. Tests using mineral compositions from mantle xenoliths and experiments showed that these thermometers suffer from either low precision or systematic errors or both (see review of Nimis and Grütter 2010). In peridotitic systems, the olivine-garnet and the clinopyroxene–garnet thermometers show the lowest precision, with possible discrepancies exceeding 200 °C. The Harley (1984) orthopyroxene–garnet formulation shows the highest precision, but systematically overestimates at T < 1100 °C and underestimates at T > 1100 °C (by up to ~150 °C, on average, for mantle xenoliths). All these discrepancies may reflect a modest sensitivity to T of the exchange reaction (particularly for olivine-garnet), oversimplification in the solid solution models used (especially for garnet and clinopyroxene), and Fe3+/ΣFe ratios in the minerals different from those in the calibration samples. Matjuschkin et al. (2014) experimentally demonstrated the large effect of neglected Fe3+ on orthopyroxene–garnet and olivine-garnet thermometers. Their experiments in the Na-free CaO–FeO–Fe2O3–MgO–Al2O3–SiO2 system at 5 GPa and 1000–1400 °C suggest that temperature estimates may be significantly improved if calculations are made using Fe2+ instead of total Fe in the garnet. Mössbauer data for orthopyroxene and garnet in peridotitic xenoliths, however, suggest that the partitioning of Fe3+ between these two minerals is affected by pressure and Na content in the orthopyroxene (Nimis et al. 2015). Therefore, a test over a range of relevant pressures in a Na-bearing system would be desirable. Nimis and Grütter (2010) empirically corrected the Harley (1984) thermometer using two-pyroxene thermometry of well equilibrated mantle xenoliths as calibration. The corrected version generally yields lower T estimates at low T and higher T estimates at high T relative to the uncorrected thermometer and may be more accurate for orthopyroxene–garnet pairs equilibrated under ‘average’ mantle redox conditions. On the other hand, it may produce large errors (> to ≫100 °C) if redox conditions are far from ‘average’, especially for strongly oxidized conditions at very high pressure and when both P and T are calculated by iteration in combination with an orthopyroxene–garnet barometer (Nimis et al. 2015). Considering that at deep lithospheric levels diamond is stable over a relatively wide range of redox conditions (∆logfO2 FMQ ≈ –5 to –2), this problem can be serious. Also, application to some xenolith suites tends to produce frequent high-P–T outliers (Grütter, pers. comm.). In spite of its recognized systematic discrepancies, the original Harley (1984) version may still be preferable for diamond thermobarometry due to its lower P dependence, which reduces error propagation during iterative P–T calculations. However, the systematic deviations of the Harley (1984) thermometer at T far from 1100 °C, especially if all Fe is treated as Fe2+, should be considered in data evaluation.The clinopyroxene–garnet thermometer represents the only viable method to determine T from major element compositions of eclogitic inclusions in diamonds. There are many versions of this thermometer in the literature. The most recent calibration (Nakamura 2009) is based on an expanded experimental database covering mafic and ultramafic compositions across the P and T ranges relevant to diamond (1.5–7.5 GPa, 800–1820 °C) and incorporates updated solid solution models for both clinopyroxene and garnet. Older, more simplified, but in some cases still very popular versions of this thermometer do not reproduce the same experiments equally well and often show systematic deviations with changing T and composition. The nominal uncertainty of the Nakamura (2009) thermometer (±74 °C at the 1-sigma level) is still about double that of the best performing thermometers for peridotitic inclusions. None of the available versions of the clinopyroxene–garnet thermometer include corrections for the content of jadeite in clinopyroxene, therefore application to clinopyroxenes with Na > 0.5 apfu is not recommended. This excludes 15% of reported compositions for inclusions in eclogitic diamonds. As regards the Fe3+ problem, Purwin et al. (2013) and Matjuschkin et al. (2014) measured Fe3+/ΣFe ratios of coexisting clinopyroxenes and garnets in a CaO–FeO–Fe2O3–MgO–Al2O3–SiO2 system at 2.5–5.0 GPa and 800–1400 °C and found no systematic deviations in the results of clinopyroxene–garnet Fe–Mg-exchange thermometry using the Krogh (1988) thermometer when all Fe was treated as Fe2+. However, since the system investigated was Na-free, it did not allow for the incorporation of Fe3+ in clinopyroxene as aegirine component. This mechanism was shown to be relevant in clinopyroxenes from garnet peridotites (Woodland 2009) and may be even more so in eclogitic clinopyroxenes. The effect of Na on Fe3+ uptake in clinopyroxene probably contributes to the lower precision of the clinopyroxene–garnet thermometer in applications to natural samples.Trace element-based clinopyroxene–garnet thermometers. The partitioning of REE between clinopyroxene and garnet is sensitive to T and has been used as a base for a clinopyroxene–garnet thermometer (Witt-Eickschen and O’Neill 2005; Sun and Liang 2015; Pickles et al. 2016). A possible advantage of REE thermometry over classical Fe–Mg-exchange thermometry is that REE diffusivity is small compared to that of Fe and Mg. Therefore, REE thermometry is predictably more robust against reequilibration of touching inclusions in diamonds during kimberlite transport or subsequent cooling. The version by Pickles et al. (2016) is particularly interesting, as it is independent of other thermometers and reproduces experimental temperatures moderately well (mean deviation of ±76 °C) with a small P dependence (~50 °C/1 GPa). This method is also independent of variations in Fe3+/ΣFe ratio and may thus pose a promising alternative to the traditional clinopyroxene–garnet Fe–Mg exchange thermometer. A practical disadvantage is the requirement of high-quality REE data, which cannot be obtained with routine electron microprobe analysis. To the author’s knowledge, there are no T data for diamonds using this method.Nitrogen-aggregation thermochronometry. Nitrogen is the most common impurity in natural diamonds. With time, nitrogen atoms in the diamond crystal lattice tend to aggregate, proceeding from initial C-centers (single nitrogen) in Type Ib diamond, to A-centers (nitrogen pairs) in Type IaA diamond and, finally, to B-centers (nitrogen in tetrahedral arrangement) in Type IaB diamond. The first transformation is completed in relatively short geological times (for a N content of 1000 ppm: ~100 yr at 1350 °C, and ~16 m.y at 950 °C; Taylor et al. 1996). Therefore, most natural diamonds are of the intermediate IaAB variety, the proportion of B-centers being a function of nitrogen content, temperature and time (Taylor et al. 1990; Mendelssohn and Milledge 1995). If the ages of the diamond and of its host kimberlite are known, one can calculate the integrated average temperature under which the diamond has resided in the mantle since its formation, i.e., the mantle residence temperature, Tres. Radiometric dating of inclusions within a particular diamond is rarely available. More commonly, the age of the diamond is assumed on the basis of radiometric dating of other diamonds from the same source. However, nitrogen aggregation is much more sensitive to temperature than to time and can be used as an effective thermometer, assuming the diamond has resided in the mantle at constant T since formation (Taylor et al. 1990, 1996). A diagram of nitrogen content vs the relative percentage of B-centers permits appreciation of age uncertainties on thermometry, since isotherms can be drawn for different mantle residence times (Fig. 2). Chronometry errors of one b.y. propagate thermometry errors of only a few tens of degrees on calculated Tres, unless the diamond is unusually young.Nitrogen content and aggregation can be measured by Fourier-Transform Infrared (FTIR) spectroscopy, using an infrared microscope in transmission mode. This method logically provides nitrogen content and aggregation data averaged over the volume of diamond analyzed. Working on thinned (ideally, a few hundred µm) samples at high spatial resolution may help to detect zoning and to obtain more robust estimates from distinct diamond growth zones. The reader is referred to Kohn et al. (2016) for practical guidelines on spatially resolved FTIR thermochronometry and for interpretation of data for zoned diamonds. Detection limits for nitrogen content in diamonds are typically on the order of 5–10 ppm for high-quality spectra. Nitrogen contents below detection limit (Type-II diamonds) are relatively rare in lithospheric diamonds, but are more common in sublithospheric diamonds and are clearly unsuitable for nitrogen-aggregation thermochronometry. It must be clear that, contrary to thermodynamic equilibria used in inclusion thermobarometry, nitrogen aggregation is a kinetic process and so does not provide a snapshot of temperature conditions at a certain time. For instance, a short period of elevated temperature, e.g., due to circulation of hot fluids in the mantle during or after diamond growth, may have a large effect on nitrogen aggregation and could lead to ambiguous interpretations of calculated Tres. In general, unless the diamond is very young relative to its kimberlite host, heating on a time scale of ~1 m.y. will have small effects on the calculated Tres, which will approach ambient mantle conditions rather than conditions during diamond formation (Fig. 3a). If the diamond experienced secular cooling over a b.y. time scale subsequent to formation, the calculated Tres may both significantly underestimate the formation T and overestimate the final ambient T (Fig. 3b).Orthopyroxene–garnet. The Al content of orthopyroxene in equilibrium with garnet is strongly sensitive to P and forms the basis of the most widely used barometer for garnet-bearing peridotites. In this barometer, P mostly depends on the Al content in the orthopyroxene, but the compositions of both minerals must be considered to obtain robust estimates. Of the many calibrations of this barometer, that of Brey and Köhler (1990) is by far the most popular, but the older Nickel and Green (1985) calibration reproduces experimental pressures for most peridotitic systems to somewhat better precision (Nimis and Grütter 2010). Still, the Nickel and Green (1985) version was formulated under simple model-system assumptions and requires correction of Al-component activity to avoid P underestimation for orthopyroxenes with non-negligible contents of jadeite component. Carswell and Gibb (1987) and Carswell (1991) proposed two simple correction schemes, based on different assumptions on how Ti is coupled with Na and Al in orthopyroxene. The two corrections only apply to orthopyroxenes with Na > (Cr + Fe3+ + 2·Ti) or, respectively, Na > (Cr + Fe3+ + Ti) atoms per formula unit, Fe3+ usually being neglected in practical applications. That of Carswell and Gibb (1987), which was also adopted by Brey and Köhler (1990) for their barometer, produces slightly smaller P variations, but there is as yet no solid experimental basis to prefer one correction over the other. The T-dependence of the orthopyroxene–garnet barometer is reasonably small (~0.2–0.3 GPa / 50 °C), but still sufficiently large to require an accurate independent estimate of T. This fact should be considered when this barometer is used in combination with the Harley (1984) orthopyroxene–garnet thermometer, which is known to overestimate at low T and underestimate at high T (see above): the artificial compression of T estimates produced by the Harley (1984) thermometer will necessarily result in non-natural compression also of P estimates. In general, errors in input T will displace the calculated P–T estimates roughly along a typical cratonic geotherm, thus limiting the ‘apparent’ scatter in P–T plots for samples equilibrated on the geotherm (Fig. 4). This may be an advantage if the aim is to define the mantle geotherm, but may give the misperception of a high overall precision by masking the true dispersion of the data.Clinopyroxene–garnet. Barometry of bimineralic eclogites and of eclogitic clinopyroxene–garnet inclusion pairs in diamonds has long been a problem for mantle scientists. There have been several attempts to develop a suitable barometer based on the P-sensitive equilibrium grossular + pyrope ↔ diopside + Ca-Tschermak. According to a recent test by Beyer et al. (2018), the Beyer et al. (2015) formulation reproduces pressures for experiments in natural systems significantly better than the earlier formulations of Simakov and Taylor (2000) and Simakov (2008). Nonetheless, in the Beyer et al. (2018) test on natural xenoliths from the Jericho kimberlite, Slave craton, P estimates for eclogites were systematically lower (by ~1 GPa) than orthopyroxene–garnet P estimates for peridotites recording similar T. As a result, the Jericho eclogites oddly appeared to fall on a hotter geotherm. The T-dependence of the clinopyroxene–garnet barometer is ca. 0.25 GPa / 50 °C, i.e., similar to that of the orthopyroxene–garnet barometer. Therefore, also in this case, errors in input T would displace P–T estimates roughly along the geotherm and cannot explain the observed discrepancy between peridotite-system and eclogite-system thermobarometry results.A further test of clinopyroxene–garnet thermobarometry is illustrated in Figure 5, for touching or non-touching inclusions in diamonds and diamond- or graphite-bearing xenoliths. The P–T estimates reflect iterations of the Beyer et al. (2015) barometer and the Nakamura (2009) Fe–Mg exchange thermometer. Since the uncertainty of the barometer increases rapidly with decreasing concentrations of [4]Al in clinopyroxene, application to clinopyroxenes with Si contents >1.985 atoms per 6-oxygen formula is not recommended by Beyer et al. (2015). These compositions are relatively common for inclusions in diamonds (35% of 224 inclusions) and diamondiferous xenoliths (42% of 233 xenoliths), thus numerous samples had to be excluded. Even if further severe filters were applied to discard potentially low-quality chemical analyses, the results are disconcerting: although graphite-bearing xenoliths plot in the graphite stability field, most diamond-bearing xenoliths and inclusions in diamonds yield too low pressures well outside the diamond stability field (Fig. 5). Iteration of earlier formulations of the same barometer and thermometer does not improve the results (cf. Fig. 10 in Shirey et al. 2013). The reason for this discrepancy is unclear. Simakov and Taylor (2000) cautioned that a barometer based on the Ca-Tschermak content in clinopyroxene may lead to incorrect estimates when applied to kyanite- or SiO2-oversaturated assemblages, but the Beyer et al. (2015) barometer shows no systematic deviations in calculated pressures for experiments in these systems. The large chemical complexity of eclogitic garnets and, especially, eclogitic clinopyroxenes might in part explain the poor performance of the barometer in applications to natural systems, in which variable amounts of Na, Fe3+, Cr, K and vacancy-bearing (i.e., Ca-Eskola) components occur. A further possible explanation is the high sensitivity of the barometer to even small errors in the chemical analyses of clinopyroxene. Beyer et al. (2015) provided a formula to predict relative errors on calculated pressures due to propagation of uncertainties on SiO2 contents, assuming an uncertainty of 1% relative in electron microprobe analysis: %err = 1.94 × 10–8e10.18398[Si apfu]. These model errors do not consider the additional errors that derive from iterative calculation of P and T (Fig. 5). A 1% relative increase in SiO2 contents may move many inaccurate P–T estimates into the diamond field, suggesting that application of the clinopyroxene–garnet barometer may require very high-quality, well-standardized electron microprobe analyses to yield meaningful results. Many of the analyses reported in the literature are possibly of routine quality and do not meet the necessary standards for robust clinopyroxene–garnet barometry. Still, analytical quality alone is unlikely to account for all of the observed discrepancies (Fig. 5).Sun and Liang (2015) proposed that REE partitioning between clinopyroxene and garnet could be used not only as a thermometer (see above) but also as a barometer. Unfortunately, the discrepancies between calculated PREE and run pressures for a set of independent validation experiments appear to be very large (up to ~3 GPa). A test conducted on twelve quartz eclogites, two graphite eclogites and nine diamond eclogites passed the appropriate stability-field constraints for all but two of the nine diamond eclogites; these fell at 0.2 and 0.5 GPa lower P than the diamond-graphite phase boundary. The proposed REE thermobarometry approach is promising, but further tests are necessary before this method can be used with confidence in diamond studies.If trace element data are available, the partitioning of Li between clinopyroxene and garnet can also be used as a barometer (Hanrahan et al. 2009). Pressures for fifteen calibration experiments on a Mg-rich eclogitic composition at 4–13 GPa and 1100–1500 °C are reproduced to a very reasonable ±0.2 GPa (1 sigma) and the T dependence of the barometer is small (~0.2 GPa / 50 °C). Preliminary tests yielded conditions compatible with diamond for three diamond-bearing xenoliths and for non-touching inclusion pairs in four diamonds, whereas two other diamonds yielded conditions up to 2 GPa lower than the diamond stability field. Hanrahan et al. (2009) suggested that the two spurious P estimates could reflect disequilibrium between the non-touching inclusions. However, the use of non-touching inclusions is essential for this barometer, as the high diffusivity of Li may lead to reequilibration of small touching inclusions during and after transport to surface. Similar to the REE barometer, further tests are needed before this method can be recommended for diamond studies.Monomineral barometers.Nimis and Taylor (2000) developed an empirical Cr-inclinopyroxene barometer for chromian diopsides in equilibrium with garnet. In combination with the Nimis and Taylor (2000) single-clinopyroxene thermometer, this barometer uniquely permits derivation of both P and T conditions based on the composition of a single mineral inclusion. The T-dependence of the Cr-in-clinopyroxene barometer (0.15–0.25 GPa / 50 °C for diamond inclusion compositions) is generally lower than that of the orthopyroxene–garnet barometer (~0.2–0.3 GPa / 50 °C) and the P-dependence of the complementary single-clinopyroxene thermometer is very small (Fig. 4). Therefore, if P and T are calculated by iteration, errors in input T or P will displace a clinopyroxene P–T estimate at an angle relative to a cratonic geotherm and will contribute to ‘apparent’ scatter in P–T results for samples equilibrated on a steady-state geotherm. The Nimis and Taylor (2000) Cr-in-clinopyroxene barometer tends to progressively underestimate above 4.5 GPa (by up to ~1 GPa at 7 GPa; Nimis 2002; Ziberna et al. 2016). This is a known artifact and the discrepancy should be considered when comparing Cr-in-clinopyroxene estimates with those obtained using other methods. Nimis et al. (2020) proposed an empirical correction to minimize the differences with P estimates obtained using the orthopyroxene–garnet barometer at high P, at the cost of somewhat lower precision. These authors cautioned that their correction represented a temporary measure to reduce inconsistencies between independent barometric estimates. More recently, Sudholz et al. (2021b) provided an experimental recalibration of the Cr-in-clinopyroxene barometer that yields results very similar to the empirically corrected calibration for clinopyroxenes associated with diamonds or hosted in well equilibrated peridotitic xenoliths at P > 5 GPa (Fig. 6a). At lower P the two revised calibrations diverge slightly, but the Nimis et al. (2020) calibration maintains somewhat better overall consistency with orthopyroxene–garnet barometry (Fig. 6b,c). For proper application of the Cr-in-clinopyroxene barometer, the analyses must undergo compositional filtering in order to select clinopyroxenes in equilibrium with garnet and eliminate certain compositions for which analytical uncertainties propagate excessive uncertainties on the calculated P. Ziberna et al. (2016) provided a useful cookbook to perform the necessary compositional screening, which slightly improved an earlier one proposed by Grütter (2009). According to this cookbook, some compositions require that electron microprobe analyses are carried out using beam currents and counting times higher than routine to minimize analytical uncertainties, whereas some unfavorable compositions should simply be discarded. In particular, the recommended limitation to Cr/(Cr+Al)mol > 0.1 excludes about two thirds of reported inclusions classified as websteritic. When applied to properly filtered graphite- or diamond-bearing xenoliths and inclusions in peridotitic diamonds, the Cr-in-clinopyroxene barometers yield results consistent with or within ~0.3 GPa of the stability field of the respective carbon phase (Nimis and Taylor 2000; see also Fig. 8a). The lack of adequate filtering in previously published thermobarometric results greatly contributed to the larger overall scatter in clinopyroxene P–T plots relative to those for orthopyroxene–garnet pairs (cf. Stachel and Harris 2008).Griffin and Ryan (1995) devised an empirical barometer based on the Cr content of garnet in equilibrium with orthopyroxene and spinel. This Cr-in-garnet barometer was later refined by Ryan et al. (1996) and further revised, using a different Cr/Ca-in-garnet approach, by Grütter et al. (2006). The last two versions are the most widely used, typically in combination with a Ni-in-garnet thermometer. That of Grütter et al. (2006) gives up to ~10% lower estimates and a better agreement with petrological constraints imposed by the graphite–diamond transition, and is therefore preferable. If the coexistence of garnet with spinel is unknown, as is the case for many garnet inclusions in diamonds, the Cr-in-garnet methods only provide an estimate of minimum pressure.Simakov (2008) proposed a simplified version of the clinopyroxene–garnet barometer for eclogites that solely relies on the Ca-Tschermak content in the clinopyroxene. The accuracy of this monomineral barometer cannot be better than that of the native two-phase barometer and is thus prone to the same problematic issues (see above). Ashchepkov et al. (2017) calibrated or recalibrated a series of empirical monomineral thermobarometers, which should be suitable for peridotitic and eclogitic garnets and clinopyroxenes. When applied to mantle xenoliths and diamond inclusions from individual localities, these methods produce strongly scattered P–T estimates (e.g., Figs. 9–12 in Ashchepkov et al. 2017), which are difficult to compare with the more regular P–T distributions that are typically obtained with other, more widely tested methods. The usefulness of the Ashchepkov et al. (2017) monomineral formulations for diamond thermobarometry is therefore questionable.With increasing pressure beyond ~7 GPa, mantle garnets incorporate increasing proportions of majoritic (i.e., pyroxene-like) component by substitution of Si for octahedrally coordinated Al and Cr. The Na and Ti contents also tend to increase. Collerson et al. (2000) utilized the relationship observed between majorite fractions and pressure in experiments on mafic and ultramafic systems to devise an empirical single-mineral barometer. The method was later revised by Collerson et al. (2010), Wijbrans et al. (2016) and Beyer and Frost (2017). Only Wijbrans et al. (2016) proposed two distinct calibrations, one for peridotitic and another for eclogitic systems. Experimental tests by Beyer and Frost (2017) and Beyer et al. (2018) showed systematic deviations or decreased precision for all but the Beyer and Frost (2017) formulation. This was particularly relevant for eclogitic compositions, which were underrepresented in previous calibration datasets. The Beyer and Frost (2017) barometer was calibrated at pressures of 6 to 20 GPa and temperatures of 900 to 2100 °C. These conditions cover sublithospheric mantle conditions up to ~575 km depth, corresponding to the lower transition zone. Experimental pressures are reproduced with a standard error of estimate of 0.86 GPa. In relative terms, considering the enormous range of pressures covered by the calibration, this uncertainty is not much larger than that of the barometers used for lithospheric diamonds. The effect of temperature on pressure estimates is apparently negligible. Note that application of the majorite barometer to lithospheric garnets equilibrated at P < 6 GPa is not recommended, because under such conditions the barometer becomes extremely sensitive to analytical errors (Beyer and Frost 2017). Tao et al. (2018) showed that high Fe3+ contents in the garnet tend to produce underestimated pressures with previous majorite barometers and provided a further revised calibration of the Collerson et al. (2010) barometer that explicitly takes into account Fe3+. The Tao et al. (2018) calibration may be preferable when Fe3+ data are available and measured Fe3+/ΣFe ratios are greater than 0.2. All these majorite barometers require that garnet occurs in equilibrium with a clinopyroxene phase. This may be an issue, because with increasing pressure pyroxene progressively dissolves into garnet and at pressures above ~15–19 GPa, depending on bulk composition and temperature, mantle rocks no longer contain pyroxene (Irifune et al. 1986; Irifune 1987; Okamoto and Maruyama 2004). Pyroxene-undersaturated condition can be suggested by the presence of CaSiO3 phases along with majoritic garnet as inclusions in diamonds or by high Ca contents in majoritic garnet. Under such conditions, the barometer would give erroneously low pressures (Harte and Cayzer 2007). If there is no independent evidence that the garnet was pyroxene-saturated, any pressure estimates close to that of the clinopyroxene-out reaction should be considered as estimates of the minimum P of entrapment. Thomson et al. (2021) devised a novel type of majorite barometer using machine-learning algorithms. Compared to traditional majorite barometers, this new barometer is apparently insensitive to petrological limitations (most notably, the absence of clinopyroxene) and reproduces experimental majoritic garnet compositions over a wider range of bulk compositions and experimental pressures (6–25 GPa) with much better overall precision (mostly within ±2 GPa). Thomson et al. (2021) caution that the compositions of majoritic garnet inclusions in diamonds lie in a region with relatively fewer experiments and machine-learning regressions may not be reliable in extrapolation. Thus, pressure predictions for inclusions in diamonds may have somewhat larger uncertainties. Still, this is the only available barometric method that allows estimating pressures for majoritic garnet inclusions beyond the clinopyroxene-out reaction. When applied to inclusions in diamonds, differences between pressure estimates obtained with the Thomson et al. (2021) and Beyer and Frost (2017) barometers lie in the range –2.5 to +6 GPa. Whatever the majorite barometric method used, if the garnet reequilibrated with clinopyroxene after entrapment, as is typically the case in unmixed inclusions (see below), underestimation of P will occur.Elastic methods. When an inclusion is entrapped in a diamond, both the inclusion and the diamond are initially at the same pressure. On eruption, the pressure acting on the diamond drops to 1 atm, but a residual pressure, Pinc, of up to several GPa may develop on the inclusion as a consequence of the inclusion and host having different elastic properties. If we know the Pinc and the elastic properties of the inclusion and host, we can back-calculate the conditions under which the two minerals would be at the same pressure. These conditions describe a line in P–T space, known as the entrapment isomeke, and the entrapment conditions (i.e., the diamond formation P–T) will be one of the points along the isomeke. This forms the basis for elastic barometry of diamonds. The principles, problems and on-going developments of this method are described in detail in Angel et al. (2022, this volume).Projection onto a geotherm. When only thermometric estimates are available due to a lack of suitable barometers, a tentative estimate of pressure can be obtained by projecting temperature estimates on a reference geotherm. This will most conveniently be a geotherm derived by fitting P–T data for mantle xenoliths or xenocrysts from the same kimberlitic source. The FITPLOT computational software by Mather et al. (2011) provides a useful aid to model geotherms from xenolith P–T data. There may be significant uncertainties in some of the input parameters that are used for the modeling, specifically the thickness and heat production of the upper and lower crust and the mantle potential temperature, but in most cases these uncertainties will not significantly affect the shape of the geotherm within the lithospheric diamond window, which will be chiefly constrained by the xenolith P–T data. Nonetheless, assuming different mantle potential temperatures will affect the estimated thickness of the lithosphere and the depth extent of the conductive portion of the geotherm. This may have some relevance for some old, high-T diamonds, as mantle potential temperatures may have varied over geological time. If this variation is not considered, thermometric estimates may unduly be projected onto the adiabatic portion of the geotherm and lead to severe overestimation of P. Values for the present-day mantle potential temperature and its change with time can be found in Katsura et al. (2010) and Ganne and Feng (2017). The output data that describe the relevant section of the model geotherm can conveniently be fitted through a polynomial expression to derive P as a function of T. The final P–T estimate will be given by the intersection of this polynomial with the isotherm described by the chosen thermometer. Note that this isotherm will not generally be a flat line in a P–T plot, due to the generally significant P-dependence of the thermometers (Fig. 4).The basic assumption underlying this approach is that the inclusions last equilibrated at P–T conditions lying on the mantle geotherm at the time of eruption. This is generally acceptable for touching inclusion pairs, which can reequilibrate in mostly the same way as touching minerals in xenoliths (but see caveat in section below dedicated to touching vs. non-touching inclusions), whereas it is much less obviously so for non-touching inclusions. Comparisons between P–T estimates for inclusions and xenoliths from the same sources showed that the assumption often holds also for non-touching inclusions (Nimis 2002). Nonetheless, there is also evidence of some diamond inclusions recording conditions significantly hotter or colder than the xenolith geotherm (Griffin et al. 1993; Sobolev and Yefimova 1998; Nimis 2002; Stachel et al. 2003, 2004; Weiss et al. 2018; Nimis et al. 2020). Also, multiple non-touching inclusions within individual diamonds may record a range of conditions, indicating thermal fluctuations during the growth history of the diamond (Griffin et al. 1993). Projection onto a geotherm should therefore be used with caution and relatively large uncertainties should be allowed. Assuming a maximum possible T difference from a xenolith geotherm of ~250 °C, consistent with available estimates for diamonds using single-clinopyroxene (Nimis and Taylor 2000), single-garnet (Canil 1999) and orthopyroxene–garnet (Harley 1984) thermometers, projection on a typical cratonic geotherm may result in a P uncertainty of up to ~1.5 GPa (Fig. 4).Projection onto the adiabatic portion of the mantle geotherm may provide a necessary independent constraint for the application of elastic barometric methods to sublithospheric diamonds (e.g., Anzolini et al. 2019).Mineral stability. Lithospheric mantle rocks provide little opportunity to estimate conditions of diamond formation based only on the stability of mineral assemblages. For example, across the entire diamond window, the only significant mineralogical change is the spinel-to-garnet transition in peridotite. The spinel-to-garnet reaction is not univariant and in depleted cratonic peridotites with strongly elevated Cr/Al, spinel + garnet assemblages may persist over a large range of P–T conditions (e.g., MacGregor 1970; Webb and Wood 1986; Klemme 2004; Ziberna et al. 2013). In general, spinel is stabilized to higher pressures in more refractory, Cr-rich and Al-poor bulk compositions. This allows magnesiochromite to be incorporated as inclusions in diamonds that were formed in very deep, refractory environments (e.g., 6.5 GPa, corresponding to a depth of over 200 km, for a diamond from the Udachnaya kimberlite, Siberia; Nestola et al. 2019a).Mineral stabilities become more interesting in the sublithospheric mantle, where a number of characteristic mineralogical changes occur (Harte 2010; Harte and Hudson 2013). Although these changes do not permit bracketing of P–T conditions with a resolution comparable to that of conventional thermobarometers, in some cases they constrain the formation of a diamond to within specific mantle depth regions. For instance, the coexistence of Fe-bearing periclase with enstatite, interpreted to be inverted bridgmanite (previously known as MgSi-perovskite), has long provided evidence that some diamonds were formed at lower-mantle depths (Scott Smith et al. 1984; Moore et al. 1986). Even in the absence of periclase, inverted bridgmanite from the lower mantle can readily be distinguished from upper-mantle enstatite by its very low Ni contents (Stachel et al. 2000). The finding of a rare inclusion of ringwoodite in a diamond from Juina, Brazil, unequivocally demonstrated its origin from the mantle transition zone (Pearson et al. 2014). Also, the reconstructed compositions of a suite of exsolved inclusions in Juina diamonds nicely matched those of minerals expected to form in a basaltic system in the lower mantle (Walter et al. 2011).However, using incomplete mineral assemblages found as inclusions in diamonds as depth markers is not always free from ambiguity. For instance, inclusions of breyite (previously known as CaSi-walstromite) have long been considered to represent retrogressed CaSiO3-perovskite from depths greater than ~600 km (e.g., Harte et al. 1999; Joswig et al. 1999), but there is compelling evidence that at least some breyite inclusions originated at much shallower depths within the upper mantle (Brenker et al. 2005; Anzolini et al. 2016). Woodland et al. (2020) recently provided experimental support for a possible upper-mantle origin of breyite. Also, Fe-bearing periclase is the most common inclusion of interpreted lower-mantle origin (Kaminsky 2012), but there is evidence that Fe-bearing periclase participates in mineral parageneses straddling the upper mantle-lower mantle boundary (Hutchison et al. 2001) and that it may be a co-product of diamond-forming reactions that may occur in the transition zone and even in the overlying upper mantle (Thomson et al. 2016; Bulatov et al. 2019).Data quality. A sometimes underrated prerequisite for robust thermobarometry is precise and accurate chemical analyses. Inclusions in diamonds are often small and their chemical analysis can be challenging. Moreover, some thermobarometers may be particularly sensitive to analytical uncertainties on elements that occur at critical concentration levels. For instance, Ziberna et al. (2016) explored propagation of analytical errors on P estimates using the single-clinopyroxene barometer of Nimis and Taylor (2000) and found that routine electron microprobe analytical conditions may be insufficient for meaningful barometry of many inclusions in diamonds. Clearly, not only poor counting statistics but also bad sample preparation and improper standardization can be an issue and may lead to spurious results. A survey of published chemical data for over 600 clinopyroxene inclusions in lithospheric diamonds reveals that ~30% of them have oxide total weight percentages < 98.5% or > 101%, or cation sums < 3.98 or > 4.03 atoms per formula unit and, thus, are definitely not of good quality! Note that the above cut-offs generously take into account the possible effects of considering all Fe as Fe2+ in electron microprobe analyses.The next critical step when studying inclusions that were separated from coexisting mineral phases is the definition of their original paragenesis. Even if single-mineral thermobarometers may often come in handy, they invariably assume that the investigated mineral last equilibrated with specific other minerals. As for clinopyroxene and garnet, which are major constituents in several types of mantle rocks, simple compositional screening based on either major or trace element concentrations may help to discriminate between the most typical paragenetic varieties (e.g., Griffin et al. 2002; Grütter et al. 2004; Nimis et al. 2009; Ziberna et al. 2016). Unusual compositions that are not sensitively discriminated by available classification schemes or fall outside the compositional range over which the thermobarometers were calibrated, as well as false positives that unduly survive compositional filtering, may undermine the reliability of thermobarometric data. This problem may be minimized by investigating large datasets, when this is permitted by the number and nature of the available samples.Even when all cautionary steps have been made to select suitable compositional data, thermobarometric estimates are subject to errors. As discussed in the previous section, some of these errors are systematic and lead to predictable inconsistencies between estimates obtained using different methods. This fact should always be considered when comparing data obtained using different thermobarometers or even different formulations of the same thermobarometer. Thermobarometric methods with demonstrated internal consistency must be prioritized if the purpose is to explore the distribution of data for heterogeneous inclusion populations.In this respect, thermobarometry of eclogitic diamonds remains problematic. Considering the compositional limitations required by the ‘best’ available barometric methods (i.e., SiCpx < 1.985 apfu and NaCpx < 0.5 apfu), ~45% of over 200 reported eclogitic clinopyroxene–garnet inclusion pairs are automatically excluded. In addition, the reliability of barometric methods for eclogitic inclusions is questionable (see above).When P estimates are too uncertain or not available, for example, due to a lack of suitable barometers for the specific inclusions investigated, projection of temperatures onto the local xenolith geotherm can provide a last-resort solution to estimate P. The resulting uncertainties can be too large for an accurate characterization of genesis conditions for individual diamonds, but the procedure can have some utility for exploring general trends in large datasets (e.g., Korolev et al. 2018; Nimis et al. 2020). It probably remains the safest available method for barometry of eclogitic inclusions, notwithstanding potential over- or underestimation for diamonds formed at conditions hotter or, respectively, colder than ambient mantle at the time of eruption. A fairly popular alternate has been to calculate T for these diamonds at a fixed P of 5 GPa, approximately corresponding to the average P for diamonds worldwide. This method is not as effective and may overcompress the range of T estimates, due to the generally significant P-dependence of the thermometers (Fig. 4).Touching vs. non-touching. When a pair of touching mineral grains included in a diamond reequilibrate to changing external conditions, their total encapsulated volume remains constrained by the surrounding diamond. This forces the included minerals to follow a distinct pressure path and to achieve final compositions that are different from those of the same minerals outside the diamond (see Angel et al. 2015 and Ferrero and Angel 2018 for more general discussion of this phenomenon). The most common changes in ambient conditions are probably isobaric cooling or heating after diamond growth, in response to secular variations in mantle thermal state or following short-term thermal pulses related to diamond-forming processes. In most cases of diamond-hosted inclusions, cooling will cause a decrease in the inclusion P even if the external P remains constant, leading to underestimation of the actual diamond P using chemical thermobarometers. The extent of this underestimation can be calculated if the thermoelastic properties of the inclusion and host minerals are known (Angel et al. 2015). For various combinations of garnet, pyroxene and olivine inclusions in diamond, the error would amount to ~0.3 GPa for a temperature drop of 100 °C. Heating will have the opposite effect. These errors are of a similar magnitude as those of the orthopyroxene–garnet and single-clinopyroxene barometers and may thus contribute to some of the scatter in P–T plots for touching inclusions. The resulting errors may be lower, however, if plastic deformation in the diamond is able to reduce the under- or overpressures developed on the inclusions.If some thermal reequilibration occurs, the use of single-mineral thermobarometers on touching inclusions may generate further errors when some key minerals are not part of the inclusion assemblage. For instance, touching inclusions of clinopyroxene-orthopyroxene devoid of garnet will reliably record the ambient thermal state at the time of eruption, but will be poor candidates for single-clinopyroxene barometry. In response to a thermal change, the clinopyroxene will easily reequilibrate with orthopyroxene, but not with garnet, and so a false apparent encapsulation P will be calculated at the final equilibrium T (Nimis 2002). Touching inclusions of clinopyroxene ± garnet ± olivine would instead produce estimates close to the diamond formation conditions, because the composition of the clinopyroxene in orthopyroxene-free assemblages would be little affected by a variation in temperature. Polymineralic inclusions consisting of touching clinopyroxene-orthopyroxene–garnet ± olivine will fully reequilibrate and will provide P and T conditions close to those at the time of eruption (except for the small systematic errors described in the previous paragraph).Non-touching inclusions are immune to the above problems, because the inert diamond protects them from post-entrapment reequilibration, and are unequivocally the best candidates to determine the diamond formation P–T, provided reliable monomineral thermobarometers can be used. In all other cases, non-touching inclusion pairs are prone to potential errors, because the individual inclusions may have been incorporated at different times and under different conditions (see above) and may thus not be in mutual equilibrium. The largest T difference between multiple inclusions in individual diamonds was reported for garnets in a diamond from the Mir kimberlite, Siberia (Griffin et al. 1993). The calculated difference varies greatly depending on which calibration of the Ni-in-garnet thermometer is used, i.e., ca. 400 °C with the Ryan et al. (1996) calibration or ca. 230 °C with the Canil (1999) calibration. Disequilibrium amongst non-touching inclusions in diamonds, however, appears to be the exception rather than the rule, as multiple inclusions in individual diamonds most often yield similar P–T estimates once the systematic deviations between different thermobarometers are taken into account (Stachel and Harris 2008; Stachel and Luth 2015).Syngenesis vs. protogenesis. For several decades, thermobarometry of inclusions in diamonds has relied on the assumption of syngenesis, i.e., the inclusions formed or completely recrystallized during the growth of diamond. A corollary of this assumption was that the inclusions record the conditions of diamond formation, excepting for potential post-entrapment modifications in touching inclusions. The main proof for syngenesis was that diamond typically imposed its shape on the inclusions. This widely accepted paradigm was challenged by Nestola et al. (2014), who found evidence of protogenesis (i.e., formation before the diamond) for some olivine inclusions that would have been classified as syngenetic based on morphological criteria. Similar evidence was then found for other olivine, clinopyroxene, garnet and magnesiochromite inclusions (Milani et al. 2016; Nestola et al. 2017, 2019b; Nimis et al. 2019). The possibility that inclusions in diamonds are protogenetic suggests that they might record conditions predating diamond formation. This might be particularly relevant for isolated inclusions, if their host diamonds were formed from hot fluids or melts intruding through relatively cool mantle rocks. Did the preexisting mineral grains have enough time to fully reequilibrate to the changing thermal conditions during diamond growth and before their incorporation was complete?The time required by a mineral to adjust its composition to new physicochemical conditions strongly depends on temperature and grain size, and can be estimated from diffusion kinetics. For instance, a clinopyroxene grain with a diameter of 200 µm would fully reset its Ca/Mg ratio (the main temperature sensor in pyroxene thermometry) in less than 100,000 yr above 1200 °C and in a few m.y. below 900 °C (Fig. 7). Whether these time spans are short or long enough for complete reequilibration depends on the diamond growth rate, which is not precisely known. It is tempting to infer that incomplete reequilibration of protogenetic inclusions, especially the largest and shallowest (i.e., coldest) ones, may account for some of the scatter in P–T plots for non-touching inclusions (Fig. 8). Also, the similarity of P–T conditions occasionally observed for inclusions in diamonds and xenoliths from the same sources may alternatively be ascribed to diamond formation from thermally-equilibrated media (Nimis 2002) or to lack of reequilibration of protogenetic inclusions during short-lived diamond-forming processes (Nestola et al. 2014).Exsolved inclusions. Some inclusions in diamonds show evidence of post-entrapment unmixing. Reported examples are clinopyroxenes with orthopyroxene exsolution lamellae (± coesite), representing unmixed high-T clinopyroxenes, and intergrowths of more or less majoritic garnet and clinopyroxene, representing unmixed high-P majoritic garnets.Exsolved clinopyroxene inclusions are uncommon and have only been reported in diamonds from Namibian placers and from the Voorspoed kimberlite, South Africa (Leost et al. 2003; Viljoen et al. 2018). The coexistence of clinopyroxene and orthopyroxene makes these inclusions ideal for pyroxene thermometry, but the exsolved assemblage does not provide barometric estimates. Since pyroxene thermometers have a small P-dependence, projection onto the local xenolith geotherm provides in these cases reasonable estimates of the depth of origin of the diamonds (Nimis et al. 2020). If the compositions and relative proportions of the exsolved phases are known (e.g., by combining electron microprobe data with 2D or, better, 3D image analyses), the integrated pre-exsolution composition of the clinopyroxene can be reconstructed. The integrated composition can provide a T estimate at the time of diamond growth, assuming equilibrium with a separate orthopyroxene during entrapment, or a minimum T estimate of diamond formation, in the absence of orthopyroxene. In the reported cases, thermometry of the reconstructed clinopyroxenes provided evidence for diamond formation at temperatures near the mantle adiabat and ~200 °C hotter than the ambient mantle at the time of eruption. Attempts to apply the single-clinopyroxene barometer to the reconstructed compositions yielded unreasonably low pressure estimates, suggesting formation in strongly modified mantle environments (Nimis et al. 2020).Exsolution textures are common in majoritic garnet inclusions in sublithospheric diamonds (Moore and Gurney 1989; Wilding 1990; Harte and Cayzer 2007). These textures are interpreted to reflect decompression during slow ascent of the diamonds in convecting mantle or mantle plumes. The presence of exsolutions implies that pressure estimates based on majoritic garnet compositions are minima. Using reconstructed pre-exsolution garnet compositions will yield conditions closer to those of diamond formation. Even so, the resulting estimates may be minima, because the garnet may have originated at depths beyond the clinopyroxene-out reaction (see previous Barometry section).Figure 8 shows a compilation of P–T data for lithospheric diamonds from worldwide sources, obtained using the most robust available thermobarometer combinations. The compilation comprises clinopyroxene inclusions in 100 lherzolitic diamonds and orthopyroxene–garnet inclusion pairs in 77 harzburgitic, 12 lherzolitic and 10 websteritic diamonds. The great majority of the clinopyroxene inclusions were isolated within their host diamonds and should generally record conditions close to those of diamond formation. The orthopyroxene–garnet pairs consist of both touching and non-touching assemblages. The touching pairs may have reequilibrated after their incorporation and should provide conditions close to the final conditions of residence in the mantle. In addition, P–T data were calculated, using the same methods, for diamondiferous xenoliths for which suitable mineral analyses were available.In comparing the two plots for single-clinopyroxene and orthopyroxene–garnet thermobarometry (Fig. 8a,b), it is important to consider the known systematic inconsistencies between the thermobarometric methods used (see above). The orthopyroxene–garnet thermometer tends to overestimate temperature below 1100 °C and to underestimate above 1100 °C. This accounts at least in part for the more restricted temperature range obtained for the orthopyroxene–garnet inclusions. Nonetheless, disregarding a single obvious outlier, all P–T values fall within 0.3 GPa of the diamond stability field, confirming the general reliability of the thermobarometers. Most of the P–T results lie between the 35- and 40-mW/m2 conductive geotherms of Hasterok and Chapman (2011), i.e., at conditions typical for cratonic lithospheres. Most exceptions record higher temperature conditions, which in the case of clinopyroxene may reach the mantle adiabat. All of these exceptions consist of non-touching inclusions. A possible explanation is that several diamonds were formed from thermally non-equilibrated media derived from the sublithospheric convective mantle. Another potential source of mantle temperature inhomogeneity during diamond formation can be incomplete cratonization at the time of diamond growth. For instance, in the Kaapvaal craton a stable geothermal gradient was established by about 2.5 Ga and followed by secular cooling (e.g., Michaut and Jaupart 2007; Brey and Shu 2018) and re-heating (Griffin et al. 2003), but many Kaapvaal diamonds are significantly older than 2.5 Ga (see review in Shirey and Richardson 2011) or may have formed during important perturbations of the local lithosphere (Griffin et al. 2003; Korolev et al. 2018). Therefore, it is not surprising that many of these diamonds may record conditions hotter or colder than those at the time of kimberlite eruption (e.g., Nimis et al. 2020).The touching orthopyroxene–garnet inclusion pairs deserve further discussion. In the available dataset, the majority of them (23 out of 31) refer to diamonds from the De Beers Pool (Kimberley, South Africa). These inclusions were studied by Phillips et al. (2004), who carefully documented differences between P–T estimates recorded by touching vs. non-touching inclusions. These workers found that the non-touching inclusions yielded higher P–T values than the touching inclusions and ascribed this observation to post-entrapment reequilibration of the touching inclusions under decreasing T conditions, possibly accompanied by mantle uplift and decompression. They also correctly noted that post-entrapment elastic modifications were not of sufficient magnitude to reconcile the observed P–T differences between touching and non-touching inclusions. However, Nimis et al. (2020) observed that both touching and non-touching inclusions in the Phillips et al. (2004) dataset record a colder geothermal regime than mantle xenoliths and xenocrysts in the same kimberlites, casting doubts on the secular cooling hypothesis. Whereas the relatively cold signature of the non-touching inclusions might reflect old conditions pre-dating the final thermal equilibration of the lithosphere recorded by the xenoliths, the data for the touching inclusions are difficult to interpret. Weiss et al. (2018) speculated that the low temperatures recorded by the touching inclusions could be related to late infiltration of cold, slab-derived fluids, remnants of which were preserved in some cloudy and coated diamonds from the same locality. Yet it remains unclear why infiltration of these fluids should leave virtually no traces in the sampled mantle xenoliths.Touching orthopyroxene–garnet inclusions from other localities are probably too few to be statistically significant. Nonetheless, in all these cases the touching inclusions yield temperatures significantly lower than non-touching inclusions in other diamonds from about the same depth (Fig. 8b). Additional thermobarometric evidence that diamond formation may be followed by cooling was provided by Stachel and Luth (2015) and Viljoen et al. (2018). However, Stachel and Luth (2015) also described touching inclusions yielding temperatures similar to or even slightly higher (up to 40 °C) than non-touching inclusions in the same diamonds.Diamond’s depth systematics. Irrespective of the nature of the inclusions and the thermobarometric methods used, the frequency distribution of pressures for lithospheric diamond is essentially unimodal, with a mode at ~6 GPa, corresponding to depths around ~190 km (Fig. 8; see also Stachel and Harris 2008 and Stachel 2014 for broadly similar distributions obtained using different thermobarometers and partly different datasets). The slight differences between the pressure distributions obtained for clinopyroxene and orthopyroxene–garnet inclusions may largely be ascribed to recognized inconsistencies between the different thermobarometer combinations used and, perhaps, to sampling bias. Therefore, these differences are probably not significant. The hard constraints imposed by the graphite–diamond and lithosphere-asthenosphere boundaries cannot fully explain the observed depth distributions, which show decreased diamond frequencies approaching these boundaries (Fig. 8). Nimis et al. (2020) obtained very similar depth distributions with depth modes at 180 ± 10 km for diamonds from three individual South African kimberlite sources by using various combinations of thermobarometric methods on different inclusion types. They also found no clear correlation with the depth distributions of mantle xenocrysts from the same sources calculated using the same thermobarometric methods. They concluded that the observed diamond depth distributions are unrelated to the kimberlite sampling efficiency and, thus, have genetic meaning and likely global significance.The global significance of the observed depth distributions is also supported by the abundant TNi-in-garnet data for garnet inclusions (Fig. 9). These data show a mode at ~1200 °C, but no pressure constraints are generally available. Assuming that the garnets last equilibrated at ‘average’ conditions between the 35- and 40- mW/m2 model geotherms, similar to the clinopyroxene and orthopyroxene–garnet inclusions (Fig. 8), the Ni-in-garnet temperature mode would correspond, again, to a pressure mode of ~6 GPa and a depth mode of ~190 km. The gentler increase of frequencies to ~1200 °C and their steeper decrease at higher temperatures (Fig. 9) also mirror the frequency variations with increasing pressure to ~6 GPa and beyond exhibited by the clinopyroxene and orthopyroxene–garnet inclusions (Fig. 8).The diamond depth distributions in the lithosphere may be the result of both constructive and destructive processes (Nimis et al. 2020). In particular, the progressive decrease in diamond concentrations at depths shallower than ~170–190 km may reflect (i) decreasing precipitation rates for diamond from ascending parental media or (ii) progressive decrease in the amount of these deep-sourced media at shallower levels. Probably, both mechanisms concur to limit the diamond endowment in the shallow portion of the diamond window. The rapid decrease of diamonds near the base of the lithosphere may reflect (i) a positive balance between the opposed effects that an increase of P and an increase of T may have on carbon solubility in some mantle fluids/melts along specific P–T paths or (ii) a progressive decrease in diamond endowment due to reactions with carbon-undersaturated asthenospheric fluids/melts. These two alternative hypotheses are difficult to prove or disprove, because our knowledge of the behavior of natural carbon-bearing melts and fluids at high P–T is still limited. The prevalence of melt-metasomatized lithologies near the base of typical lithospheres coincides with diminished diamond frequency and suggests that melt-driven metasomatism may be an important destructive agent for diamond (e.g., McCammon et al. 2001). Remobilization of carbon from these deep lithospheric roots may eventually contribute to build up the diamond concentration at shallower depths (Nimis et al. 2020).The recognition that depths around ~190 km are particularly favorable for diamonds may have interesting implications for their mineralogical exploration. Areas characterized by the occurrence of xenocrystic mantle minerals (e.g., clinopyroxenes or garnets) from these particular depths may be considered as high-priority targets. Nonetheless, since the xenocrysts and diamond depth distributions may be uncorrelated, even limited xenocryst records from the diamond’s most favorable depth range may be indicative of significant diamond potential (Nimis et al. 2020).Figure 10 shows the global distribution of pressure estimates for diamonds containing majoritic garnets. At present, these are the only types of inclusions that provide abundant, quantitative barometric data for sublithospheric diamonds. A difficulty in investigating these data lies in the fact that they may represent minimum P estimates. As explained in the previous section, this limitation is due to the frequent occurrence of exsolutions and, for all but the Thomson et al. (2021) barometer, to the possible absence of accompanying clinopyroxene in the original mineral assemblage from which the majoritic garnet inclusion was separated on entrapment. If one considers only inclusions that are reportedly free of clinopyroxene exsolutions, two distinct unimodal distributions are obtained for peridotitic and eclogitic/pyroxenitic garnets, respectively (Fig. 10). Using the Beyer and Frost (2017) barometer, the peridotitic inclusions are mostly clustered in the 175–225 and 225–275 km depth intervals, near the lithosphere-asthenosphere boundary. Most of these inclusions would thus more appropriately be classified as deep lithospheric. The eclogitic/pyroxenitic inclusions are, on average, much deeper and mostly fall in the depth interval between 300 and 500 km. If one considers also inclusions that do contain exsolved pyroxenes, the peak of eclogitic/pyroxenitic inclusions at ~400 km becomes more prominent. Using the Thomson et al. (2021) barometer, the distribution for the eclogitic/pyroxenitic inclusions is stretched to slightly higher pressures, with a maximum estimated depth of 613 ± 44 km, whereas many of the peridotitic inclusions are shifted to significantly greater depths, falling well into the sublithospheric region (Fig. 10). Some of the deepest (>450 km) majoritic garnets fall in the region where clinopyroxene disappears and thus the Beyer and Frost (2017) barometer may underestimate pressure for these samples. However, the Thomson et al. (2021) machine-learning barometer, which is immune to this effect, does not suggest a much deeper origin for these garnets (Fig. 10). Differences between depth estimates for these very deep samples using the two alternative barometers are between –65 and +81 km (average +17 km).It should be considered that, in some cases, in the absence of accurate imaging data, the occurrence of fine clinopyroxene exsolutions may have been overlooked. The use of improved BSE and EBSD imaging might allow recognition of exsolved clinopyroxene in nominally ‘clinopyroxene-free’ majoritic inclusions (cf. Harte and Cayzer 2007; Zedgenizov et al. 2014). The thirteen Juina majoritic inclusions for which no coexisting clinopyroxene has been reported to date give depth estimates in a restricted range between 353 and 455 km (using the Beyer and Frost 2017 barometer) or 368 to 462 km (using the Thomson et al. 2021 barometer) and are in part responsible for the prominent peak at ~400 km in Figure 10. If these inclusions also contained exsolutions and the available electron microprobe data do not faithfully reflect their original compositions, then the overall peak could be shifted to somewhat greater depths or, at least, the overall depth distribution would be smoother than shown in Figure 10. Thomson et al. (2021) suggested that all majoritic garnet inclusions in South American diamonds probably contain exsolutions. They also noted that in the eight inclusions for which both exsolved and reconstructed bulk compositions were available the amount of exsolution is a simple function of the calculated pressure. Accordingly, they devised an empirical correction that can be applied to all inclusions for which reconstructed bulk compositions are not available. When this correction is applied to South American diamonds, the peak for eclogitic majoritic garnet inclusions is smoothed and shifted to ca. 100 km greater depth. It remains unclear if this correction, which is based on a restricted number of inclusions, has any general validity at either global or local scale. In any case, the high abundance of depth estimates between ~300 and ~500 km, even for inclusions that are certainly clinopyroxene-free or whose original composition was reconstructed, suggests that most of the eclogitic/pyroxenitic majoritic garnets indeed originated from this depth range (Harte 2010), or perhaps from a slightly more extended range of ~300 to ~600 km if the Thomson et al. (2021) correction for South American diamonds is valid. This range overlaps with independent depth estimates obtained from phase stability constraints for other sublithospheric diamonds containing Ca-silicate inclusions (300–360 km; Brenker et al. 2005; Anzolini et al. 2016) and with a minimum depth estimate obtained from elastic barometry for a single ferropericlase inclusion (~450 km; Anzolini et al. 2019). Of further interest, the 300–600 km range corresponds to the depths at which carbonated MORB slabs are most likely to intersect their solidus, providing an ideal environment for focused melting and potential diamond growth (Thomson et al. 2016, 2021). Considering that many periclase-magnesiowüstite inclusions, for which barometric data are not available, may also have formed from subduction-related melts (Thomson et al. 2016; Nimis et al. 2018) the actual proportion of diamonds from the 300–600 km interval may be even larger than shown in Figure 10. The reader is referred to Walter et al. (2022, this volume) for a thorough discussion on these and related aspects.A second favorable region for diamond formation is probably located near the upper mantle–lower mantle boundary and uppermost lower mantle (Harte 2010). Evidence for diamond formation in this region is provided by a number of key mineral associations within individual diamonds, mostly consisting of various combinations of periclase, (Mg,Fe)SiO3, (Mg,Fe)2SiO4, jeffbenite (formerly known as TAPP), breyite, corundum and SiO2, and of a NaAl-pyroxene phase (believed to have formed as highly majoritic garnet) alongside with periclase or jeffbenite (Hutchison et al. 2001; Harte 2010). Although in several diamonds only part of the key mineral assemblages is actually represented, the general consistency between the observed mineral associations led Harte (2010) to suppose that a large number of diamonds could come from the relatively narrow depth range of 600 to 800 km. Most inclusions in diamonds from this depth range are ultramafic, as opposed to the mafic inclusions that dominate in the 300–600 km interval.Mineral inclusions are very rare in diamonds and occur in only about 1% of examined stones, with significant differences between individual localities (Stachel and Harris 2008). Moreover, only a few mantle minerals or mineral assemblages that may occur as inclusions in diamonds are suitable for thermobarometry. Therefore, only a small proportion of natural diamonds may yield robust information on the P–T conditions of their formation, and these diamonds are mostly peridotitic. This proportion may increase in the future, in parallel with the development of new thermobarometric methods or refinement of existing ones. A significant field of potential and desirable improvement concerns the eclogitic inclusions, which represent a major fraction of inclusions in lithospheric diamonds, but for which existing barometric methods still provide inconsistent results. A dedicated test on clinopyroxene–garnet pairs similar to that performed on chromian diopsides by Ziberna et al. (2016) might shed more light on the minimum analytical quality that is required for reliable barometry of these inclusions. Extending the applicability of non-traditional methods, such as elastic barometry, to non-ideal configurations and to a more diverse list of inclusion minerals might also have a dramatic effect on the number of diamonds suitable for thermobarometry, although the need for elastically preserved inclusions generally restricts applications to relatively small inclusions not surrounded by fractures. An interesting advantage of the elastic method is that it can rely on non-destructive techniques and, thus, can leave the samples intact and ready for further investigation with other methods. Such investigations require the use of undamaged samples in which internal strains are well preserved. Legacy destructive techniques used to expose minerals included in diamond, unfortunately, prohibit future examination of these unique samples for purposes of elastic barometry.Reanalysis of some previously investigated inclusions or diamondiferous xenoliths might provide a relatively effortless way to increase the number of diamond P–T data. Four of 195 published chromian clinopyroxene analyses and 6 of 156 published orthopyroxene analyses had to be discarded while preparing this review only because of sub-standard analytical quality. Moreover, twenty-five clinopyroxenes were rejected because their compositions fell just off the major-element compositional fields that are used to classify garnet-associated chromian diopsides (Ziberna et al. 2016). Trace element analysis of these diopsides might have allowed more robust discrimination (e.g., based on Sc/V relationships; Nimis et al. 2009) of the garnet-associated samples and many of them might have turned out to be suitable for thermobarometry. Reintegrating all these samples would increase the number of P–T data for chromian diopside-bearing diamonds by a remarkable 27% and for diamonds overall by 17%.An enlargement of the number of P–T data for diamonds would have significant positive outcomes in terms of sample bias reduction. Sample bias is presently a major problem, since the number of localities for which a statistically significant number of sufficiently reliable P–T estimates are currently available is limited at best to three and all are from the Kaapvaal craton (Nimis et al. 2020). Consequent benefits of sample bias reduction would be a better assessment of diamond depth distributions at individual localities, and a more robust evaluation of diamond P–T systematics and of their significance in terms of diamond forming processes and preservation.If the number of P–T data for diamonds is important, their internal consistency may be even more so. In fact, most of the combinations of thermobarometric methods that may be applied to diamond inclusions do show some degree of inconsistency with one another. These inconsistencies mostly stem from simplified thermodynamic treatment of solid solutions in minerals and from uncertainties in the data used for their calibration and derived either from experiments or from independent xenolith thermobarometry. The internal consistency of thermobarometric methods is particularly critical when comparing data for different inclusion types, as commonly is necessary in diamond studies. This is another field of potential scientific advance that may eventually contribute to an improved understanding of diamond forming processes.This review is the result of the knowledge, expertise and generosity of many people. Fruitful collaboration through the years with many of them, and particularly with W.R. Taylor, H. Grütter, L. Ziberna, F. Nestola and R.J. Angel, has been of decisive importance in producing my personal contributions to diamond thermobarometry. Yet many more people have in fact contributed to this review through their enormous scientific work and by providing the community with invaluable data over about five decades. Most of them appear in the reference list below. I express my sincere gratefulness to all of these people and, in particular, to T. Stachel, who generously shared a diamond inclusion database, which served as a useful basis to build the updated database used in this work. I am also grateful to A. Thomson, for help in majorite barometry calculations, and to H. Grütter, G. Brey and editor T. Stachel for their very helpful and thorough reviews and comments.
求助全文
通过发布文献求助,成功后即可免费获取论文全文。 去求助
来源期刊
自引率
0.00%
发文量
0
×
引用
GB/T 7714-2015
复制
MLA
复制
APA
复制
导出至
BibTeX EndNote RefMan NoteFirst NoteExpress
×
提示
您的信息不完整,为了账户安全,请先补充。
现在去补充
×
提示
您因"违规操作"
具体请查看互助需知
我知道了
×
提示
确定
请完成安全验证×
copy
已复制链接
快去分享给好友吧!
我知道了
右上角分享
点击右上角分享
0
联系我们:info@booksci.cn Book学术提供免费学术资源搜索服务,方便国内外学者检索中英文文献。致力于提供最便捷和优质的服务体验。 Copyright © 2023 布克学术 All rights reserved.
京ICP备2023020795号-1
ghs 京公网安备 11010802042870号
Book学术文献互助
Book学术文献互助群
群 号:481959085
Book学术官方微信