Next Article in Journal
Mountain Permafrost: A Reflection on the Periglacial Environment in Mongolia
Next Article in Special Issue
Ochotopteris—An Endemic Fern of the Mid-Cretaceous Arctic
Previous Article in Journal
Implementation of the Displacement Discontinuity Method in Geotechnical Case Studies
Previous Article in Special Issue
High-Latitude Depositional Systems, Provenance, and Basinal Setting of the Late Cretaceous Cantwell Basin, Denali National Park and Preserve, Alaska: A Stratigraphic Framework for Paleontological and Paleoclimatic Studies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploring the Deuterium Excess of Cretaceous Arctic Paleoprecipitation Using Stable Isotope Composition of Clay Minerals from the Prince Creek Formation (Maastrichtian) in Northern Alaska

1
Kansas Geological Survey, University of Kansas, Lawrence, KS 66047, USA
2
Department of Geology, University of Kansas, Lawrence, KS 66045, USA
3
Department of Geosciences, University of Alaska, Fairbanks, AK 99775, USA
4
Bureau of Economic Geology, University of Texas, Austin, TX 78712, USA
*
Author to whom correspondence should be addressed.
Geosciences 2023, 13(9), 273; https://doi.org/10.3390/geosciences13090273
Submission received: 29 June 2023 / Revised: 30 August 2023 / Accepted: 8 September 2023 / Published: 10 September 2023

Abstract

:
We report estimated stable isotope compositions of Artic paleoprecipitation using phyllosilicates sampled from three paleosols and two bentonites in the Prince Creek Formation (Maastrichtian) in northern Alaska. Previous studies reported a deuterium excess in estimates of Arctic paleoprecipitation from the Late Cretaceous by combining hydrogen and oxygen proxy sources, including pedogenic minerals, dinosaurian tooth enamel phosphates, pedogenic siderites, and n-alkane biomarkers. The new dataset produced in this study removes uncertainty on possible explanations (photosynthetic and transpiration) of the deuterium excess by producing stable hydrogen and oxygen isotopic signatures from the same source material. The δD of the phyllosilicates range from −171‰ to −72‰ VSMOW and δ18O ranges from 5.0 to 11.8‰ VSMOW. By assuming a MAT of 6.3 °C and calculating uniquely derived fractionation equations for each phyllosilicate, we report estimated isotopic composition of Late Cretaceous paleoprecipitation with an average δD value of −133‰ VSMOW, corresponding to an average δ18O value of −20.3‰ VSMOW. The estimates of Late Cretaceous paleoprecipitation do not intersect the Global Meteoric Water Line and reveal a reported deuterium excess ranging from 7 to 46 per mil. These results confirm the presence of a deuterium excess in Late Cretaceous Arctic paleoprecipitation and provide new insight to assessing possible explanations for this phenomenon.

1. Introduction

The Cretaceous period has been the focus of a substantial amount of paleoclimate research, given its status as an ancient greenhouse world with higher atmospheric pCO2 levels, warmer global temperatures, and reduced equator-to-pole temperature gradients [1,2]. However, gaps in our understanding of the Cretaceous greenhouse still exist, particularly surrounding the polar regions. This is often expressed as a conflict between resolving paleoclimate proxy data, which suggests a warm Cretaceous Arctic, and that of climate model simulations [3]. In an effort to resolve this conflict, new research has focused on collecting and analyzing samples that can reveal the relationship of paleotemperature and stable isotopic (δ18O and δD) composition of paleoprecipitation. This work includes studies on pedogenic siderite, dinosaurian tooth enamel phosphates, and pedogenic clay minerals producing δ18O data and n-alkane biomarkers producing δD data from terrestrial deposits that capture the Cretaceous record in the Arctic [4,5].
The results from this previous research revealed low δ18O of paleoprecipitation corresponding to higher δD than would be expected, given modern relationships of δ18O and δD plotting along the Global Meteoric Water Line [5]. However, the question became if this deuterium excess recorded in n-alkane biomarkers was the result of photosynthetic and transpiration fractionation events occurring within plants producing higher δD values or if it was related to the paleoprecipitation. Here, we address this concern by producing both δ18O and δD data from pedogenic clays and bentonites, thus removing any photosynthetic or transpiration fractionation components from arctic vegetation. These data not only allow us to address the question of the existence and magnitude of the deuterium excess but also allow us to discuss the possible mechanisms that could contribute to its existence.

2. Materials and Methods

Samples from the Prince Creek Formation used in this study were previously collected from three field localities on the North Slope of Alaska. The Colville Basin of northern Alaska formed in the Aptian to Albian as a result of subsidence to the north of the Brooks Range orogenic belt [6,7]. Uplift and erosion of the Brooks Range drove clastic wedges, including the Prince Creek Formation, into the basin during the Late Cretaceous and Paleogene [8]. Bluffs of the Prince Creek Formation along the Colville, Kogosukruk, and Kikiakrorak rivers dip to the northeast (<3°), exposing alluvial sandstone, siltstone, shale, coal and bentonites [9,10]. The entire Prince Creek Formation is known to range in age from Campanian to Paleocene; however, the samples used in this study are thought to be early Maastrichtian, based on palynological and geochronological constraints [11]. Plate reconstructions indicate that the Prince Creek Formation was deposited at a paleolatitude of 82–85 °N [12,13].
Flaig et al. [9,10] interpreted the strata of the Prince Creek Formation to record deposits of meandering fluvial channels, levees, crevasse splays, and associated floodplains (which include lakes, ponds, swamps, and paleosols). Soil-forming processes on floodplains were continually interrupted by alluvial input, resulting in weakly developed paleosols similar to modern aquic Entisols and Inceptisols [10]. At least some of these paleosols were influenced by the presence of volcanic ash [14]. Macro- and micromorphological features in the paleosols indicate repeated, possibly seasonal, wetting and drying of floodplains [10,11]. Overall, the Prince Creek Formation preserves a Cretaceous, paleopolar low-gradient, muddy coastal plain succession [9,10]. The Prince Creek Formation also contains the densest concentration of high-latitude dinosaur bones of any high-latitude location in the Northern or Southern Hemisphere [15], making it an important ancient greenhouse ecosystem for continuing study [16].
With respect to samples collected and analyzed for this study, locality 06SH is located downriver and up-stratigraphic section from localities PFDV 17 and KKT, and is presumed to contain slightly younger sediments. Samples 06SH9.2, 06SH14.4, and 06SH15.1 were collected from a series of stacked, compound-cumulative paleosols exposed along the modern Colville River (Figure 1 and Figure 2). Detailed description and analyses of the paleosols by Flaig et al. [10] identified the paleosols as having characteristics similar to Aquic Inceptisols and Entisols. Sample 06SH9.2 was collected from a dark organic rich B-horizon (10YR 2/1) measuring approximately 70 cm thick. Flaig et al. [10] noted several features within the paleosol, including platy structure, organic fragments, high amounts of bioturbation, common Fe-oxide nodules/coatings, and root traces. Sample 06SH14.4 was collected from a dark olive gray (5Y 3/2) B-horizon measuring approximately 60 cm thick. Noted features included blocky structure and organic fragments [10]. Sample 06SH15.1 was collected from an olive gray (5Y 4/1) B-horizon measuring approximately 80 cm thick. Noted features included blocky structure, micro-sphaerosiderite, root traces, common Fe-oxide nodules/coatings, and papules [10]. Samples PFDV 17 and KKT were collected from bright olive-green bentonite layers measuring approximately 45 cm and 30 cm thick. Bulk samples from each locality and their exact stratigraphic position are indicated by blue stars in Figure 2.
Bulk paleosol material was disaggregated and centrifuged to isolate the <0.2 µm fraction, regarded to be dominated by pedogenic clays in paleosol profiles [17,18,19,20,21,22,23]. The isolated <0.2 µm-size fraction samples were then treated using the following solutions to remove non-phyllosilicate materials: (1) 10% acetic acid solution for 24 h to remove calcite [24,25], (2) sodium citrate–bicarbonate–dithionite solution to remove secondary iron oxy-hydroxides, and (3) 30% H2O2 solution to remove organic matter.
Mineralogy of the chemically treated <0.2 µm fraction samples was determined using X-ray diffraction (XRD). Three oriented aggregates were created for each sample: (1) potassium saturation at room temperature, (2) magnesium saturation at room temperature, and (3) magnesium saturation followed by glycerol solvation at room temperature. Oriented aggregates were analyzed at the Roy M. Huffington Department of Earth Sciences at Southern Methodist University (SMU: Dallas, TX, USA) using a Rigaku Ultima III X-ray diffractometer with a step scan of 0.05° over a spectrum of 2–30° 2θ. Relative abundance of clay minerals in each sample was determined using the area of the background-subtracted intensities of the 001 peak for glycerol-solvated samples (17–20 Å for smectite, 9.9–10.1 Å for illite, 7.1–7.2 Å for kaolinite) [18,22,23,26]. The chemically treated samples were then split into three aliquots for stable isotope and elemental analysis.
The first aliquot was prepared as a pressed powder pellet using 0.25 µm polished hardened steel disks and coated with ~150-Å layer of high-purity carbon in a carbon evaporator. The pressed pellets were analyzed for major and minor elemental composition using a Cameca SXFive electron microprobe housed at the Texas A&M Materials and Characterization Facility.
The second and third aliquots were analyzed for stable oxygen and hydrogen isotope composition at extraction laboratories housed at the Roy M. Huffington Department of Earth Sciences at Southern Methodist University (SMU: Dallas, TX, USA). Oxygen isotope analyses were conducted following the methods of Clayton and Maydeda [27]. Samples were heated overnight in nickel-rod bombs connected to a gas extraction line while kept at 100–150 °C to remove sorbed and interlayer water. Samples were then reacted with BrF5 at ~560 °C overnight to produce O2 gas. The O2 gas was converted to CO2 using heated graphite rods in high-vacuum glass extraction lines. The CO2 gas was then cryogenically captured and measured on a Finnigan MAT 252 isotope ratio mass spectrometer (IRMS) operating in dual inlet mode. Repeated oxygen isotope analyses of the 3 pedogenic phyllosilicate samples yield an analytical error of 0.2‰.
Hydrogen isotope analyses were conducted following the methods of Savin and Epstein [24]. Samples were initially outgassed at 150 °C for a minimum of 1 h under closed-system conditions in 0.16 bar of O2 gas to remove sorbed and interlayer water and destroy any recalcitrant organic matter not destroyed by previous H2O2 treatment. The samples were then dehydroxylated at 850 °C under closed-system conditions in 0.16 bar O2 gas. The low- and high-temperature water fractions were then converted to H2 gas by passage over depleted uranium metal at 760 °C as two separate fractions. H2 gas yields were measured on a mercury manometer with an uncertainty of ±1 µmol, and the δD values were measured on a Finnigan MAT 252 IRMS operating in dual inlet mode at SMU. Replicate hydrogen isotope analyses of phyllosilicate samples yield an analytical error of ±4‰. The oxygen and hydrogen isotopic composition of the <0.2 µm fraction are reported in conventional delta notation (δ) in parts per thousand (‰) relative to the Vienna Standard Mean Ocean Water (V-SMOW; [28]).
Sample 06KKT-20.5 was collected from a bentonite layer along the Colville River and sample PFDV 17 was collected from a bentonite layer along the Kogosukruk River (Figure 1, Figure 2 and Figure 3). Previous work by Salazar-Jaramillo et al. [14,29] analyzed mineralogy and oxygen isotope composition of the clay fraction of both bentonite samples. We obtained bulk samples and separated the clay size fraction <2 µm. The isolated <2 µm-size fraction samples were then treated using the following solutions to remove non-phyllosilicate materials: (1) 10% acetic acid solution for 24 h to remove calcite [24,25], (2) sodium citrate–bicarbonate–dithionite solution to remove secondary iron oxyhydroxides, and (3) 30% H2O2 solution to remove organic matter. Mineralogy of the chemically treated <2 µm fraction samples was determined using X-ray diffraction (XRD). Three oriented aggregates were created for each sample and analyzed at the Geology Department at the University of Kansas (KU: Lawrence, KS) using a Bruker D2 Phaser diffractometer with a step scan of 0.05° over a spectrum of 2–30° 2θ. Results from XRD analyses were used to confirm previous mineralogical analyses reported by Salazar-Jaramillo et al. [14].
Hydrogen isotope analyses of the clay fraction from the two bentonite samples were conducted at the Laboratory for Stable Isotope Science (LSIS), University of Western Ontario, London, Ontario, Canada following methods described by Sharp et al. [30]; VanDeVelde and Bowen [31]; Bauer and Venemann [32]; Qi et al. [33]; and Kanik et al. [34]. Approximately 2 mg of each sample was weighed into 3.5 × 5 mm silver capsules. Duplicates of each sample were run as well as three standards: USGS 57, USGS 58, and in-house biotite standard (GBS). Capsules were folded shut and dried in a vacuum oven at 200 °C for 18 h. Samples were then immediately transferred to a helium (He)-flushed, Costech Zero-Blank autosampler attached to the furnace port of the TCEA. The autosampler was then immediately purged under helium flow of 0.85 bar to limit clay mineral rehydration by atmospheric moisture. The samples and standards were maintained under He flush in the autosampler for 4 h prior to isotopic analyses. The H-isotope data are reported in delta (δ) notation relative to VSMOW-SLAP (Standard Light Antarctic Precipitation) calibrated scale [35]. Phyllosilicate standards USGS 57 (biotite δD = −91.5‰ VSMOW) and USGS 58 (muscovite δD = −28.4‰ VSMOW) were used to calibrate the H-isotope measurements to VSMOW [33].

3. Results

3.1. Clay Mineralogy and Chemical Composition

X-ray diffraction analyses of the <0.2 µm-size fraction from paleosol samples collected from locality 06SH revealed a mixture of smectite (001) peak (18.5 Å), illite/mica (001) peak (10 Å), and kaolinite (001) peak (7.2 Å) (Figure 4, Table 1). The average kaolinite content of the three paleosol samples was ~21%, with the remainder accounted for by 2:1 phyllosilicates. The chemical composition of end-member 2:1 phyllosilicates was calculated using wt% oxide data collected from the <0.2 µm-size fraction (Table 1). The wt% kaolinite of each sample was taken into account based on results from X-ray diffraction spectra. The mole fraction of Al2O3 and SiO2 within mineralogical mixtures of each sample was calculated from the wt% oxide data in Table 1. Estimated mole fractions of Al and Si from kaolinite were subtracted from the phyllosilicate mixtures to calculate an end-member 2:1 phyllosilicate chemical formula for each sample. X-ray diffraction analyses of the <2 µm-size fraction of the bentonite samples 06KKT-20.5 and PFDV 17 confirmed the results of Salazar-Jaramillo et al. [14], revealing a monomineralic signature (pure montmorillonite; Supplementary Table S1).

3.2. δ18O and δD of Meteoric Water from Pedogenic Phyllosilicates

To calculate the δ18O and δD of meteoric water from which the pedogenic phyllosilicates precipitate requires known fractionation factors for both oxygen and hydrogen for each sample as well as crystallization temperature. Given the mineralogical mixture present in the three paleosol samples, unique oxygen and hydrogen fractionation factors were calculated for each sample based on mineralogy and end-member chemical composition (Table 2).
Oxygen fractionation factors for 2:1 phyllosilicates were calculated using the bond-model approach of Savin and Lee [36]. Hydrogen fractionation factors for 2:1 phyllosilicates were calculated using Equation (1), which was developed by Tabor and Montañez [20]:
1000 l n D α = 2.2 × 10 6 / T 2 7.7 + ( 2 X A l 4 X M g 68 X F e )
where XAl, XMg, and XFe are the mole fractions of aluminum, magnesium, and iron present in the octahedral layer of the 2:1 phyllosilicate mineral.
The calculated hydrogen and oxygen fractionation factors for the 2:1 phyllosilicates were then weighted and combined with oxygen and hydrogen fractionation factors for kaolinite, proposed by Sheppard and Gilg [37] (Equations (2) and (3)):
1000 l n 18 α k a o l i n t e - w a t e r = 2.76 × 10 6 / T 2 6.75
1000 l n D α k a o l i n t e - w a t e r = 2.2 × 10 6 / T 2 7.7
The δ18O of meteoric water reported for the bentonite samples by Salazar-Jaramillo et al. 2019 was calculated using the smectite fractionation factor proposed by Sheppard and Gilg [37]. This was chosen due to its monomineralic composition (pure montmorillonite) and close alignment to the Fe-rich end member of the fractionation curve [29]. To calculate the δD of meteoric water, this study calculated unique hydrogen fractionation factors by using chemical data reported by Salazar-Jaramillo et al. [29] and applying the equation developed by Tabor and Montañez [20] (Equation (1)).
A crystallization temperature of 6.3 °C was assumed for all samples based on estimates produced by Spicer and Herman [38]. This temperature has also been assumed by previous attempts to calculate the δ18O of meteoric water [4,29,39].
The stable oxygen isotope composition of the pedogenic phyllosilicates from the paleosol samples ranges from 10.3‰ to 11.8‰ VSMOW with a reported analytical error of ±0.2‰, while hydrogen ranges from −72‰ to −93 ‰ VSMOW with a reported analytical error of ±4‰ (Table 3). When combined with their respective unique fractionation factors and an assumed crystallization temperature of 6.3 °C, the estimated δ18O meteoric water values range from −17.2‰ to −18.8‰ VSMOW and the estimated δD meteoric water values range from −86‰ to −106‰ VSMOW (Table 3, Figure 5).
The stable hydrogen isotope composition of the clay fraction from the bentonite samples ranges from −131‰ to −172‰ VSMOW with a reported analytical error of ±4‰. When combined with previous δ18O data from Salazar-Jaramillo et al. [29], the respective unique hydrogen fractionation factors, and an assumed crystallization temperature of 6.3 °C, the estimated δ18O meteoric water values average −23.6‰ [29] and the estimated δD meteoric water values range from −142‰ to −179‰ VSMOW (Table 3, Figure 5).
Table 3. Measured phyllosilicate oxygen and hydrogen isotope compositions, oxygen and hydrogen isotope fractionation equations, and estimated meteoric water isotopic compositions.
Table 3. Measured phyllosilicate oxygen and hydrogen isotope compositions, oxygen and hydrogen isotope fractionation equations, and estimated meteoric water isotopic compositions.
SampleFormationδ18Omix
V-SMOW (±0.3%)
δDmix
V-SMOW (±4%)
Wt. % Kaolinite (±5%)a 103ln18α2:1-waterb 103lnDα2:1-water = −2.2*106*T−2+c 103ln18αmix-waterd 103lnDαmix-water= −2.2*106*T−2+e δDmeteoric water V-SMOWf δ18Ometeoric water V-SMOW
Paleosol
06 SH 9.2Prince Creek11.8−3232.84 × 106 × T−2–7.03−18.512.82 × 106 × T−2–6.96−13.40−106−17.2
06 SH 14.4Prince Creek10.3−83152.83 × 106 × T−2–6.90−19.432.82 × 106 × T−2–6.87−15.48−95−18.8
06 SH 15.1Prince Creek11.5−72252.81 × 106 × T−2–7.00−18.282.80 × 106 × T−2–6.93−12.97−86−17.3
Bentonite
PFDV 17Prince Creek5.03−13602.55 × 106 × T−2–4.05−19.382.55 × 106 × T−2–4.05−19.38−147−23.6
PFDV 17 duplicatePrince Creek5.03−13202.55 × 106 × T−2–4.05−19.382.55 × 106 × T−2–4.05−19.38−142−23.6
06KKT-20.5Prince Creek4.96−16902.55 × 106 × T−2–4.05−15.442.55 × 106 × T−2–4.05−15.44−177−24.6
06KKT-20.5 duplicatePrince Creek4.96−17202.55 × 106 × T−2–4.05−15.442.55 × 106 × T−2–4.05−15.44−179−24.6
a Calculated oxygen isotope fractionation equations between the <0.2 µm phyllosilicate and water. Fractionation equations were calculated using the bond-model data of Savin and Lee [36] in conjunction with the calculated chemical composition end-member 2:1 phyllosilicates presented in Table 2. b Calculated hydrogen isotope fractionation equations between the <0.2 µm phyllosilicate and water. Fractionation equations were calculated based on the molar fraction of Al, Mg, and Fe in the octahedral layer, as proposed by Gilg and Sheppard [40] and Tabor and Montañez [20]. c Calculated oxygen isotope fractionation equations between mixed phyllosilicate mineralogies in the <0.2 µm phyllosilicate fraction and water. Fractionation equations were calculated based on the molar fraction of oxygen contributed by kaolinite and 2:1 phyllosilicates in each sample [37]. d Calculated hydrogen isotope fractionation equations between mixed phyllosilicate mineralogies in the <0.2 µm phyllosilicate fraction and water. Fractionation equations were calculated based on the molar fraction of hydrogen contributed by kaolinite and 2:1 phyllosilicates in each sample [37]. e Calculated δD of meteoric water in equilibrium with the phyllosilicate mixture, assuming a crystallization temperature of 6.3 °C. f Calculated δ18O of meteoric water in equilibrium with the phyllosilicate mixture, assuming a crystallization temperature of 6.3 °C.
Figure 5. Hydrogen and oxygen isotope plot of estimated ranges of meteoric water isotope compositions from the Prince Creek Formation. The single dot and diamonds represent average composition, with the whiskers showing the full range of δ18O and δD. Results are compared to the modern Global Meteoric Water Line (black line) and local meteoric water line for Alaska (blue line; Rozanski, K. et al. [41]). The green dot represents average δ18O (estimated from dinosaur tooth enamel phosphate, pedogenic siderite, and pedogenic clays) and of δD (estimated from n-alkanes extracted from carbonaceous mudstones) presented in Ludvigson et al. [5]. The open green box surrounding the green dot is the 1 sigma of the dataset. The blue diamond represents the average δ18O and δD estimated from pedogenic clays sampled from site 06 SH (this study). The orange diamond represents the average δ18O and δD estimated from bentonites sampled from sites PFDV 17 and 06 KKT-20.5 (this study).
Figure 5. Hydrogen and oxygen isotope plot of estimated ranges of meteoric water isotope compositions from the Prince Creek Formation. The single dot and diamonds represent average composition, with the whiskers showing the full range of δ18O and δD. Results are compared to the modern Global Meteoric Water Line (black line) and local meteoric water line for Alaska (blue line; Rozanski, K. et al. [41]). The green dot represents average δ18O (estimated from dinosaur tooth enamel phosphate, pedogenic siderite, and pedogenic clays) and of δD (estimated from n-alkanes extracted from carbonaceous mudstones) presented in Ludvigson et al. [5]. The open green box surrounding the green dot is the 1 sigma of the dataset. The blue diamond represents the average δ18O and δD estimated from pedogenic clays sampled from site 06 SH (this study). The orange diamond represents the average δ18O and δD estimated from bentonites sampled from sites PFDV 17 and 06 KKT-20.5 (this study).
Geosciences 13 00273 g005

4. Discussion

4.1. Stable Isotope Composition of Phyllosilicates and Use as Paleoclimate Proxies

Materials collected from paleosols can contain a variety of components, including pedogenic clays, detrital and burial authigenic clays, and diagenetically altered pedogenic clays. In order to ensure a true paleoclimate signal is being recovered from analyzed paleosol materials, several conditions need to be met, including:
  • Phyllosilicates analyzed from a paleosol are authigenic.
  • Phyllosilicates have not been diagenetically altered since the time of formation.
  • The relationship between the oxygen and hydrogen isotope compositions of soil water is known.
In order to satisfy condition 1, paleosol samples analyzed in this study were separated and chemically treated in an effort to isolate a fraction thought to be the best representation of a mineral fraction derived from pedogenesis (see Section 2). The fine clay fraction, <0.2 µm, is regarded as dominated by pedogenic clays in a paleosol profile [17,18,19,20]. Furthermore, mineralogical analyses show the <0.2 µm fraction is dominated by 2:1 smectite and I/S clay minerals with no detrital material such as quartz present. Therefore, we argue the <0.2 µm fraction is pedogenic in origin. The two bentonite samples are argued to have formed in situ as the result of weathering of primary ashfall deposits, resulting in a monomineralic signature. Given its mineralogical composition, we are satisfied with the <2 µm size fraction for paleoclimate analyses.
Several lines of evidence are also presented to satisfy condition 2. Evidence presented by Burns et al. [42] suggests the Prince Creek Formation experienced relatively shallow burial depths of approximately 600 to 2000 m based on downhole vitrinite reflectance profiles and sonic-porosity logs along the Colville River. Vitrinite reflectance values (n = 4, avg. 0.3825; [43,44]) indicate a maximum burial temperature of 48 °C, which is too low to have caused substantial diagenetic changes in paleosol mineralogy [45]. This is also confirmed by studies conducted by Ludvigson et al. [5], which indicated organic matter sampled from the Prince Creek Formation had very low thermal maturity and the n-alkane biomarkers were suitable for δD and δ13C isotopic analyses as their values had not been thermally altered. Additionally, there are well-preserved micromorphological features present in the paleosols [14] and the paleosols remain unlithified. Clay mineral analyses by Salazar-Jaramillo et al. [14] showed no ordering of illite during I/S interstratification analyses of paleosols in the Prince Creek Formation, suggesting little or no diagenetic transformations of illite [46]. Furthermore, the results presented in Figure 5 indicate reasonable oxygen and hydrogen isotopic values of estimated meteoric water that are within the range of previously published isotopic estimates of meteoric water from the Prince Creek Formation using alternative proxies, including pedogenic siderite, dinosaurian tooth enamel phosphates, and n-alkane biomarkers [5].
The bentonites sampled for this study are argued to be the products of in situ weathering of volcanic ash [14,29]. The <2 µm-size clay fraction from the bentonite samples contains pure montmorillonite with no I/S mixed layer clays, indicating that illitization in the Prince Creek Formation is purely a low-temperature pedogenic process [14].
The third condition is met with the assumption that processes that govern the global meteoric water line have been constant for well-developed soils throughout geologic history [41,47,48]. This, combined with uniquely derived oxygen and hydrogen fraction factors for each sample and crystallization temperatures, allows for calculations of meteoric water. The possibility does exist that the authigenic phyllosilicates collected from each paleosol may be composed of several different fractions crystallized under different climate regimes throughout the development of the soil. This scenario cannot be ruled out, and therefore soil-water δ18O and δD estimates are considered to represent average paleoclimatic conditions that persisted throughout the interval(s) of pedogenesis that included phyllosilicate crystallization.

4.2. Implications of δ18O and δD of Meteoric Water from Pedogenic Phyllosilicates

When the calculated δ18O and δD of meteoric water are compared to the Global Meteoric Water Line, all the samples show a deuterium excess, plotting above the expected relationship (Figure 5). This deuterium excess measures between 7–46 per mil, with an average deuterium excess of 28 per mil. There is a significant difference in the estimated δ18O of meteoric water between the paleosol (average −17.8‰ VSMOW) and bentonite samples (−24.1‰ VSMOW). This could be the result of geographic, orographic, or geochronological differences between the paleosol and bentonite localities. Regardless, these oxygen isotope estimates of meteoric water fall within previously reported data from paleosol siderites, dinosaur tooth enamel phosphate, and pedogenic clay minerals [4,5], and both the bentonite and paleosol localities record deuterium excess despite the large shift in oxygen isotopic composition of meteoric water.
Previous studies using oxygen isotopes from pedogenic siderites, dinosaur tooth enamel phosphate, and pedogenic clay minerals to calculate δ18O of meteoric water and hydrogen isotopes from n-alkane biomarkers to calculate δD of meteoric water first reported an apparent deuterium excess of 40–60 per mil in the Cretaceous Prince Creek Formation [5]. However, the hydrogen isotopes measured from n-alkane biomarkers of plants living in an arctic environment with continuous photosynthetic activity and transpiration during the growing season could produce an apparent deuterium excess due to kinetic fractionation that preferentially concentrates deuterium (2H) over protium (1H) in the leaf tissue [5]. Although it appears that part of the deuterium excess reported from n-alkane biomarkers may be the result of photosynthetic and transpiration fraction components given the larger reported deuterium excess in Ludvigson et al. [5], this study has confirmed the deuterium excess by measuring both oxygen and hydrogen isotopes from phyllosilicates, thus removing any photosynthetic or transpiration fractionation components of Arctic vegetation.
With the exclusion of photosynthetic/transpiration fractionation, Ludvigson et al. [5] proposed two other possible processes that could contribute to the deuterium excess, the first being changes in the zonal patterns of evaporation and precipitation at lower latitudes. This includes relationships between D-excess based on relative humidity, precipitation recycling, seasonality, air mass trajectory, and others [49,50,51,52,53,54]. The second proposed process is the potential influence of hydrogenotrophic methanogenesis occurring in Cretaceous soils. Carbon isotopic data from paleosol siderites collected at localities throughout the Prince Creek Formation have δ13C values that are characteristic of environments in which hydrogenotrophic methanogenesis is the dominant anaerobic CH4-producing pathway [5,55,56]. During hydrogenotrophic methanogenesis, the preferential uptake of 1H can enrich the source water in deuterium without impacting the oxygen isotopic composition [57,58,59]. In order to evaluate the degree of influence of hydrogenotrophic methanogenesis on the deuterium excess, the next step would be to use previously collected data to quantitatively evaluate methane fluxes from Cretaceous Arctic soils. This would be further aided by additional phyllosilicate stable isotopic data from Cretaceous paleosols to determine the extent both temporally and geographically of the deuterium excess persisting during the Cretaceous and aid in the development of new climate models.

5. Conclusions

Our data confirm the presence of an apparent deuterium excess of Cretaceous Arctic paleoprecipitation first reported by Ludvigson et al. [5]. Results of stable oxygen and hydrogen isotopic analyses combined with elemental and chemical data of pedogenic clays reveal an average δ18O of paleoprecipitation of −20.3‰ VSMOW and average δD values of −122‰ VSMOW. The data produced from three paleosol samples and two bentonite samples reveal a deuterium excess ranging from 7 to 46 per mil with an average reported deuterium excess of 28 per mil. These results exclude any influence from photosynthetic and transpiration fractionation components in Arctic vegetation that could plague previous attempts to define the deuterium excess using n-alkane biomarker data. This leaves two processes that could explain the observed deuterium excess: (1) mass-balance changes in zonal patterns of evaporation and precipitation at lower latitudes, and/or (2) concentration of 2H in the groundwaters of methane-emitting Arctic soils. Further research is needed to constrain the influence of each of these processes and to understand the extent to which both temporally and geographically the deuterium excess was present during the Cretaceous and how this knowledge can aid future climate modeling.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/geosciences13090273/s1. Table S1 Raw X-ray diffraction data for <0.2 µm fraction of pedogenic phyllosilicates from locality 06SH and the <2 µm size fraction for bentonite samples from localities PFDV 17 and 06KKT. Table S2 Bond model equations used to calculate 2:1 end-member oxygen isotope fractionation for pedogenic phyllosilicates from locality 06SH.

Author Contributions

Conceptualization, K.A., G.L. and M.S.; methodology, K.A.; validation, K.A., G.L. and M.S.; formal analysis, K.A.; investigation, K.A., G.L., M.S., P.M. and P.F.; resources, K.A.; data curation, K.A. and P.M.; writing—original draft preparation, K.A., G.L., M.S., P.M. and P.F.; writing—review and editing, K.A., G.L., M.S., P.M. and P.F.; visualization, K.A., G.L., M.S. and P.F.; supervision, K.A.; project administration, K.A.; funding acquisition, K.A. and M.S. All authors have read and agreed to the published version of the manuscript.

Funding

Financial support for the collection of paleosol and bentonite samples used in this research was provided by the National Science Foundation Office of Polar Programs (OPP-0425636 to P.J.M.). Financial support for chemical and isotopic analyses of the paleosol and bentonite samples was provided by the Kansas Geological Survey (to K.A.A.) and KU Endowment (to M.S.).

Data Availability Statement

All new data reported in this article are included in Tables and Supplementary documents. Other data included in this analysis appeared in earlier publications.

Acknowledgments

We acknowledge the work and expertise of Fred Longstaffe at the Laboratory for Stable Isotope Science (LSIS), University of Western Ontario, London, Ontario for hydrogen isotope analyses of samples and Andrew Mott from the Texas A&M Materials and Characterization Facility for electron microprobe analyses conducted for this study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Barron, E.J. Warm, equable Cretaceous: The nature of the problem. Earth Sci. Rev. 1983, 19, 305–338. [Google Scholar] [CrossRef]
  2. Ruddiman, W.F. Earth’s Climate: Past and Future, 2nd ed.; W. H. Freeman and Company: New York, NY, USA, 2008. [Google Scholar]
  3. Poulsen, C.J. Palaeoclimate: A balmy Arctic. Nature 2004, 432, 814. [Google Scholar] [CrossRef] [PubMed]
  4. Suarez, C.A.; Ludvigson, G.A.; Gonzalez, L.; Fiorillo, A.; Flaig, P.; McCarthy, P.J. Use of Multiple Oxygen Isotope Proxies for Elucidating Arctic Cretaceous Palaeo-Hydrology. In Isotopic Studies in Cretaceous Research; Bojar, A.-V., Melinte-Dobrinescu, M.C., Smit, J., Eds.; Geological Society London, Special Publications: London, UK, 2013; Volume 382, pp. 185–202. [Google Scholar]
  5. Ludvigson, G.A.; Diefendorf, A.F.; Suarez, M.B.; González, L.A.; Corcoran, M.C.; Schlanser, K.; Flaig, P.P.; McCarthy, P.J.; van der Kolk, D.; Houseknecht, D.; et al. Stable Isotope Tracers of Cretaceous Arctic Paleoprecipitation. Geosciences 2022, 12, 143. [Google Scholar] [CrossRef]
  6. Moore, T.E.; Wallace, W.K.; Bird, K.J.; Karl, S.M.; Mull, C.G.; Dillon, J.T. Geology of Northern Alaska. In The Geology of Alaska: The Geology of North America; Plafker, G., Berg, H.C., Eds.; v. G-1; The Geological Society of America: Boulder, CO, USA, 1994; pp. 49–140. [Google Scholar]
  7. Cole, F.; Bird, K.J.; Toro, J.; Roure, F.; O’Sullivan, P.B.; Pawlewicz, M.; Howell, D.G. An integrated model for the tectonic development of the frontal Brooks Range and Colville Basin 250 km west of the Trans-Alaska Crustal Transect. J. Geophys. Res. 1997, 102, 20685–20708. [Google Scholar] [CrossRef]
  8. Mull, C.G.; Houseknecht, D.W.; Bird, K.J. Revised Cretaceous and Tertiary Stratigraphic Nomenclature in the Colville Basin, Northern Alaska; Professional Paper; United States Geological Survey: Reston, VA, USA, 2003; Volume 1673, pp. 1–51.
  9. Flaig, P.P.; McCarthy, P.J.; Fiorillo, A.R. A tidally influenced, high-latitude coastal plain: The Upper Cretaceous (Maastrichtian) Prince Creek Formation, North Slope, Alaska. In From River to Rock Record: The Preservation of Fluvial Sediments and Their Subsequent Interpretation; Davidson, S., Lelu, S., North, C., Eds.; SEPM Special Publication: Tulsa, Oklahoma, 2011; Volume 97, pp. 233–264. [Google Scholar]
  10. Flaig, P.P.; McCarthy, P.J.; Fiorillo, A.R. Anatomy, evolution and paleoenvironmental interpretation of an ancient Arctic coastal plain: Integrated paleopedology and palynology from the Upper Cretaceous (Maastrichtian) Prince Creek Formation, North Slope, Alaska. In New Frontiers in Paleopedology and Terrestrial Paleoclimatology; Driese, S.G., Nordt, L.D., Eds.; SEPM Special Publication: Tulsa, Oklahoma, 2013; Volume 104, pp. 179–230. [Google Scholar]
  11. Fiorillo, A.R.; McCarthy, P.J.; Flaig, P.P.; Brandlen, E.; Norton, D.; Jacobs, L.; Zippi, P.; Gangloff, R.A. Paleontology and Paleoenvironmental Interpretation of the Kikak-Tegoseak Dinosaur Quarry (Prince Creek Formation: Late Cretaceous), Northern Alaska: A Multi-Disciplinary Study of an Ancient, High-Latitude Ceratopsian Dinosaur Bonebed, In New Perspectives on Horned Dinosaurs. The Royal Tyrell Museum Ceraptopsian Symposium; Ryan, M.J., Chinner-Algeier, B.J., Eberth, D.A., Eds.; University Press: Bloomington, IN, USA, 2010; pp. 456–477. [Google Scholar]
  12. Brouwers, E.M.; Clemens, W.A.; Spicer, R.A.; Ager, T.A.; Carter, L.D.; Sliter, W.V. Dinosaurs on the North Slope, Alaska: High latitude latest Cretaceous environments. Science 1987, 25, 1608–1610. [Google Scholar] [CrossRef] [PubMed]
  13. Witte, W.K.; Stone, D.B.; Mull, C.G. Paleomagnetism, paleobotany, and paleogeography of the Cretaceous North Slope, Alaska. In Alaska North Slope Geology; Tailleur, I., Weimer, P., Eds.; The Pacific Section: Alameda, CA, USA; SEPM: Tulsa, Oklahoma; Alaska Geological Society: Anchorage, AK, USA, 1987; Volume 1, pp. 571–579. [Google Scholar]
  14. Salazar Jaramillo, S.; McCarthy, P.J.; Trainor, T.; Fowell, S.J.; Fiorillo, A.R. Origin of clay minerals in alluvial paleosols, Prince Creek Formation, North Slope, Alaska: Influence of volcanic ash on pedogenesis in the Late Cretaceous Arctic. J. Sediment. Res. 2015, 85, 192–208. [Google Scholar] [CrossRef]
  15. Rich, T.H.; Gangloff, R.A.; Hammer, W.H. Polar Dinosaurs. Science 2002, 295, 979–980. [Google Scholar] [CrossRef]
  16. Fiorillo, A.R.; McCarthy, P.J.; Kobayashi, Y.; Suarez, M.A. Cretaceous dinosaurs across Alaska show the role paleoclimate played in structuring large ancient herbivore populations. Geosciences 2022, 12, 161. [Google Scholar] [CrossRef]
  17. Stern, L.A.; Chamberlain, C.P.; Reynolds, R.C.; Johnson, G.D. Oxygen isotope evidence of climate change from pedogenic clay minerals in the Himalayan molasses. Geochim. Cosmochim. Acta 1997, 61, 731–744. [Google Scholar] [CrossRef]
  18. Tabor, N.J.; Montañez, I.P.; Southard, R.J. Paleoenvironmental reconstruction from chemical and isotopic compositions of Permo-Pennsylvanian pedogenic minerals. Geochim. Cosmochim. Acta 2002, 66, 3093–3107. [Google Scholar] [CrossRef]
  19. Vitali, F.; Longstaffe, F.J.; McCarthy Plint, P.J.; Guy, A.; Caldwell, W.; Glen, E. Stable isotopic investigation of clay minerals and pedogenesis in an interfluve paleosol from the Cenomanian Dunvegan Formation, N.E. British Columbia, Canada. Chem. Geol. 2002, 192, 269–287. [Google Scholar] [CrossRef]
  20. Tabor, N.J.; Montañez, I.P. Oxygen and hydrogen isotope compositions of pedo-genic phyllosilicates: Development of modern surface domain arrays and implications for paleotemperature reconstructions. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2005, 223, 127–146. [Google Scholar] [CrossRef]
  21. Myers, T.S.; Tabor, N.J.; Jacobs, L.L.; Mateus, O. Palaeoclimate of the Late Jurassic of Portugal: Comparison with the western United States. Sedimentology 2012, 59, 1695–1717. [Google Scholar] [CrossRef]
  22. Rosenau, N.A.; Tabor, N.J. Oxygen and hydrogen isotope compositions of paleosol phyllosilicates: Differential burial histories and determination of Middle-Late Pennsylvanian low-latitude terrestrial paleotemperatures. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2013, 392, 383–397. [Google Scholar] [CrossRef]
  23. Andrzejewski, K.A.; Tabor, N.J. Paleoenvironmental and paleoclimatic reconstruction of Cretaceous (Aptian-Cenomanian) terrestrial formations of Texas and Oklahoma using phyllosilicates. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2020, 543, 109491. [Google Scholar] [CrossRef]
  24. Savin, S.M.; Epstein, S. The oxygen and hydrogen isotope geochemistry of clay minerals. Goechim. Cosmochim. Acta 1970, 34, 25–42. [Google Scholar] [CrossRef]
  25. Lawrence, J.R.; Taylor, H.P. Deuterium and 18O correlation: Clay minerals and hydroxides in Quaternary soils compared to meteoric waters. Geochim. Cosmochim. Acta 1971, 35, 993–1003. [Google Scholar] [CrossRef]
  26. Moore, D.M.; Reynolds, R.C. X-ray Diffraction and the Identification and Analysis of Clay Minerals; Oxford University Press: New York, NY, USA, 1997; p. 378. [Google Scholar]
  27. Clayton, R.N.; Maydeda, T.K. The use of bromine pentafluoride in the extraction of oxygen from oxides and silicates for isotopic analysis. Geochim. Cosmochim. Acta 1963, 27, 43–52. [Google Scholar] [CrossRef]
  28. Gonfiantini, R. Advisory Group Meeting on Stable Isotope Reference Samples for Geochemical and Hydrological Investigations; Report International Atomic Energy Agency: Vienna, Austria, 1984. [Google Scholar]
  29. Salazar-Jaramillo, S.; McCarthy, P.J.; Ochoa, A.; Fowell, S.J.; Longstaffe, F.J. Data supporting Maastrichtian paleoclimate variables applying a multi proxy approach to a paleosol profile, Arctic Alaska. Data Brief 2020, 29, 105191. [Google Scholar] [CrossRef] [PubMed]
  30. Sharp, Z.D.; Atudorei, V.; Durakiewicz, T. A rapid method for determination of hydrogen and oxygen isotope ratios from water and hydrous minerals. Chem. Geol. 2001, 178, 197–210. [Google Scholar] [CrossRef]
  31. VanDeVelde, J.H.; Bowen, G.J. Effects of chemical pretreatments on the hydrogen isotope composition of 2:1 clay minerals. Rapid Commun. Mass Spectrom. 2013, 27, 1143–1148. [Google Scholar] [CrossRef] [PubMed]
  32. Bauer, K.K.; Vennemann, T.W. Analytical methods for the measurement of hydrogen isotope composition and water content in clay minerals by TC/EA. Chem. Geol. 2014, 363, 229–240. [Google Scholar] [CrossRef]
  33. Qi, H.; Coplen, T.B.; Gehre, M.; Vennemann, T.W.; Brand, W.A.; Geilmann, H.; Olack, G.; Bindeman, I.N.; Palandri, J.; Huang, L.; et al. New biotite and muscovite isotopic reference materials, USGS57 and USGS58, for δ2H measurements—A replacement for NBS 30. Chem. Geol. 2017, 467, 89–99. [Google Scholar] [CrossRef]
  34. Kanik, N.J.; Longstaffe, F.J.; Kuligiewicz, A.; Derkowski, A. Systematics of smectite hydrogen-isotope composition: Structural hydrogen versus absorbed water. Appl. Clay Sci. 2022, 216, 106338. [Google Scholar] [CrossRef]
  35. Coplen, T.B. New guidelines for reporting stable hydrogen, carbon, and oxygen isotope-ratio data. Geochim. Cosmochim. Acta 1996, 60, 3359–3360. [Google Scholar] [CrossRef]
  36. Savin, S.M.; Lee, M. Isotopic Studies of Phyllosilicates. Rev. Mineral. Geochem. 1988, 19, 189–223. [Google Scholar]
  37. Sheppard, S.M.F.; Gilg, H.A. Stable isotope geochemistry of clay minerals. Clay Miner. 1996, 31, 1–24. [Google Scholar] [CrossRef]
  38. Spicer, R.A.; Herman, A.B. The Late Cretaceous environment of the Arctic: A quantitative reassessment based on plant fossils. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2010, 295, 423–442. [Google Scholar] [CrossRef]
  39. Ludvigson, G.A.; González, L.A.; Fowle, D.A.; Roberts, J.A.; Driese, S.G.; Villarreal, M.A.; Smith, J.J.; Suarez, M.B. Paleoclimatic applications and modern process studies of pedogenic siderite. In New Frontiers in Paleopedology and Terrestrial Paleoclimatology; Driese, S.G., Nordt, L.C., McCarthy, P.J., Eds.; SEPM Special Publication: Broken Arrow, OK, USA, 2013; Volume 104, pp. 79–87. [Google Scholar]
  40. Gilg, H.A.; Sheppard, S.M.F. Hydrogen isotope fractionation between smectites and water. Terra Nova 1995, 7, 329. [Google Scholar]
  41. Rozanski, K.; Araguás-Araguás, L.; Gonfiantini, R. Isotopic patterns in modern global precipitation. Climate Change in Continental Isotopic Records. Am. Geophys. Monogr. 1993, 78, 1–36. [Google Scholar]
  42. Burns, W.M.; Hayba, D.O.; Rowan, E.L.; Houseknecht, D.W. Estimating the amount of eroded section in a partially exhumed basin from geophysical well logs: An example from the North Slope. In Studies by the U.S. Geological Survey in Alaska; Special Paper; Haeussler, P.J., Galloway, J.P., Eds.; U.S. Geological Survey: Reston, VA, USA, 2005; Volume 1732, pp. 1–18. [Google Scholar]
  43. Robinson, M.S. Kerogen Microscopy of Coal and Shales from the North Slope of Alaska; Public Data File; Alaska Division of Geological and Geophyscial Surveys: Fairbanks, AK, USA, 1989; Volume 89. [Google Scholar]
  44. Johnson, M.J.; Howell, D.G. Thermal Maturity of Sedimentary Basins in Alaska: An Overview; U.S. Geological Survey: Reston, VA, USA, 1996; Volume 2142, pp. 1–131.
  45. Barker, C.E.; Pawlewicz, M.J. The correlation of vitrinite reflectance with maximum paleotemperature in humic organic matter. In Paleogeothermics; Springer: New York, NY, USA, 1986; pp. 79–93. [Google Scholar]
  46. Meunier, A. Clays; Springer: Berlin/Heidelberg, Germany, 2005; pp. 1–472. [Google Scholar]
  47. Craig, H. Isotopic variations in meteoric waters. Science 1961, 133, 1702–1708. [Google Scholar] [CrossRef] [PubMed]
  48. Yapp, C.J. Climatic implications of surface domains in arrays of δD and δ18O from hydroxyl minerals: Goethite as an example. Geochim. Cosmichim. Acta 2000, 64, 2009–2025. [Google Scholar] [CrossRef]
  49. Dansgaard, W. Stable isotopes in precipitation. Tellus 1964, 16, 436–468. [Google Scholar] [CrossRef]
  50. Bowen, G.J.; Cai, Z.; Fiorella, R.P.; Putman, A.L. Isotopes in the Water Cycle: Regional- to Global-Scale Patterns and Applications. Annu. Rev. Earth Planet. Sci. 2019, 47, 453–479. [Google Scholar] [CrossRef]
  51. Craig, H.; Gordon, L.I. Deuterium and Oxygen-18 Variations in the Ocean and the Marine Atmosphere. In Proceedings of the Conference on Stable Isotopes in Oceanographic Studies and Paleotemperatures, Pisa, Italy, 26–30 July 1965; Tongiorgi, E., Ed.; pp. 9–130. [Google Scholar]
  52. Gat, J.R.; Bowser, C.J.; Kendall, C. The contribution of evaporation from the Great Lakes to the continental atmosphere: Estimate based on stable isotope data. Geophys. Res. Lett. 1994, 21, 557–560. [Google Scholar] [CrossRef]
  53. Good, S.P.; Mallia, D.V.; Lin, J.C.; Bowen, G.J. Stable isotope analysis of precipitation samples obtained via crowdsourcing reveals the spatiotemporal evolution of Superstorm Sandy. PLoS ONE 2014, 9, e91117. [Google Scholar] [CrossRef]
  54. Cropper, S.; Solander, K.; Newman, B.D.; Tuinenburg, O.A.; Staal, A.; Theeuwen, J.J.E.; Xu, C. Comparing deuterium excess to large-scale precipitation recycling models in the tropics npj. Clim. Atmos. Sci. 2021, 4, 60. [Google Scholar] [CrossRef]
  55. Talbot, M.R.; Kelts, K. Primary and diagenetic carbonates in the anoxic sediments of Lake Bosumtwi, Ghana. Geology 1986, 14, 912–916. [Google Scholar] [CrossRef]
  56. Mozley, P.S.; Wersin, P. Isotopic composition of siderite as an indicator of depositional environment. Geology 1992, 20, 817–820. [Google Scholar] [CrossRef]
  57. North, J.C.; Frew, R.D.; Van Hale, R. Can stable isotopes be used to monitor landfill leachate impact on surface waters? J. Geophys. Explor. 2006, 88, 49–53. [Google Scholar] [CrossRef]
  58. Andrei, F.; Barbieri, M.; Sappa, G. Application of 2H and 18O isotopes for tracing municipal solid waste landfill contamination of groundwater: Two Italian case histories. Water 2021, 13, 1065. [Google Scholar] [CrossRef]
  59. Castañeda, S.; Sucgang, R.; Almoneda, R.; Mendoza, N.; David, C. Environmental isotopes and major ions for tracing leachate contamination from a municipal landfill in Metro Manila, Philippines. J. Environ. Radioact. 2012, 110, 20–37. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Sample locations from the Prince Creek Formation along the Colville and Kogosukruk Rivers in northern Alaska.
Figure 1. Sample locations from the Prince Creek Formation along the Colville and Kogosukruk Rivers in northern Alaska.
Geosciences 13 00273 g001
Figure 2. Measured stratigraphy including all stratigraphic features and sample horizons from the Prince Creek Formation in northern Alaska. Figure includes the stratigraphic location of samples analyzed for this study, indicated by blue stars. (A) Representative stratigraphic column and paleoenvironmental interpretations for the PFDV 17 sample locality along the Kogosukruk River. (B) Representative stratigraphic column and paleoenvironmental interpretations for the 06KKT sample locality along the Colville River. (C) Representative stratigraphic column and paleoenvironmental interpretations for the 06SH sample locality along the Colville River.
Figure 2. Measured stratigraphy including all stratigraphic features and sample horizons from the Prince Creek Formation in northern Alaska. Figure includes the stratigraphic location of samples analyzed for this study, indicated by blue stars. (A) Representative stratigraphic column and paleoenvironmental interpretations for the PFDV 17 sample locality along the Kogosukruk River. (B) Representative stratigraphic column and paleoenvironmental interpretations for the 06KKT sample locality along the Colville River. (C) Representative stratigraphic column and paleoenvironmental interpretations for the 06SH sample locality along the Colville River.
Geosciences 13 00273 g002
Figure 3. Outcrop imagery of locality PFDV 17 located along the Kogosukruk River showing the sedimentologic, stratigraphic and stratal architecture framework of the site. Figure includes part of the PFDV 17 stratigraphic chart. Outlined red box and green arrow show the location of the sampled bentonite layer along the outcrop belt and a closeup image of the bentonite. Pocketknife is 9 cm for scale.
Figure 3. Outcrop imagery of locality PFDV 17 located along the Kogosukruk River showing the sedimentologic, stratigraphic and stratal architecture framework of the site. Figure includes part of the PFDV 17 stratigraphic chart. Outlined red box and green arrow show the location of the sampled bentonite layer along the outcrop belt and a closeup image of the bentonite. Pocketknife is 9 cm for scale.
Geosciences 13 00273 g003
Figure 4. X-ray diffraction of glycerol-solvated clays in the <0.2 µm size fraction from paleosols sampled at site 06 SH in the Prince Creek Formation, shown in stratigraphic order. Spectra show the samples are dominated by smectite (001; 18.5 Å) with smaller amounts of illite/mica (10 Å) and kaolinite (001; 7.2 Å).
Figure 4. X-ray diffraction of glycerol-solvated clays in the <0.2 µm size fraction from paleosols sampled at site 06 SH in the Prince Creek Formation, shown in stratigraphic order. Spectra show the samples are dominated by smectite (001; 18.5 Å) with smaller amounts of illite/mica (10 Å) and kaolinite (001; 7.2 Å).
Geosciences 13 00273 g004
Table 1. Electron microprobe and X-ray diffraction data for the <0.2 µm-size fraction of phyllosilicates for paleosol samples (this study) and the <2 µm-size fraction of bentonite samples [14]. Weight percentage (wt. %) of clay mineralogy was calculated using X-ray diffraction data, and results reported have an analytical error of ±5%.
Table 1. Electron microprobe and X-ray diffraction data for the <0.2 µm-size fraction of phyllosilicates for paleosol samples (this study) and the <2 µm-size fraction of bentonite samples [14]. Weight percentage (wt. %) of clay mineralogy was calculated using X-ray diffraction data, and results reported have an analytical error of ±5%.
SampleFormationNa2OMgOMnOAl2O3SiO2K2OCaOFe2O3TiO2Total2:1 Phyllosilicate (wt.%)Kaolinite (wt.%)
Paleosol
06 SH 9.2Prince Creek2.511.680.0124.5756.992.560.035.561.0494.957723
06 SH 14.4Prince Creek2.281.670.0223.4057.392.500.036.511.0194.818515
06 SH 15.1Prince Creek2.012.180.0424.2957.233.210.115.251.0995.417525
Bentonite
* PFDV 17Prince Creek2.241.820.0219.3166.960.491.374.710.1997.111000
** 06 KKT-20.5Prince Creek0.191.840.0120.6760.780.071.916.040.2191.721000
* Geochemical analyses produced from XRF data reported in Salazar-Jaramillo et al. [14] ** Geochemical analyses produced from microprobe data reported in Salazar-Jaramillo et al. [14].
Table 2. Calculated chemical formulae for end-member 2:1 phyllosilicate mixtures.
Table 2. Calculated chemical formulae for end-member 2:1 phyllosilicate mixtures.
SampleFormationChemical Formulae
Paleosol
06 SH 9.2Prince Creek(K0.26Na0.39)(Al1.37Fe0.33Mg0.20Ti0.06)(Si3.61Al0.39)O10(OH)2
06 SH 14.4Prince Creek(K0.23Na0.32)(Al1.37Fe0.36Mg0.18Ti0.06)(Si3.67Al0.33)O10(OH)2
06 SH 15.1Prince Creek(K0.33Na0.31Ca0.01)(Al1.33Fe0.32Mg0.26Ti0.07)(Si3.61Al0.39)O10(OH)2
Chemical formulae were calculated from the oxide data reported in Table 1, based on the presence of twelve oxygen atoms and two hydrogen atoms in each unit-cell [26].
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Andrzejewski, K.; Ludvigson, G.; Suarez, M.; McCarthy, P.; Flaig, P. Exploring the Deuterium Excess of Cretaceous Arctic Paleoprecipitation Using Stable Isotope Composition of Clay Minerals from the Prince Creek Formation (Maastrichtian) in Northern Alaska. Geosciences 2023, 13, 273. https://doi.org/10.3390/geosciences13090273

AMA Style

Andrzejewski K, Ludvigson G, Suarez M, McCarthy P, Flaig P. Exploring the Deuterium Excess of Cretaceous Arctic Paleoprecipitation Using Stable Isotope Composition of Clay Minerals from the Prince Creek Formation (Maastrichtian) in Northern Alaska. Geosciences. 2023; 13(9):273. https://doi.org/10.3390/geosciences13090273

Chicago/Turabian Style

Andrzejewski, Kate, Greg Ludvigson, Marina Suarez, Paul McCarthy, and Peter Flaig. 2023. "Exploring the Deuterium Excess of Cretaceous Arctic Paleoprecipitation Using Stable Isotope Composition of Clay Minerals from the Prince Creek Formation (Maastrichtian) in Northern Alaska" Geosciences 13, no. 9: 273. https://doi.org/10.3390/geosciences13090273

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop