<<

Earth and Planetary Science Letters 396 (2014) 201–212

Contents lists available at ScienceDirect

Earth and Planetary Science Letters

www.elsevier.com/locate/epsl

Constraining the cause of the end- extinction with coupled records of carbon and calcium isotopes ∗ Adam B. Jost a, ,RolandMundilb,BinHec,ShaunT.Brownd,e, Demir Altiner f, Yadong Sun g, Donald J. DePaolo d,e, Jonathan L. Payne a a Department of Geological and Environmental Sciences, Stanford University, Stanford, CA 94305, USA b Berkeley Center, Berkeley, CA 94709, USA c State Key Lab of Isotope Geochemistry, Chinese Academy of Sciences, Guangzhou, Guangdong 510640, PR d Department of Earth and Planetary Science, University of California, Berkeley, CA 94720, USA e Geochemistry Department, Earth Science Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94270, USA f Department of Geological Engineering, Middle East Technical University, Ankara 06531, Turkey g State Key Laboratory of Geobiology and Environmental Geology, China University of Geosciences, Wuhan, Hubei 430074, PR China article info abstract

Article history: Anegativeδ13C excursion in carbonate sediments near the Guadalupian/ (Middle/Late ) Received 26 October 2013 boundary has been interpreted to have resulted from a large carbon cycle disturbance during the end- Received in revised form 18 March 2014 Guadalupian (ca. 260 Ma). However, the carbon isotope data alone are insufficient Accepted 6 April 2014 to uniquely determine the type and magnitude of perturbation to the global carbon cycle. Calcium Available online 3 May 2014 isotopes can be used to further constrain the cause of a carbon isotope excursion because the carbon Editor: J. Lynch-Stieglitz and calcium cycles are coupled via CaCO3 burial. In this study, we present coupled carbon and calcium Keywords: isotope records from three Guadalupian–Lopingian (G/L) sections in China (Penglaitan and ) and 13 44/40 Turkey (Köserelik Tepe). The δ Candδ Ca records differ among our studied sections and do not Emeishan co-vary in the same manner. No section shows δ13Candδ44/40Ca changes consistent with massive, Middle Permian rapid volcanic CO2 emissions or methane clathrate destabilization. Additionally, many sections with large diagenesis (>3h) changes in δ13C exhibit δ18O evidence for diagenetic alteration. Only one section exhibits a large large igneous province excursion in the δ44/40Ca of but the absence of a similar excursion in the δ44/40Ca of apatite suggests that the limestone excursion reflects a mineralogical control rather than a perturbation to the global calcium cycle. Hence, we interpret the large isotopic changes observed in some sections to have resulted from local burial conditions or diagenetic effects, rather than from a large carbon and calcium cycle disturbance. Perturbations to the global carbon and calcium cycles across the G/L transition were much less intense than the disturbances that occurred across the subsequent Permian– boundary. This finding is consistent with the much smaller magnitude of the end-Guadalupian extinction relative to the end-Permian. © 2014 Elsevier B.V. All rights reserved.

1. Introduction Clapham, 2012; Stanley and Yang, 1994). Given strong evidence for a causal link between Siberian Traps volcanism and the end- 1.1. Overview Permian extinction (Payne and Clapham, 2012), the emplacement of the Emeishan large igneous province in the latest Guadalupian One of the least understood extinction events of the Phanero- raises the possibility of a similar trigger for the end-Guadalupian zoic occurred at the end of the Guadalupian (ca. 260 Ma) extinction (Courtillot and Renne, 2003). In the case of the end- (Fig. 1)(Bambach, 2002; Raup and Sepkoski, 1982). This extinc- Permian extinction, calcium and carbon isotope records provide tion event was temporally distinct, and causally independent of the evidence that the extinction event was linked to acidification subsequent end-Permian mass extinction (ca. 252 Ma) (Clapham following massive CO2 release caused by Siberian Traps volcan- and Payne, 2011; Clapham et al., 2009; Jin et al., 1994; Payne and ism (Hinojosa et al., 2012; Payne et al., 2010). Because no calcium isotope records exist for the end-Guadalupian, it is currently un- * Corresponding author. known whether a similar change in ocean chemistry occurred dur- E-mail address: [email protected] (A.B. Jost). ing this earlier biodiversity crisis. http://dx.doi.org/10.1016/j.epsl.2014.04.014 0012-821X/© 2014 Elsevier B.V. All rights reserved. 202 A.B. Jost et al. / Earth and Planetary Science Letters 396 (2014) 201–212

1.2. Calcium isotopes and the ocean calcium cycle

We express calcium isotope ratios as the relative abundance of 44Ca versus 40Ca using standard δ notation:   44Ca 40 sample δ44/40Ca =  Ca  − 1 × 1000 44Ca 40Ca BSE We report our δ44/40Ca data relative to bulk silicate earth (BSE) rather than seawater because the calcium isotope composition of seawater has been shown to vary over the (Nielsen et al., 2011). On the BSE scale, modern seawater is approximately 0.9h, and SRM-915a is −1.0h (Nielsen et al., 2011). Calcium iso- topes in CaCO3 sediments are fractionated relative to seawater by h Fig. 1. Middle–Late Permian timescale illustrating the different epoch and 1.0 to 1.9 during the formation of CaCO3 sediments (Blättler et nomenclature, as well as the conodont zonation for the Capitanian and early Wuchi- al., 2012; DePaolo, 2004; Gussone et al., 2005); in part, this range apingian. Ages are from the International Chronostratigraphic Chart (v2013/01) of reflects the ∼0.9h difference in fractionation between calcite and the International Commission on Stratigraphy (ICS). aragonite (Blättler et al., 2012; Gussone et al., 2005; Marriott et al., 2004; Nielsen et al., 2011). Because calcium influxes have similar Geochemical proxy data from marine suggest the isotopic compositions and the ocean is well-mixed with respect end-Guadalupian extinction was associated with an environmen- to calcium due to its long residence time (Hippler et al., 2003; 44/40 tal crisis. The redox chemistry of deep-sea cherts indicates that Schmitt et al., 2001), substantial changes in the δ Ca of sea- the end-Guadalupian was the start of long-lasting and wide-spread water are likely either due to mass flux imbalances or changes in ocean anoxia, which later increased in intensity at the end of the fractionation. Changes in fractionation associated with shifts in the primary CaCO polymorph being deposited in the appear to Permian (Isozaki, 1997). A negative shift in δ13C occurs in several 3 explain long-term changes in the δ44/40Ca of seawater (Blättler et stratigraphic sections near the Guadalupian/Lopingian (G/L) bound- al., 2012; Farkaš et al., 2007). ary (Bond et al., 2010b; Isozaki et al., 2007; Kaiho et al., 2005; We can distinguish whether δ44/40Ca variation in limestone Lai et al., 2008; Shen et al., 2013; Tierney, 2010; Wang et al., 2004; resulted from changing mineralogy or from changing seawater Wei et al., 2012; Wignall et al., 2009a; Yan et al., 2013). These δ13C δ44/40Ca by using other Ca-bearing minerals (Fantle, 2010). This fluctuations are consistent with some perturbation to the carbon approach has been applied to the calcium isotope record cycle, and have been variously attributed specifically to sea-level using marine barite (Griffith et al., 2008a, 2008b, 2011) and to change (Wang et al., 2004), ocean stratification and anoxia (Isozaki the Permian–Triassic boundary interval using conodont microfos- sils (Hinojosa et al., 2012). Here, we use the δ44/40Ca of et al., 2007; Tierney, 2010), methane release (Bond et al., 2010a; (hydroxyapatite) to differentiate changing mineralogy from global Retallack and Jahren, 2008; Retallack et al., 2006), and Emeishan changes in seawater δ44/40Ca. In the case of a global flux imbal- volcanism (Wei et al., 2012). ance, the δ44/40Ca records from carbonates and conodonts should 13 Because any negative δ C excursion may be explained through mirror each other. If the local sediment mineralogy shifted from several mechanisms, identifying the precise cause requires addi- aragonite to calcite, the δ44/40Ca of carbonates would increase but tional geochemical proxies. Calcium isotopes can provide further the δ44/40Ca of conodonts would not. However, if there was a constraint on the possible causes of δ13C excursions because the global change from aragonite to calcite deposition, we expect to 44/40 carbon and calcium cycles are linked through weathering and find a positive excursion in carbonate δ Ca, and a negative shift in conodont δ44/40Ca as seawater δ44/40Ca adjusts to the new car- the burial of CaCO (Blättler et al., 2011; De La Rocha, 2000; 3 bonate fractionation factor (Fig. 2). Holmden et al., 2012a; Payne et al., 2010). For example, the large The δ44/40Ca composition of conodonts may be influenced by 13 negative δ C excursion associated with the end-Permian event has vital effects. We attempt to minimize this issue by analyzing 13 been attributed to volcanic input of C-depleted CO2 (Payne and monogeneric samples, and assume that any vital effects are con- Kump, 2007), ocean stagnation and overturn (Grice et al., 2005; stant during our study interval. Isozaki, 1997; Kump et al., 2005; Nielsen and Shen, 2004), and col- lapse of the biological pump (Rampino and Caldeira, 2005). How- 2. Geologic setting ever, the associated excursions in both δ13C and δ44/40Ca can only 2.1. Penglaitan, Province, China be accounted for by volcanic CO2 input and subsequent ocean acid- ification (Hinojosa et al., 2012; Payne et al., 2010). We collected limestone rock and conodont samples from the To better constrain the nature of the end-Guadalupian event, G/L global stratotype section and point (GSSP) at Penglaitan, south here we present coupled calcium and carbon isotope records for China (Fig. 3). Limestone samples were collected over 150 m of the G/L boundary interval at three different stratigraphic sections: section exposed along the Hongshui River, beginning near the base Penglaitan (Guangxi, China), Chaotian (Sichuan, China), and Kösere- of Member IV of the Maokou Formation (interbedded radiolar- lik Tepe (Adana, Turkey). We compare these new data with existing ian chert, micritic limestone, and sandstone), through Member V δ13C and δ18O data from sections in China, Iran, and Japan to il- (a.k.a., the Laibin Limestone, a massive, fossiliferous limestone), and terminating in the lower Heshan Formation (interbedded ra- lustrate the variations in these isotope proxy records. We show diolarian chert and micritic limestone) (Jin et al., 2006). The G/L that the geochemical changes during this interval are not compat- boundary lies within the uppermost Laibin Limestone, at the first ible with significant global acidification in response to Emeishan appearance datum (FAD) of postbitteri postbitteri (Jin et al., volcanism or methane release, and that only a small carbon cycle 2006). Conodont samples were collected beginning 30 m below the perturbation, if any, could have taken place. G/L boundary and through the remainder of the section. A.B. Jost et al. / Earth and Planetary Science Letters 396 (2014) 201–212 203

Fig. 2. Predicted calcium isotope records for seawater, carbonates, and conodonts in four hypothetical scenarios: In steady state, the carbonates and conodonts are offset from seawater based on their fractionation factor. The global flux imbalance represents a temporary change in carbonate mass balance; in this case, the calcium outflux exceeds influx for a brief interval. A local mineralogy shift (here from aragonite to calcite) affects the fractionation factor for carbonate, but does not haveaneffectontheglobal calcium isotope compositions of seawater. A global mineralogy (aragonite to calcite) shift affects the fractionation factor for carbonate permanently, and changes the isotopic composition of seawater and conodonts.

nous limestone followed by massively-bedded, algal-rich wacke- stone and packstone containing chert beds and nodules. Conodonts are not abundant at this section, so biostratigraphic constraints are poor. The upper limestone portion of the Maokou Formation belongs to postserrata zone, and the low- ermost unit of the late Capitanian mudstone of the Maokou Fm. to the J. shannoni zone (Isozaki et al., 2008; Lai et al., 2008). No con- odonts were recovered from the lower Wuchiaping Formation at Chaotian, nor the nearby Shangsi section (Sun et al., 2008).

2.3. Köserelik Tepe, Adana Province, Turkey Fig. 3. Late Permian paleogeographic reconstruction (Ron Blakey, NAU Geology) with the localities featured in our study: (A) Penglaitan, Guangxi, China, (B) Chaotian, The Köserelik Tepe section is located in the Adana Province of Sichuan, China, and (C) Köserelik Tepe, Adana, Turkey; as well as the localities from previous studies discussed here: (1) Tieqiao, Guangxi, China, (2) Kamuru, Miyazaki, Turkey (Fig. 3), 80 km southeast of the city of Kayseri. We sam- Japan, (3) Abadeh, Fars, Iran, and (4) Xiong Jia Chang and Gouchang, Guizhou, China. pled limestone rock over 350 m of exposed outcrop, all belonging to the Yıgılıtepe˘ Formation (Demirtaslı, ¸ 1967; Özgul et al., 1973), The Capitanian/ boundary was located in the with our section beginning in the lower Capitanian (just above field according to Jin et al. (2006). The position of the Wor- the youngest quartz arenite deposit), and ending in the upper dian/Capitanian boundary was estimated based on conodont bios- Wuchiapingian. The Yıgılıtepe˘ Formation contains shallow-water tratigraphy from the nearby Tieqiao section (Mei et al., 1998). shelf limestone deposits, with abundant algae and foraminifera, along with corals, , and in lesser abundances. The section exhibits four transgressive–regressive cycles, which are 2.2. Chaotian, Sichuan Province, China reflected in sedimentary transitions from lime mudstone to fos- siliferous packstone to wackestone with some cycles capped with The Chaotian section is located in northern Sichuan Province dolomite. (Fig. 3), approximately 30 km north of the city of Guangyuan, and At Köserelik Tepe, the last appearance of fusuline foraminifera has been the focus of several sedimentological (He et al., 2010; (schwagerinid) occurs at 131.8 m. Samples between 196 to 218 m Isozaki et al., 2008, 2004; Saitoh et al., 2013b) and geochemical are rich in Shanita foraminifera, which places this part of the studies (Lai et al., 2008; Saitoh et al., 2013a). We sampled lime- section in the upper Capitanian. The level at 223 m marks the stone over approximately 100 m of exposed outcrop along the Jial- beginning of the Paraglobivalvulina mira-Reichelina Zone, which ing River, beginning in the lower Capitanian portion of the Maokou corresponds to the lower Wuchiapingian (Altiner et al., 2000; Formation, and ending in the mid-upper Wuchiapingian portion of Gaillot and Vachard, 2007). Therefore, we tentatively place the the Wuchiaping Formation. The lower 45 m of Maokou Formation Capitanian/Wuchiapingian boundary at 223 m. we sampled consists of thickly-bedded, fossiliferous packstones and wackestones, in some parts containing thick chert beds and 3. Methods large chert nodules. The upper 15 m of the Maokou Formation (up- 3.1. Sample preparation per Capitanian) consists of dark gray lime mudstones with abun- dant interbedded chert, phosphate nodules, and dolomite (Isozaki We cut limestone samples using a tile saw and prepared pow- et al., 2008; Saitoh et al., 2013a, 2013b). The top of the Maokou ders using a bench-top drill press equipped with a 0.8 mm dental and the G/L boundary at this location is at the base of a 1.5 m- drill bit. We collected only micrite where possible, and avoided thick volcanic ash bed, referred to as the Wangpo Tuff, although skeletal grains and sparry calcite cements. We extracted conodont this layer is also argued to be erosional products of Emeishan elements from bulk rock by dissolution with 10% acetic acid and flood basalts (He et al., 2010). The overlying Wuchiaping Formation heavy liquid separation (sodium polytungstate) at China University begins with sandy limestone, which is overlain by dark, bitumi- of Geosciences in Wuhan, China. 204 A.B. Jost et al. / Earth and Planetary Science Letters 396 (2014) 201–212

3.2. Trace element analysis 2σ < 0.1h based on duplicate measurements, we assigned them a 2σ of 0.1h, which is our long-term reproducibility of NIST-915a. Limestone powders were digested in 1 M acetic acid overnight, centrifuged, and decanted to separate the dissolved fraction from 4. Results the insoluble fraction and avoid contamination from non-carbonate phases such as clays and organic matter. Elemental concentrations 4.1. Penglaitan of Ca, Mg, Sr, and Mn were determined by inductively-coupled plasma optical emission spectrophotometry (ICP-OES) at Stanford At Penglaitan, δ13C values start near +5.3h in the upper Wor- University. dian, then decrease to values between +3.2 and +4.0h for most of the Capitanian. In the uppermost Capitanian, δ13C values in- 3.3. Carbon and oxygen isotope analysis crease slightly to just over +5.0h, and then steadily decrease to near +3.3h at the end of the section (Fig. 4). Carbon isotope val- All carbon and oxygen isotope analyses were performed at Stan- ues are typical for the Middle Permian (+3to+5h)(Korte et ford University. Three hundred and sixty-five samples were ana- al., 2005), and agree with previous measurements from Penglaitan lyzed using a Finnigan MAT252 coupled to a Kiel III carbonate de- (Chen et al., 2011; Wang et al., 2004). vice. Between 60 and 250 μg of limestone powder were reacted in The largest change in δ13C at Penglaitan is a small positive ◦ phosphoric acid (H3PO4)at70 C for 600 s. Precision on SRM-8544 shift of approximately 1h below the G/L boundary, followed by (NBS-19) and internal standard M2 were 0.06h and 0.035h (1σ ) an approximately 2h decrease until the end of the measured sec- for oxygen and carbon, respectively. tion. Our records do not extend far enough into the Wuchiapingian The remaining 105 samples were analyzed using a Thermo to indicate at what value the shift ends, though data from the Finnigan Gasbench interfaced with a Thermo Finnigan Delta Plus nearby Tieqiao section suggest δ13C returned to Middle Permian XL mass spectrometer via a Thermo Finnigan ConFlo III unit. Be- values (∼+5h)(Tierney, 2010). Carbon isotope values are signif- tween 200 and 700 μg of limestone powder were weighed into icantly positively correlated with oxygen isotope values (Table 1), sealed vials that were flushed with He gas and reacted with 0.5 but not with any other geochemical parameter (δ44/40Ca, Mn/Sr, ◦ mL of phosphoric acid for 1 h at 72 C. External precision of oxygen Sr/Ca, Mg/Ca) at Penglaitan. and carbon isotope data is <0.1h (1σ ), based upon repeated mea- Oxygen isotope values at Penglaitan vary between −8.8 and surements of two internal lab standards (calibrated against NBS 18, −1.9h, with the majority between −7.0 and −3.0h. There is no NBS 19, and LSVEC). All δ13C and δ18O values are reported relative apparent stratigraphic trend in δ18O at this section, except for a to V-PDB. small (∼1h) shift towards lower values above the G/L boundary. Oxygen isotopes at Penglaitan are not correlated with any other 3.4. Calcium isotope analysis geochemical parameter except δ13C. The bulk limestone samples from the first 128 m have no ob- All calcium isotope analyses were performed at the Center for vious trend in δ44/40Ca and have a mean value of −0.38h.Inthe Isotope Geochemistry, University of California, Berkeley. Bulk lime- uppermost Capitanian, in the Laibin Limestone, δ44/40Ca reaches stone samples were digested following the same method used for amaximumof+0.01h before returning to a mean of −0.33h the trace element analysis. After separation from the insoluble above 133 m, which is similar to the values found before the G/L fraction, limestone samples were dried and redissolved in 3.0 M boundary excursion (Fig. 4). Background values in this study are HNO3. For conodont samples, we constructed monogeneric sam- comparable to those presented in the lower-resolution Phanerozoic ples (Jinogondolella, Clarkina) using 1–5 individual elements with compilation by Farkaš et al. (2007), and also to those presented in little or no residual carbonate coating. We cleaned samples in the high-resolution end-Permian study by Hinojosa et al. (2012). three 10-min sonic baths each of acetone, 1.0 M ammonium ac- δ44/40Ca at Penglaitan is not correlated with any other geochemi- etate, and 1.0 M acetic acid (9 washes total), and then digested cal parameter. them in 6 M HNO3. The conodont data begin approximately 20 m below the bound- After complete dissolution, we spiked the dissolved samples ary and continue 10 m above the base of the Lopingian. The with a 42Ca- and 48Ca-enriched tracer to allow for correction of δ44/40Ca record from the conodonts lacks a clear positive or neg- mass-dependent fractionation occurring during the thermal ion- ative shift through the sampled interval (Fig. 4). Conodont values ization process. We loaded the spiked solution onto an element- reach a minimum of −0.92h, and a maximum of −0.51h.These specific resin column (DGA Eichrom Technologies) to isolate and values are similar to those of Late Permian and con- + purify Ca2 from the sample matrix. Samples were run in a odonts (Hinojosa et al., 2012). Thermo-Fisher Triton multi-collector thermal ionization mass spec- Mn/Sr values vary between 1 and 3 through most of the section. trometer (TIMS). Calcium isotope values are reported relative to Mn/Sr values are highest near the G/L boundary interval, but with Bulk Silicate Earth (BSE) (40Ca/44Ca = 47.162 and 42Ca/44Ca = most values <5. 0.31221). 4.2. Chaotian 3.5. Data analysis At Chaotian, δ13C values vary mostly between +4.0 and +4.5h Co-variation among all geochemical data was calculated using for the first 45 m of the section before beginning to drop steadily Spearman’s Rank test, and significance (p < 0.05) was assigned af- to almost +3h at 48 m. The interval between 48 m and 70 m ter a Bonferroni correction was applied to account for the number is minimally sampled due to very poor outcrop exposure in an of correlations being assessed (N = 50). organic-rich mixed chert and lime mudstone interval. Carbon iso- We fitted a trend line and a 2σ error envelope to our cal- tope values starting at 60 m suggest a 9h negative shift to −6.1h cium isotope records by resampling our record 200 times using at 62 m in the latest Capitanian. At 67 m, δ13C vary between randomly selected values for each sample level from a normal dis- −2.2h and +1.5h and rebound to pre-G/L values mostly between tribution defined by the mean and standard deviation of each sam- +4.5 and +5.0h by 74 m (Fig. 5). ple. Mean resampled values were plotted using the smooth.spline To fill in the chemostratigraphic record where we were un- function in R (smooth parameter, spar = 0.50). For samples with able to sample outcrop, we used data from Saitoh et al. (2013a), A.B. Jost et al. / Earth and Planetary Science Letters 396 (2014) 201–212 205

Fig. 4. The δ13C, δ18O, and δ44/40Ca record from Penglaitan. The filled data points are bulk carbonate data, and the x-shapes are conodont data. Conodont are from Mei et al. (1998), using the stratigraphic heights from Chen et al. (2011). Error bars on the δ44/40Ca data are 2σ from replicate measurements. Numbers represent FAD of conodont species: 1 – J. shannoni;2–J. altudaensis;3–J. prexuanhaenensis;4–J. xuanhaenensis;5–J. granti;6–Clarkina postbitteri hongshuiensis;7–C. postbitteri postbitteri; 8–C. dukouensis;9–C. asymmetrica.

Table 1 Oxygen isotope values from our samples above and below the Spearman’s rank test results for δ13Cversusδ18O for each section. Significance is muddy G/L boundary interval vary between −7.3 and −3.9h. designated to p-values less than a Bonferroni-corrected 95% confidence threshold, 18 = 13 18 The poorly-exposed mudstone interval has δ O values between which is 0.05/50 0.001. Correlations between δ Candδ Oaresignificantat − h − h Penglaitan, Tieqiao (Tierney, 2010 only), Kamura, and Iran. 12.6 and 3.5 (Saitoh et al., 2013a). Where data from both studies overlap stratigraphically, δ18O values are within 1h of Section ρ p-value each other (Fig. 5). Penglaitan 0.41 <0.001 − Calcium isotope values of the limestone are constant within er- Chaotian 0.25 0.03 − h Köserelik Tepe −0.20 0.01 ror for the first 50 m of the section and average around 0.39 . Tieqiao Values above the G/L boundary begin slightly less than the pre- (Tierney, 2010)0.70 <0.001 boundary average at −0.54h,decreaseto−0.70h, and then in- Chaotian crease to −0.32h by just above 70 m. Values then decrease to (Saitoh et al., 2013a) −0.08 0.47 − h Kamura around 0.50 for the remainder of the section (Fig. 5). Most of (Isozaki et al., 2007) −0.64 <0.001 these changes are within error of each other. Iran Mn/Sr values for our Chaotian samples were all very low (Liu et al., 2012)0.53 <0.001 (<0.13). There are no published Mn/Sr data for the muddy in- Tieqiao terval. There is no significant correlation between any of the geo- (Yan et al., 2013)0.38 0.03 Guizhou chemical proxies we measured at Chaotian, nor is there between (Wignall et al., 2009a; Bond et al., 2010b)NANA the δ13C and δ18O values reported in Saitoh et al. (2013a) (Ta- ble 1). who utilized a core from the same section to measure δ13Cinthe 4.3. Köserelik Tepe poorly-exposed muddy interval in high detail. We fit the data from Saitoh et al. (2013a) into our record based on the position of the At Köserelik Tepe, δ13C values start at +4.1h and increase to G/L boundary, and made minor adjustments to match levels with values between +4.5 and +5.5h by 10 m. At 50 m, there is a similar δ13C values from both datasets (Fig. 5). The combined data 22 m stratigraphic gap due to poor exposure, and samples con- 13 show that the decline in δ C beginning at around 45 m acceler- tinue to have δ13C values between +4.0 and +5.8h range, mostly ates significantly just below 50 m (where our section ends, around >+4.8h. At 130 m, δ13C values begin to decrease to a minimum the FAD of Jinogondolella shannoni, Saitoh et al., 2013a) and reaches of +3.3h at 162 m. After a 14 m gap, values continue between a minimum of −3h. Values between 50 and 70 m vary between +4.0and+5.5h between 176 and 204 m, and again begin de- −6h and +3h, with the majority between −1h and +3h. creasing to a minimum of +1.1h. At this minimum, though, δ13C 206 A.B. Jost et al. / Earth and Planetary Science Letters 396 (2014) 201–212

Fig. 5. The δ13C, δ18O, and δ44/40Ca record from Chaotian. Error bars on the δ44/40Ca data are 2σ from replicate measurements. Open circles are data from Saitoh et al. (2013a), correlated with our section based on lithology and matches with our δ13C data. Conodont biozones are from Saitoh et al. (2013a). Numbers represent FAD of conodont species: 1 – Jinogondolella shannoni. values vary between +1.1 and +4.0h, not including one dolomi- major (>3h) δ13C excursions (Chaotian, Guizhou, and Tieqiao), tized sample (sample KST-246, Mg/Ca = 0.58) with a value of while others contain more minor excursions (1–3h)(Kamura, 13 −6.6h, located stratigraphically just above the δ C minimum. Köserelik Tepe), or no discernable excursion (0–1h) (Penglaitan, 13 After a 4.5 m gap above this dolomitized horizon, δ Cbegins Abadeh). The timing and duration of the largest excursions are not at +4.3h and gradually increases to a maximum of +7.0h at consistent across sections. Based on , the 307 m, and then gradually decreases to values between +4.4 and start of the excursion varies from as early as the J. postserrata zone +4.9h by 338 m, the end of the section (Fig. 6). Carbon isotopes (Chaotian, Saitoh et al., 2013a)toaslateasC. postbitteri postbitteri here are positively correlated with Sr/Ca (p < 0.001, ρ = 0.31) and negatively correlated with Mn/Sr (p < 0.001, ρ =−0.37). (Tieqiao, Yan et al., 2013), while the end varies from as early as Oxygen isotope values vary between −2.5 and −10.7h J. xuanhanensis (Guizhou, Bond et al., 2010b)toaslateasC. duk- throughout the entire section. Samples below 50 m and above ouensis (Tieqiao, Yan et al., 2013). The implied duration of this 200 m are most variable in their δ18O composition. Samples be- excursion varies by section, ranging from <1Myupto∼4My tween 50 and 200 m appear to cycle between relatively high values (Sun et al., 2010). − − h − − h ( 4.0to 2.5 ) at 50 m, to lower values ( 9.3to 5.6 )at There are significant positive correlations between δ13C and − 100 m, and then back to higher values (mostly between 5.0 and δ18O at Tieqiao (in Tierney, 2010 only) and Iran, and a negative −3.3h)(Fig. 6). Oxygen isotopes at Köserelik Tepe are significantly correlation at Kamura (Table 1). There are no δ18Odatareported positively correlated with Mg/Ca (p < 0.001, ρ = 0.49). from Guizhou. To explore which records may have been affected Calcium isotope values at Köserelik Tepe vary between −0 70h . 18 13 and −0.23h for the Capitanian. At the positive δ13Cexcursionin by diagenesis, we plotted δ Oversusδ C for every section, 18 − h the early Wuchiapingian, δ44/40Ca values decrease to −0.87h, and and identified samples with δ O values less than 7 (Figs. 7 18 then return to an average of −0.56h, though all data are within and 8). Limestone with δ O < −7h would imply precipitation ◦ error of each other (Fig. 6). Calcium isotopes do not co-vary signif- at >45 C(Hays and Grossman, 1991), assuming a δ18Oofsea- icantly with any other geochemical parameter. water equal to −1h (relative to V-SMOW) for an ice-free world Mn/Sr values at this section are mostly <0.2 and only three (Savin, 1977). While samples with δ18O greater than −7h may 13 samples are greater than 0.8. In addition to δ C, Mn/Sr also nega- have experienced diagenetic alteration of their elemental and iso- =− tively co-varies with Sr/Ca (p < 0.001, ρ 0.65). topic chemistry, we reason that samples with δ18O less than −7h

13 must have experienced such alteration because temperatures above 4.4. Comparison with previously published δ Crecords ◦ 45 C do not reflect seawater conditions, but formation or alter- ation at depth. This process highlights data points from all sec- We compiled published δ13C and δ18O records spanning the G/L interval from Tieqiao, Guangxi province (Tierney, 2010; Yan tions (except Guizhou), most severely at Tieqiao and Kamura, and et al., 2013), western Guizhou province (Bond et al., 2010a; moderately severely at Chaotian and Köserelik Tepe (Fig. 7). If we Wignall et al., 2009a), Kamura, Japan (Isozaki et al., 2007), and removed these data points, δ13C excursions at Tieqiao and Kamura Abadeh, Iran (Liuetal.,2012)(Fig. 7). Some of the sections exhibit would be absent. A.B. Jost et al. / Earth and Planetary Science Letters 396 (2014) 201–212 207

Fig. 6. The δ13C, δ18O, and δ44/40Ca record from Köserelik Tepe. Error bars on the δ44/40Ca data are 2σ from replicate measurements.

Fig. 7. Comparison of δ13C across all the sections. Red data points are samples with δ18Ovalues<−7h. Data from Guizhou are grey because no δ18O data are available. Numbers represent FAD of conodont species: 1 – Jinogondolella postserrata;2–J. shannoni;3–J. altudaensis;4–J. prexuanhaenensis;5–J. xuanhaenensis;6–J. granti; 7–Clarkina postbitteri hongshuiensis;8–C. postbitteri postbitteri;9–C. dukouensis;10–C. asymmetrica. Conodont records are as reported in Chen et al. (2011) (Penglaitan and Tieqiao), Saitoh et al. (2013a) (Chaotian), and Bond et al. (2010b) (Guizhou). 208 A.B. Jost et al. / Earth and Planetary Science Letters 396 (2014) 201–212

Fig. 9. Comparison of δ44/40Ca at all the sections in this study. Hollow circles are 13 18 Fig. 8. Scatter plot of δ Cversusδ O for the studied sections and comparison bulk limestone δ44/40Ca, and x-shapes are conodont δ44/40Ca (only at Penglaitan). sections from the published literature. Data from Guizhou are not included because The dotted horizontal line is the G/L boundary. The black fit lines are made using 18 18 no δ O data are available. The dashed line is −7h for δ O. All points to the left a Monte Carlo resampling scheme and the spline.smooth function in R.Erroren- ofthedashedline(<−7h) are highlighted in red in Fig. 7. velopes are 2σ . There is no consistent pattern in δ44/40Ca among the three sections in this study, indicating that changes at each section are not related to fluctua- 44/40 44/40 5. Discussion tions in global seawater δ Ca. The positive δ Caexcursioninbulklimestone with no corresponding change in conodont δ44/40Ca below the G/L boundary at Penglaitan is most consistent with local mineralogical control on δ44/40Ca in bulk 5.1. Comparison of isotopic records limestone.

Although many have argued for a global negative excursion in should be consistent across sections. However, conodont biostratig- 13 δ C at the G/L boundary (Bond et al., 2010b; Isozaki et al., 2007; raphy indicates widely varying starting and ending times for the Kaiho et al., 2005; Lai et al., 2008; Liu et al., 2012; Saitoh et al., δ13C isotope excursion, where present. Many of the sections that 2013a; Wang et al., 2004; Wei et al., 2012; Wignall et al., 2009a; have significant δ13C changes have excursions spanning at least Yan et al., 2013), our compilation of new data and data from two conodont zones, implying the event lasted more than one mil- existing sections illustrates that the negative δ13Cexcursionis lion . not always present. Furthermore, where the negative excursion is A global shift in the δ13C composition of seawater lasting for present, the stratigraphic position and magnitude of the excursion is inconsistent. This situation is unlike the end-Permian record this long is unlikely for two reasons. Sections that do not exhibit 13 of δ13C, where many sections across the globe show large and large δ C shifts have no evidence for long depositional hiatuses 13 biostratigraphically-synchronous δ13Cexcursions(Korte and Kozur, spanning these zones. For example, there is no large δ Cshift 2010; Payne and Clapham, 2012; Shen et al., 2013). at Penglaitan, but all Capitanian conodont zones from J. shannoni The disagreement in the δ44/40Ca records among the sections to J. postbitteri postbitteri are present (Mei et al., 1998), meaning we sampled is similar to what we observed in the δ13C records the maximum length of a single time gap, if any, is less than 1 (Fig. 9). We observe a positive excursion at Penglaitan, and no ex- million years. Wignall et al. (2009b) observed a sequence boundary cursion at Chaotian or Köserelik Tepe given the error on our mea- at the base of the Laibin Limestone at Penglaitan [this has also 44/40 surements. There is no positive excursion in the conodont δ Ca been argued to be a storm deposit feature (Qiu et al., 2013)], but it 44/40 data at Penglaitan, implying that seawater δ Ca composi- could not have lasted the >1 million years required to exclude the tion was unchanging through this interval and that the observed δ13C excursion. Thus, the absence of a large excursion at Penglaitan 44/40 changes in the bulk limestone δ Ca record are diagenetically- is unlikely to be from an incomplete stratigraphic record. or facies-controlled (Fig. 2). The long duration of the possible δ13C excursion is also prob- The variability in isotopic records among sections could reflect lematic because it is extremely difficult to keep seawater DIC at a global environmental event at the G/L that was not recorded in ∼−4h for longer than its residence time (∼150 ky) without in- some areas due to stratigraphic hiatuses, or it could indicate that voking a voluminous source of 13C-depleted carbon such as vol- there was, at most, a small carbon cycle perturbation at the G/L canic CO or methane (Halverson et al., 2002). However, these pos- and that the larger changes in δ13C and δ44/44Ca at some sections 2 reflect local effects or diagenetic alteration. We discuss below why sible sources are unlikely. Numerical models suggest that volcanic 13 we prefer the latter scenario. degassing should result in a negative δ C excursion followed by 13 a positive excursion caused by input of C-depleted CO2 and sub- 5.1.1. No evidence for a global carbon/calcium cycle event sequent anoxia and high organic carbon burial (Payne and Kump, 13 If there was a global event at the G/L that was recorded in δ C 2007). Unbuffered CO2 input also decreases CaCO3 saturation (acid- + and δ44/40Ca of limestone, the timing and duration of the event ification) and increases weathering of Ca2 , thereby decreasing the A.B. Jost et al. / Earth and Planetary Science Letters 396 (2014) 201–212 209

+ relative outflux of Ca2 from the ocean (Payne et al., 2010), forcing negatively correlated, which is consistent with dolomitization of 44/40 adecreaseintheδ Ca of CaCO3 sediments. part of the section. Some have argued that the emplacement of Emeishan was a The mudstone interval at Chaotian is also rich in organic car- large source of CO2 (Ganino and Arndt, 2009; Ganino et al., 2008), bon [up to 16 wt% in some horizons (Saitoh et al., 2013b)], and has and a trigger for the extinction event (Courtillot and Renne, 2003). widely varying δ13Cvalues.Thelowδ13Cvaluesmayresultfrom However, the pattern of a linked negative and positive excursion in remineralization of 13C-depleted organic matter during deposition δ13C expected from modeling is not present in any section. Addi- or after burial. Given a lack of a negative correlation between δ13C tionally, we do not observe a negative δ44/40Ca excursion at the G/L and δ18Ointhissection(Table 1), remineralization of organic mat- boundary. This is unlike at the Permian/Triassic boundary where a ter is more likely the cause of anomalously low δ13Cwithoutlow negative excursion occurs in δ44/40Ca of bulk limestone (Payne et δ18O. al., 2010) and conodonts (Hinojosa et al., 2012). Based on these ob- Elevated Mn/Sr ratios are common in limestones with high pro- 13 servations, it is unlikely that Emeishan CO2 release impacted δ C portions of non-marine, diagenetic carbonate (Veizer, 1983). Mn/Sr or δ44/40Ca. is correlated with δ13C at Köserelik Tepe, most clearly in the sam- It is possible to generate a negative δ13C excursion of several ples below the dolomitized horizons. These samples were either permil without a detectable positive excursion by introducing a altered during exposure and meteoric alteration or during dolomi- small amount of highly 13C-depleted (δ13C −30h) carbon, such tization after burial. High Mn/Sr can also be found in samples de- as from methane clathrate destabilization (Retallack and Jahren, posited under reducing conditions. We differentiate between these 2008; Retallack et al., 2006). However, small volumes of CO2 or possibilities based on correlation with other evidence for reducing 5 CH4 need to be released extremely quickly (10 yr) to alter conditions or diagenesis. For example, Penglaitan has very elevated ocean and atmospheric chemistry. Such a scenario fails to account Mn/Sr values in some levels, but the presence of glauconite and for the long duration of the δ13C excursion at many sections. pyrite framboids and lack of correlation between Mn/Sr and any Is it possible that the error of our δ44/40Ca measurements other geochemical proxies suggest the high Mn/Sr values reflect re- obscures an important isotope excursion? Mass balance suggests ducing conditions during deposition. There are no elevated Mn/Sr that this possibility is unlikely. If we assume that the smallest values in our samples from Chaotian, but the mudstone interval δ44/40Ca fluctuation detectable is 0.2h, we can approximate what may be selectively altered. we would expect to find in δ13C. Assuming a seawater [Ca] = 2+ × 12 15 mM (Horita et al., 2002), and a Ca influx of 32 10 mol/yr 5.1.3. Potential mineralogical controls on δ44/40Ca − h at an average of 0.5 and a fractionation between seawater and Decoupling of the conodont and bulk limestone δ44/40Ca = h 2+ CaCO3 1.4 (DePaolo, 2004), the influx of Ca wouldneedto records at Penglaitan suggests a local control, potentially by a more than double for 100 kyr to yield an excursion greater than change in the dominant calcium carbonate polymorph (Fig. 2). h 2+ 0.2 . Assuming that every mole of excess Ca was a result of The oceans in the Late Permian favored aragonite as an inor- excess CO2 input to the atmosphere, a starting concentration of ganic precipitate (Stanley and Hardie, 1998); however, given that = dissolved inorganic carbon (DIC) 2 mM, and an instantaneous crinoids and brachiopods (common reef components) × 18 − h − h excess CO2 input of 3.5 10 mol at 15 , we expect a 7.7 construct calcite skeletons, and that inorganic calcite precipitates 13 excursion in δ C. An excursion of this magnitude should be clearly have been demonstrated to fractionate less than aragonite from recorded in many sections, but we only vaguely observe changes solution (Blättler et al., 2012; Gussone et al., 2005), the δ44/40Ca on that order at Chaotian and Guizhou, the former showing ev- excursion could be attributable to a local increase in calcite idence for alteration and local effects. More depleted sources of grains in an otherwise aragonite-dominated system. 13 carbon would produce even larger excursions in δ C. The positive δ44/40Ca excursion at Penglaitan coincides with a facies change, from deep-water mudstones, to grainstones rich 5.1.2. Evidence for local effects and alteration of δ13C in crinoid, , bivalve, and bryozoan fragments, which Changes in δ13C at the various sections may be due to local we interpret to reflect deposition in a shallow slope environment effects. Oxygen isotopes in limestone can be readily altered dur- (see supplemental Fig. A1). Modern brachiopods have a δ44/40Ca ing diagenesis (Given and Lohmann, 1985). Positive co-variation of composition similar to inorganic calcite, and echinoderms have δ13C and δ18O in some sections (Penglaitan, Tieqiao, Iran) indi- a δ44/40Ca composition between inorganic calcite and aragonite cates that many samples with lower δ13C have been recrystallized (Blättler et al., 2012), the latter possibly a result of vital effects or contain diagenetic cement (Fig. 8)(Banner and Hanson, 1990; from stronger biological control on mineralization (Böhm et al., Given and Lohmann, 1985). If we remove all δ13C values from sam- 2006; Gussone et al., 2006). Therefore, the positive excursion in ples with δ18Ovalues<−7h, assuming those samples are least δ44/40Ca of limestone at Penglaitan may have been generated by a likely to reflect seawater conditions, the negative carbon isotope proportional increase of these fossil grains in the sediment. If the excursions at Kamura and Tieqiao disappear, and Chaotian is left average aragonite grain is −0.6h, and the average calcite grain is as the only section with a significant δ13Cexcursion(Fig. 7). There 0h (Blättler et al., 2012), a mixture of 90% aragonite grains and are no δ18O data published for Guizhou, so we cannot apply this 10% calcite grains will yield an average sediment composition of method for those data. −0.33h. Increasing the proportion of calcite grains to 60% yields There is other evidence that the δ13C excursion in the bound- an average sediment composition of −0.03h. Precise sampling of ary interval at Chaotian does not reflect the δ13Cofseawater.The textures within thick-sections using a microdrill would reveal the mudstone interval, which hosts the negative δ13C excursion, con- relative contributions from these different carbonate components, tains dolomite (Saitoh et al., 2013a), which tends to have higher but this is beyond the scope of this study. δ18O values than normal marine limestone (Degens and Epstein, The Laibin Limestone contains a small volume of pore-filling 1964). Mudstone samples with higher δ18O may reflect higher lev- cement, which may be a source of the elevated δ44/40Ca values els of dolomitization. This is probably also true for the Kamura sec- (see Fig. A1). However, given that the volume of this cement is tion, where the Lopingian-age Mitai Formation is dolomitized and <25%, its δ44/40Ca composition would need to be much greater the underlying Guadalupian-age Iwato Formation is not (Isozaki than 0h to cause a 0.3h excursion in bulk limestone, which is et al., 2007), and at the Köserelik Tepe section, where δ18O and not likely given the range of modern carbonate compositions (see Mg/Ca are positively correlated. Also, δ13C and δ18OatKamuraare Blättler et al., 2012). 210 A.B. Jost et al. / Earth and Planetary Science Letters 396 (2014) 201–212

5.1.4. Potential diagenetic controls on δ44/40Ca 6. Conclusions Preferential diagenesis could also account for stratigraphic vari- ation in δ44/40Ca. During diagenesis, equilibrium fractionation fac- New carbon and calcium isotope data from Penglaitan, Chao- 2+ tor (α) between pore fluid Ca and CaCO3 is very close to 1.000 tian, and Köserelik Tepe sections highlight the spatial and tem- (DePaolo, 2011; Fantle and DePaolo, 2007; Jacobson and Holmden, poral variability in the Guadalupian–Lopingian chemostratigraphic 2008), which makes it only possible to change the δ44/40Ca value record. The absence of any consistent and reproducible pattern in of carbonate sediment soon after burial, when the pore fluid and δ13C and δ44/40Ca among these sections and previously studied sediment are not yet equilibrated (Fantle and DePaolo, 2007). Sea- sections indicates that the excursions present reflect local condi- water and near-surface marine pore waters are enriched in 44Ca tions or diagenetic effects, and that any changes to the isotope relative to carbonate sediment, and therefore pore water may in- compositions of carbon and calcium in the global ocean at this crease the δ44/40Ca of carbonate sediment before chemical equi- time were minor and/or below detection. This does not neces- librium is reached (Fantle and DePaolo, 2007). Assuming any pore sarily exclude the possibility of an environmental perturbation at fluid has much less calcium in it than a host limestone, the max- the end-Guadalupian, but implies that the intensity of any event imum possible change in the δ44/40Ca of limestone due to equi- was small. These geochemical findings are compatible with emerg- libration is likely less than 0.1h (Fantle and DePaolo, 2007), and ing evidence from paleobiological records that the intensity of the cannot fully explain the ∼0.3h excursion we observe at Penglai- end-Guadalupian extinction was much less severe than previously tan. thought. Low sea-level and increased mixing with meteoric waters are other proposed mechanisms for changing δ44/40Ca. For example, Acknowledgements modern groundwater flowing into Florida Bay has lower δ44/40Ca values than seawater, which affects the δ44/40Ca composition of This study was supported by NSF grants 0923669 (R.M.) locally precipitated limestone (Holmden et al., 2012b). This phe- and 0923620 (J.L.P.), and by NASA grant NNX09AN6767 (J.L.P.). nomenon has been proposed as an explanation for changes to S.T. Brown and D.J. DePaolo were supported by the Director, Office δ44/40Ca in the geologic record (Holmden et al., 2012b). We expect, of Science, Office of Basic Energy Sciences, of the U.S. Department then, that lower sea-level would increase the influence of ground- of Energy under Contract No. DE-AC02-05CH11231. The authors water calcium in shallow facies. For Penglaitan, we find the most thank D.A. Fike and 2 anonymous reviewers for their constructive positive δ44/40Ca values in the shallowest marine facies. At Chao- comments; S. Denyszyn and Y. Zhong for assistance in the field; tian, we find the most negative δ44/40Ca values in shallow facies and P. Blisniuk, J.L. Hinojosa, A. Jin, G. Li, D. Mucciarone, P. Srini- above in the lowermost Lopingian (Saitoh et al., 2013b), but other vasan, and T. Owens for assistance in the laboratory. samples from similar facies below this level are more positive. As- suming all carbonate platforms have similar groundwater δ44/40Ca Appendix A. Supplementary material as Florida Bay, mixing between seawater and groundwater cannot account for stratigraphic variation in limestone δ44/40Ca. Supplementary material related to this article can be found on- line at http://dx.doi.org/10.1016/j.epsl.2014.04.014. 5.2. Reconciling the isotopic and fossil records References The end-Guadalupian extinction has long been considered to Algeo, T.J., Twitchett, R.J., 2010. Anomalous Early Triassic sediment fluxes due be one of the major extinctions of the Phanerozoic (Bambach, to elevated weathering rates and their biological consequences. Geology 38, 2002; Stanley and Yang, 1994). However, recent compilation of 1023–1026. taxonomically standardized Late Permian fossil occurrence data, Altiner, D., Özkan-Altiner, S., Koçyigit,˘ A., 2000. Later Permian foraminiferal biofa- which includes extensive new data from south China, reveals that cies belts in Turkey: paleogeographic and tectonic implications. In: Bozkurt, E., extinction rates were actually quite low during the Capitanian Winchester, J.A., Piper, J.D.A. (Eds.), Tectonic and Magmatism in Turkey and the Surrounding Area. In: Geol. Soc. (Lond.) Spec. Publ., vol. 173, pp. 83–96. (Clapham et al., 2009; Payne and Clapham, 2012). These recent Bambach, R.K., 2002. Phanerozoic biodiversity mass extinction. Annu. Rev. Earth findings demonstrate that the end-Guadalupian extinction event Planet. Sci. 34, 127–155. (ca. 260 Ma) was substantially smaller than the subsequent end- Banner, J.L., Hanson, G.N., 1990. Calculation of simultaneous isotopic and trace ele- Permian extinction (ca. 252 Ma). Based on current data, the Capita- ment variations during water-interactions with applications to carbonate diage- nesis. Geochim. Cosmochim. Acta 54, 3123–3137. nian ranks 35th in extinction intensity among marine animals out Blättler, C.L., Jenkyns, H.C., Reynard, L.M., Henderson, G.M., 2011. Significant in- of 110 Phanerozoic stages (M.E. Clapham, pers. comm.). However, creases in global weathering during Oceanic Anoxic Events 1a and 2 indicated unlike other Permian stages (besides the ), Capita- by calcium isotopes. Earth Planet. Sci. Lett. 309, 77–88. nian extinctions preferentially affected marine animals vulnerable Blättler, C.L., Henderson, G.M., Jenkyns, H.C., 2012. Explaining the phanerozoic Ca isotope history of seawater. Geology 40 (9), 843–846. to anoxia and hypercapnia (i.e., non-motile, benthic animals such Böhm, F., Gussone, N., Eisenhauer, A., Dullo, W.-C., Reynaud, S., Paytan, A., 2006. Cal- as corals, sponges, brachiopods, bryozoans, and crinoids) (Clapham cium isotope fractionation in modern scleractinian corals. Geochim. Cosmochim. and Payne, 2011), suggesting that this extinction event was differ- Acta 70, 4452–4462. ent from typical Permian background patterns of extinction, and Bond, D.P.G., Hilton, J., Wignall, P.B., Ali, J.R., Stevens, L.G., Sun, Y., Lai, X., 2010a. was triggered by a specific environmental perturbation. The Middle Permian (Capitanian) mass extinction on land and in the oceans. 13 Earth-Sci. Rev. 102, 100–116. The lack of a quantifiable and coherent excursion in the δ C Bond, D.P.G., Wignall, P.B., Wang, W., Izon, G., Jiang, H.S., Lai, X.L., Sun, Y.D., Newton, 44/40 and δ Ca records among sections suggests that any global car- R.J., Shao, L.Y., Védrine, S., Cope, H., 2010b. The mid-Capitanian (Middle Per- bon cycle event at the end-Guadalupian was much smaller than its mian) mass extinction and carbon isotope record of South China. Palaeogeogr. end-Permian counterpart. The larger changes at the end-Permian Palaeoclimatol. Palaeoecol. 292, 282–294. Chen, B., Joachimski, M.M., Sun, Y., Shen, S., Lai, X., 2011. Carbon and conodont ap- in calcium isotopes (Hinojosa et al., 2012; Payne et al., 2010), car- atite oxygen isotope records of Guadalupian–Lopingian boundary sections: cli- bon isotopes (Payne and Clapham, 2012), marine sedimentation matic or sea-level signal? Palaeogeogr. Palaeoclimatol. Palaeoecol. 311, 145–153. patterns (Algeo and Twitchett, 2010; Payne et al., 2007), and ex- Clapham, M.E., Payne, J.L., 2011. Acidification, anoxia, and extinction: A multiple lo- tinction intensity (Clapham et al., 2009; Payne and Clapham, 2012) gistic regression analysis of extinction selectivity during the Middle and Late Permian. Geology 39, 1059–1062. demonstrate that there were fundamental differences between the Clapham, M.E., Shen, S., Bottjer, D.J., 2009. The double mass extinction revisited: two events, and that the two extinctions had different intensities, reassessing the severity, selectivity, and causes of the end-Guadalupian biotic and possibly different mechanisms. crisis (Late Permian). Paleobiology 35, 32–50. A.B. Jost et al. / Earth and Planetary Science Letters 396 (2014) 201–212 211

Courtillot, V.E., Renne, P.R., 2003. On the ages of flood basalt events. C. R. Isozaki, Y., Kawahata, H., Ota, A., 2007. A unique carbon isotope record across Géosci. 335, 113–140. the Guadalupian–Lopingian (Middle–Upper Permian) boundary in mid-oceanic Degens, E.T., Epstein, S., 1964. Oxygen and carbon isotope ratios in coexisting cal- paleo-atoll carbonates: the high-productivity “Kamura event” and its collapse in cites and dolomites from recent and ancient sediments. Geochim. Cosmochim. Panthalassa. Glob. Planet. Change 55, 21–38. Acta 28, 23–44. Isozaki, Y., Yao, J., Ji, Z., Saitoh, M., Kobayashi, N., Sakai, H., 2008. Rapid sea-level De La Rocha, C.L., 2000. Isotopic evidence for variations in the marine calcium cycle change in the Late Guadalupian (Permian) of the Tethyan side of South China: over the Cenozoic. Science 289, 1176–1178. litho- and biostratigraphy of the Chaotian section in Sichuan. Proc. Jpn. Acad. Demirtaslı, ¸ E., 1967. Pınarbası-Sarız-Ma ¸ gara˘ civarının jeoloji raporu. MTA Enst. Ra- Ser. B, Phys. Biol. Sci. 84, 344–353. poru, Rap. 1935, 129. Jacobson, A.D., Holmden, C., 2008. δ44Ca evolution in a carbonate aquifer and its DePaolo, D.J., 2004. Calcium isotopic variations produced by biological, kinetic, ra- bearing on the equilibrium isotope fractionation factor for calcite. Earth Planet. diogenic and nucleosynthetic processes. Rev. Mineral. Geochem. 55, 255–288. Sci. Lett. 270, 349–353. DePaolo, D.J., 2011. Surface kinetic model for isotopic and trace element fraction- Jin, Y., Zhang, J., Shang, Q., 1994. Two phases of the end-Permian mass extinction. ation during precipitation of calcite from aqueous solutions. Geochim. Cos- Can. Soc. Petrol. Geol. 17, 813–822. mochim. Acta 75, 1039–1056. Jin, Y., Shen, S., Henderson, C.M., Wang, X., Wang, W., Wang, Y., Cao, C., Shang, Q., Fantle, M.S., 2010. Evaluating the Ca isotope proxy. Am. J. Sci. 310, 194–230. 2006. The Global Stratotype Section and Point (GSSP) for the boundary between Fantle, M., DePaolo, D., 2007. Ca isotopes in carbonate sediment and pore fluid from the Capitanian and Wuchiapingian Stage (Permian). Episodes 29, 253–262. 2+ ODP Site 807A: the Ca(aq)-calcite equilibrium fractionation factor and calcite Kaiho, K., Chen, Z.-Q., Ohashi, T., Arinobu, T., Sawada, K., Cramer, B.S., 2005. A neg- recrystallization rates in sediments. Geochim. Cosmochim. Acta 71, ative carbon isotope anomaly associated with the earliest Lopingian (Late Per- 2524–2546. mian) mass extinction. Palaeogeogr. Palaeoclimatol. Palaeoecol. 223, 172–180. Farkaš, J., Böhm, F., Wallmann, K., Blenkinsop, J., Eisenhauer, A., Vangeldern, R., Korte, C., Kozur, H.W., 2010. Carbon-isotope stratigraphy across the Permian–Triassic Munnecke, A., Voigt, S., Veizer, J., 2007. Calcium isotope record of Phanerozoic boundary: a review. J. Asian Earth Sci. 39, 215–235. oceans: implications for chemical evolution of seawater and its causative mech- Korte, C., Jasper, T., Kozur, H.W., Veizer, J., 2005. δ18Oandδ13C of Permian bra- anisms. Geochim. Cosmochim. Acta 71, 5117–5134. chiopods: a record of seawater evolution and continental glaciation. Palaeo- Gaillot, J., Vachard, D., 2007. The Khuff Formation (Middle East) and time- geogr. Palaeoclimatol. Palaeoecol. 224, 333–351. equivalents in Turkey and South China: biostratigraphy from Capitanian to Kump, L.R., Pavlov, A., Arthur, M.A., 2005. Massive release of hydrogen sulfide to the Changhsingian times (Permian), new foraminiferal taxa, and palaeogeographi- surface ocean and atmosphere during intervals of oceanic anoxia. Geology 33, cal implications. Coloq. Paleontol. 57, 37–223. 397. Ganino, C., Arndt, N.T., 2009. Climate changes caused by degassing of sediments Lai, X., Wang, W., Wignall, P.B., Bond, D.P.G., Jiang, H., Ali, J.R., John, E.H., Sun, during the emplacement of large igneous provinces. Geology 37, 323–326. Y., 2008. Palaeoenvironmental change during the end-Guadalupian (Permian) Ganino, C., Arndt, N.T., Zhou, M.-F., Gaillard, F., Chauvel, C., 2008. Interaction of mass extinction in Sichuan, China. Palaeogeogr. Palaeoclimatol. Palaeoecol. 269, magma with sedimentary wall rock and magnetite ore genesis in the Panzhi- 78–93. hua mafic intrusion, SW China. Miner. Depos. 43, 677–694. Liu, X.-c., Wang, W., Shen, S.-z., Gorgij, M.N., Ye, F.-c., Zhang, Y.-c., Furuyama, S., Given, R.K., Lohmann, K.C., 1985. Derivation of the original isotopic composition of Kano, A., Chen, X.-z., 2012. Late Guadalupian to Lopingian (Permian) carbon and Permian marine cements. J. Sediment. Petrol. 55, 430–439. strontium isotopic chemostratigraphy in the Abadeh section, central Iran. Gond- Grice, K., Cao, C., Love, G.D., Bottcher, M.E., Twitchett, R.J., Grosjean, E., Summons, wana Res. 24 (1), 222–232. R.E., Turgeon, S.C., Dunning, W., Jin, Y., 2005. Photic zone euxinia during the Marriott, C.S., Henderson, G.M., Belshaw, N.S., Tudhope, A.W., 2004. Temperature de- Permian–Triassic superanoxic event. Science 307, 706–709. pendence of δ7Li, δ44Ca and Li/Ca during growth of calcium carbonate. Earth Griffith, E., Schauble, E., Bullen, T., Paytan, A., 2008a. Characterization of calcium iso- Planet. Sci. Lett. 222, 615–624. topes in natural and synthetic barite. Geochim. Cosmochim. Acta 72, 5641–5658. Mei, S., Jin, Y., Wardlaw, B.R., 1998. Conodont succession of the Guadalupian– Griffith, E.M., Paytan, A., Caldeira, K., Bullen, T.D., Thomas, E., 2008b. A dynamic Lopingian boundary strata in Laibin of Guangxi, China and west , USA. marine calcium cycle during the past 28 million years. Science 322, 1671–1674. Palaeoworld 9, 53–76. Griffith, E.M., Paytan, A., Eisenhauer, A., Bullen, T.D., Thomas, E., 2011. Seawa- Nielsen, J.K., Shen, Y., 2004. Evidence for sulfidic deep water during the Late Permian ter calcium isotope ratios across the transition. Geology 39, in the East Greenland Basin. Geology 32, 1037. 683–686. Nielsen, L.C., Druhan, J.L., Yang, W., Brown, S.T., DePaolo, D.J., 2011. Calcium iso- Gussone, N., Böhm, F., Eisenhauer, A., Dietzel, M., Heuser, A., Teichert, B., Reitner, tope as tracers of biogeochemical processes. In: Baskaran, M. (Ed.), Hand- J., Wörheide, G., Dullo, W., 2005. Calcium isotope fractionation in calcite and book of Environmental Isotope Geochemistry. Springer-Verlag, Berlin, Heidel- aragonite. Geochim. Cosmochim. Acta 69, 4485–4494. berg, pp. 105–124. Gussone, N., Langer, G., Thoms, S., Nehrke, G., Eisenhauer, A., Riebesell, U., Wefer, G., Özgul, N., Metin, S., Göger, E., Bingöl, I.,˙ Baydar, O., Erdogan,ˇ B., 1973. – 2006. Cellular calcium pathways and isotope fractionation in Emiliania huxleyi. Tertiary rocks of the Tufanbeyli region, eastern Taurus, Turkey. Bull. Geol. Soc. Geology 34, 625. Halverson, G.P., Hoffman, P.F., Schrag, D.P., Kaufman, A.J., 2002. A major perturbation Turkey 16, 82–100. of the carbon cycle before the Ghaub glaciation () in Namibia: Payne, J.L., Clapham, M.E., 2012. End-Permian mass extinction in the oceans: an prelude to snowball Earth? Geochem. Geophys. Geosyst. 3. ancient analog for the twenty-first century? Annu. Rev. Earth Planet. Sci. 2012, Hays, P.D., Grossman, E.L., 1991. Oxygen isotopes in meteoric calcite cements as in- 89–111. dicators of continental paleoclimate. Geology 19, 441. Payne, J., Kump, L.R., 2007. Evidence for recurrent Early Triassic massive volcanism He, B., Xu, Y.-G., Zhong, Y.-T., Guan, J.-P., 2010. The Guadalupian–Lopingian bound- from quantitative interpretation of carbon isotope fluctuations. Earth Planet. Sci. ary mudstones at Chaotian (SW China) are clastic rocks rather than acidic tuffs: Lett. 256, 264–277. implication for a temporal coincidence between the end-Guadalupian mass ex- Payne, J.L., Lehrmann, D.J., Follett, D., Seibel, M., Kump, L.R., Riccardi, A., Altiner, tinction and the Emeishan volcanism. Lithos 119, 10–19. D., Sano, H., Wei, J., 2007. Erosional truncation of uppermost Permian shallow- Hinojosa, J.L., Brown, S.T., Chen, J., DePaolo, D.J., Paytan, A., Shen, S.Z., Payne, J.L., marine carbonates and implications for Permian–Triassic boundary events. Geol. 2012. End-Permian ocean acidification confirmed using calcium isotopes in bio- Soc. Am. Bull. 119, 771–784. genic apatite. Geology 40 (8), 743–746. Payne, J.L., Turchyn, A.V., Paytan, A., DePaolo, D.J., Lehrmann, D.J., Yu, M., Wei, J., Hippler, D., Schmitt, A.D., Gussone, N., Heuser, A., Stille, P., Eisenhauer, A., Nägler, 2010. Calcium isotope constraints on the end-Permian mass extinction. Proc. T.F., 2003. Calcium isotopic composition of various reference materials and sea- Natl. Acad. Sci. USA 107, 8543–8548. water. Geostand. Newsl. 27, 13–19. Qiu, Z., Wang, Q., Zou, C., Yan, D., Wei, H., 2013. Transgressive–regressive sequences Holmden, C., Panchuk, K., Finney, S.C., 2012a. Tightly coupled records of Ca and C on the slope of an isolated carbonate platform (Middle–Late Permian, Laibin, isotope changes during the glaciation event in an epeiric sea setting. South China). Facies 60 (1), 327–345. Geochim. Cosmochim. Acta 98, 94–106. Rampino, M.R., Caldeira, K., 2005. Major perturbation of ocean chemistry and Holmden, C., Papanastassiou, D.A., Blanchon, P., Evans, S., 2012b. δ44/40Ca variability a ‘Strangelove Ocean’ after the end-Permian mass extinction. Terra Nova 17, in shallow water carbonates and the impact of submarine groundwater dis- 554–559. charge on Ca-cycling in marine environments. Geochim. Cosmochim. Acta 83, Raup, D.M., Sepkoski, J.J., 1982. Mass extinctions in the marine fossil record. Sci- 179–194. ence 215, 1501–1503. Horita, J., Zimmermann, H., Holland, H.D., 2002. Chemical evolution of seawater Retallack, G.J., Jahren, A.H., 2008. Methane release from igneous intrusion of coal during the Phanerozoic: implications from the record of marine evaporites. during Late Permian extinction events. J. Geol. 116, 1–20. Geochim. Cosmochim. Acta 66, 3733–3756. Retallack, G.J., Metzger, C.A., Greaver, T., Jahren, A.H., Smith, R.M.H., Sheldon, N.D., Isozaki, Y., 1997. Permo-Triassic boundary superanoxia and stratified superocean: 2006. Middle–Late Permian mass extinction on land. Geol. Soc. Am. Bull. 118, records from lost deep sea. Science 276, 235–238. 1398–1411. Isozaki, Y., Yao, J., Matsuda, T., Sakai, H., Ji, Z., Shimizu, N., Kobayashi, N., Kawahata, Saitoh, M., Isozaki, Y., Ueno, Y., Yoshida, N., Yao, J., Ji, Z., 2013a. Middle–Upper Per- H., Nishi, H., Takano, M., Kubo, T., 2004. Stratigraphy of the Middle–Upper Per- mian carbon isotope stratigraphy at Chaotian, South China: pre-extinction mul- mian and Lowermost Triassic at Chaotian, Sichuan, China. Proc. Jpn. Acad. Ser. B, tiple upwelling of oxygen-depleted water onto continental shelf. J. Asian Earth Phys. Biol. Sci. 80, 10–16. Sci. 67–68, 51–62. 212 A.B. Jost et al. / Earth and Planetary Science Letters 396 (2014) 201–212

Saitoh, M., Isozaki, Y., Yao, J., Ji, Z., Ueno, Y., Yoshida, N., 2013b. The appearance of biostratigraphy and its bearing on mantle plume uplift models. Lithos 119, an oxygen-depleted condition on the Capitanian disphotic slope/basin in South 20–33. China: Middle–Upper Permian stratigraphy at Chaotian in northern Sichuan. Tierney, K., 2010. Carbon and strontium isotope stratigraphy of the Permian from Glob. Planet. Change 105, 180–192. Nevada and China: implications from an icehouse to greenhouse transition. De- Savin, S.M., 1977. The history of the Earth’s surface temperature during the past 100 partment of Geological Sciences, Ohio State University. 179 pp. million years. Annu. Rev. Earth Planet. Sci. 5, 319–355. Veizer, J., 1983. Chemical diagenesis of carbonates: theory and application. In: Schmitt, A.D., Bracke, G., Stille, P., Kiefel, B., 2001. The calcium isotope composition Arthur, M.A., Anderson, T.F., Kaplan, I.R., Veizer, J., Land, L.S. (Eds.), Stable Iso- of modern seawater determined by thermal ionization mass spectrometry. Geo- topes in Sedimentary Geology. SEPM Short Course, p. 3-1-3-100. stand. Newsl. 25, 267–275. Wang, W., Cao, C., Wang, Y., 2004. The carbon isotope excursion on GSSP candi- Shen, S.-z., Cao, C.-q., Zhang, H., Bowring, S.A., Henderson, C.M., Payne, J.L., Davy- date section of Lopingian–Guadalupian boundary. Earth Planet. Sci. Lett. 220, dov, V.I., Chen, B., Yuan, D.-x., Zhang, Y.-c., Wang, W., Zheng, Q.-f., 2013. High- 57–67. 13 resolution δ Ccarb chemostratigraphy from latest Guadalupian through earliest Wei, H., Chen, D., Yu, H., Wang, J., 2012. End-Guadalupian mass extinction and neg- Triassic in South China and Iran. Earth Planet. Sci. Lett. 375, 156–165. ative carbon isotope excursion at Xiaojiaba, Guangyuan, Sichuan. Sci. China, Stanley, S.M., Hardie, L.A., 1998. Secular oscillations in the carbonate mineralogy of Ser. B, Chem. Life Sci. Earth Sci. 55 (9), 1480–1488. reef-building and sediment-producing organisms driven by tectonically forced Wignall, P.B., Sun, Y., Bond, D.P., Izon, G., Newton, R.J., Vedrine, S., Widdowson, shifts in the seawater chemistry. Palaeogeogr. Palaeoclimatol. Palaeoecol. 144, M.,Ali,J.R.,Lai,X.,Jiang,H.,Cope,H.,Bottrell,S.H.,2009a.Volcanism,mass 3–19. extinction, and carbon isotope fluctuations in the Middle Permian of China. Stanley, S.M., Yang, X., 1994. A double mass extinction at the end of the Paleozoic Science 324, 1179–1182. . Science 266, 1340–1344. Wignall, P.B., Vedrine, S., Bond, D.P.G., Wang, W., Lai, X.L., Ali, J.R., Jiang, H.S., 2009b. Sun, Y., Lai, X., Jiang, H., Luo, G., Sun, S., Yan, C., Wignall, P.B., 2008. Guadalu- Facies analysis and sea-level change at the Guadalupian–Lopingian Global Stra- pian (Middle Permain) conodont faunas at Shangsi section, northeast Sichuan totype (Laibin, South China), and its bearing on the end-Guadalupian mass Province. J. China Univ. Geosci. 19, 451–460. extinction. J. Geol. Soc. 166, 655–666. Sun, Y., Lai, X., Wignall, P.B., Widdowson, M., Ali, J.R., Jiang, H., Wang, W., Yan, C., Yan, D., Zhang, L., Qiu, Z., 2013. Carbon and sulfur isotopic fluctuations associated Bond, D.P.G., Védrine, S., 2010. Dating the onset and nature of the Middle Per- with the end-Guadalupian mass extinction in South China. Res. 24 mian Emeishan large igneous province eruptions in SW China using conodont (3–4), 1276–1282.