Skip to content
BY 4.0 license Open Access Published by De Gruyter Open Access December 2, 2022

Milankovitch cycles and the astronomical time scale of the Zhujiang Formation in the Baiyun Sag, Pearl River Mouth Basin, China

  • Ping He ORCID logo , Shangfeng Zhang ORCID logo EMAIL logo , Enze Xu ORCID logo and Chunxia Zhu ORCID logo
From the journal Open Geosciences

Abstract

The Zhujiang Formation in the Baiyun Sag, Pearl River Mouth Basin, China, is formed primarily in a deep-water continental slope environment. Its chronostratigraphic framework is based on biostratigraphy and sequence stratigraphy, and its geological dating is based on micropaleontological data. This makes it difficult to obtain precise absolute ages for various geological events. In this study, gamma ray (GR) well log data from Wells Y1, Y2, and Y3 were used as paleoclimate proxies, and spectral and wavelet analyses were used to conduct cyclostratigraphic research. The results show that the Milankovitch cycles were preserved in the Zhujiang Formation in the Baiyun Sag. Stratigraphic cycles controlled by 405 and 95 ka orbital eccentricity, 40.4 ka orbital obliquity, and 23.5 ka orbital precession cycles can be identified; the signal of stratigraphic cycles controlled by the 405 ka long eccentricity cycle is the strongest. The floating astronomical time scale is constructed based on 405 ka orbital eccentricity cycle tuning of the GR series. The precise durations of the Zhujiang Formation in Wells Y1, Y2, and Y3 are 7.13, 6.93, and 7.18 Ma, and the average deposition rates are 4.68, 5.91, and 5.33 cm/ka, respectively. The Zhujiang Formation was divided into 17 fourth-, 76 fifth-, and 174 sixth-order cycles using the 405, 95, and 40.4 ka orbital periods as the dividing scales, respectively. This study provides a quantitative method for high-precision isochronous stratigraphic division and correlation in deep-water sedimentary systems.

1 Introduction

The Milankovitch theory is an astronomical theory involving the relationship between insolation and Earth’s climate. The theory states that periodic changes in Earth’s orbital parameters (eccentricity, obliquity, and precession) can cause periodic changes in insolation, which lead to climate changes and corresponding changes in sedimentary characteristics and the environment [1]. The strata controlled by Earth’s orbital cycles record the Milankovitch cycles. By astronomical tuning of these cycles, a continuous astronomical time scale (ATS) with a resolution of 20–400 ka can be obtained in cyclostratigraphy, which can accurately determine the geological ages and durations of various geological and biological events and realize stratigraphic division and correlation with high precision [24]. At present, cyclostratigraphic research results have been obtained from Paleozoic and older strata [58] as well as from Mesozoic and Cenozoic strata [912]. These strata range from marine to continental [1316].

The Pearl River Mouth Basin is a Cenozoic petroliferous basin. Few paleomagnetic and isotopic dating data are available for the Neogene. Most researchers have established a chronostratigraphic framework through biostratigraphic and sequence stratigraphic studies [1722]. The age of the Neogene in the Pearl River Mouth Basin is primarily based on micropaleontological data [1921]. The age framework was discontinuous and inaccurate because paleontological samples were obtained from drill cuttings. The well log data can be used as a paleoclimate proxy to perform cyclostratigraphy studies, which can be used to obtain high-resolution geological age scales. Liu et al. and Tian et al. conducted related cyclostratigraphy studies using gamma ray (GR) well log data in the Huizhou Sag of the Pearl River Mouth Basin [2325].

The Baiyun Sag of the Pearl River Mouth Basin has abundant oil and gas resources, and the Zhujiang Formation is an important exploration target layer. The Zhujiang Formation of the Baiyun Sag was primarily formed in a deep-water continental slope environment, and it is difficult to divide and correlate the strata. In this study, we conducted cyclostratigraphy research of the Zhujiang Formation in the Baiyun Sag to establish a high-precision chronostratigraphic framework. Considering that factors such as hiatuses and lateral changes in strata thickness and sedimentary facies led to incomplete records of Milankovitch cycles, we selected three horizontally contrastable drilled wells (Y1, Y2, and Y3) for a comparative study to identify the Milankovitch cycles. We established a high-resolution ATS of the Zhujiang Formation in the Baiyun Sag by astronomical tuning to obtain the sedimentation time and rate. Different Milankovitch cycles were selected as scales to divide high-frequency cycles. This study provides a quantitative method for high-precision isochronous stratigraphic division of deep-water sedimentary systems.

2 Geological background

The Pearl River Mouth Basin is a Cenozoic sedimentary basin located in the northern South China Sea at 18°30′–23°30′N and 113°10′–118°00′E (Figure 1a). It is a typical passive continental margin basin. Its tectonic and sedimentary evolution can be divided into a rift valley stage from the Paleocene to the Eocene, a depression stage from the Oligocene to the Middle Miocene, and a stable subsidence stage from the Late Miocene to the present [26]. The tectonic units are distributed in a southwest to northeast direction, and from north to south, the northern uplift belt, northern depression belt, central uplift belt, central depression belt, southern uplift belt, and southern depression belt occur successively [26]. The Baiyun Sag is located in the eastern part of the central depression belt and covers an area of approximately 25,500 km2 (Figure 1b). It is the largest sedimentary depression in the Pearl River Mouth Basin and contains extremely thick deposits of fluvio-lacustrine, marine-terrestrial transitional, and shallow-deep sea facies.

Figure 1 
               Geologic setting of the Pearl River Mouth Basin, China: (a) modern map showing the location of the Pearl River Mouth Basin, China and (b) tectonic units of the Pearl River Mouth Basin and locations of the Baiyun Sag and Wells Y1, Y2, and Y3.
Figure 1

Geologic setting of the Pearl River Mouth Basin, China: (a) modern map showing the location of the Pearl River Mouth Basin, China and (b) tectonic units of the Pearl River Mouth Basin and locations of the Baiyun Sag and Wells Y1, Y2, and Y3.

The Cenozoic strata in the Baiyun Sag, from bottom to top, are the Paleogene Shenhu Formation, Wenchang Formation, Enping Formation, and Zhuhai Formation; the Neogene Zhujiang Formation, Hanjiang Formation, Yuehai Formation, and Wanshan Formation; and the Quaternary strata. The Zhujiang Formation is composed primarily of deep-water continental slope deposits (Figure 2). The lower member is composed of interbeds of light gray sandstone, gray mudstone, and silty mudstone with thin layers of limestone. The upper member is thick gray and silty mudstones interbedded with light gray sandstone and siltstone. In the seismic reflection profile, the Zhujiang Formation is located between T40 and T60. Four calcareous nannofossil zones (NN1–NN4) and five planktonic foraminiferal zones (N4–N8) were found in the Zhujiang Formation [1921,27]. In the study area, the last downhole occurrence of planktonic foraminifera Globigerina ciperoensis (age 22.9 Ma) was used as the marker of the Paleogene/Neogene boundary [19], and the last downhole occurrence of the calcareous nannofossil Helicosphaera ampliaperta (age 15.87 Ma) was used as the marker of the Lower and Middle Miocene boundary [21]. Therefore, the Zhujiang Formation was formed in the Early Miocene.

Figure 2 
               Comprehensive histogram of the Zhujiang Formation in the Baiyun Sag, Pearl River Mouth Basin.
Figure 2

Comprehensive histogram of the Zhujiang Formation in the Baiyun Sag, Pearl River Mouth Basin.

Based on seismic, logging, core, and paleontological data, the Zhujiang Formation in the Baiyun Sag was divided into five third-order sequences (SQ1–SQ5) (Figure 2). During the sequence SQ1, the northern section of the Baiyun Sag accumulated continental shelf marginal delta deposits, and the middle and southern sections accumulated neritic shelf deposits. During the sequence SQ2, the northern section of the Baiyun Sag was composed of delta and neritic shelf deposits, and the middle and southern sections formed in a deep-water continental slope environment, with deep-water fan, gravity flow, and bathyal mudstone deposits. Compared with those in SQ1, the area of delta deposits decreased, and the area of neritic shelf deposits increased during SQ2, reflecting an obvious transgression process. During the SQ3 development period, the rate of relative sea-level fall was small, and the provenance supply was insufficient. The shelf marginal delta deposition did not develop. The northern section of the Baiyun Sag contained neritic shelf deposits, while the middle and southern sections contained bathyal deposits. During the SQ4 development period, the facies belt changed little, consisting of neritic and bathyal mudstone deposits, reflecting a relatively stable or slow fall in sea level. During SQ5, deltas developed on the northern section of the Baiyun Sag, and the middle and southern sections accumulated neritic and bathyal mudstone deposits.

3 Data and methods

3.1 GR well log data

The GR well log data can indicate variation in shale and organic matter content in the strata, which is controlled by the sedimentary environment and sea level change [28,29]. Therefore, GR data can be used as a good paleoclimate proxy [30]. The GR data were obtained from Wells Y1, Y2, and Y3 in the southern Baiyun Sag. The Zhujiang Formation of the three wells is a neritic or bathyal facies with complete sedimentary records and no obvious hiatus. The GR data of the three wells were equally spaced with a spacing of 0.1524 m. Although the GR data can reflect the periodic changes in strata caused by various geological factors, the data also contain signals independent of geological factors. Therefore, the data should be preprocessed to eliminate these signals. We used the following two steps to process the data: first, outliers that were much greater than or less than the mean value were eliminated and replaced with the mean value; then, the data were detrended to eliminate low-frequency background noise.

3.2 Time series analysis

To accurately identify the main stratigraphic cycles, a method combining spectral and wavelet analyses was adopted, and the results of the three wells were compared to determine the stratigraphic cycles. Both spectral and wavelet analyses were performed using the MATLAB software. Multitaper method (MTM) spectral analysis was used to transform the data from the depth domain to the frequency domain. The abscissa in the spectrum diagram represents the frequency, the reciprocal of which is the thickness of the corresponding stratigraphic cycle. The ordinate in the spectrum diagram is the relative power, and a larger value leads to the specific sedimentary cycle appearing more frequently. We chose spectral peaks above 95% confidence and selectively used spectral peaks between 90 and 95% confidence. A continuous wavelet transform was used in the wavelet analysis. Wavelengths with high power and relative continuity in the wavelet spectrum are the thicknesses of dominant stratigraphic cycles. The digital filtering method was used to obtain the stratigraphic cycles of different periods using a Gaussian bandpass filter on the preprocessed GR data. The La2010d orbital solution was established for 15.97–23.03 Ma as an accurate astronomical target [31].

4 Results

4.1 Theoretical orbital cycles

Based on the Geologic Time Scale 2012 (GTS2012) [32], the age of the Lower Miocene is 15.97–23.03 Ma. MTM spectral analysis of the standardized eccentricity, tilt and precession (ETP), which is the sum of the normalized eccentricity, normalized obliquity, and negative normalized precession, was performed to obtain the theoretical orbital cycles for this period (Figure 3a). Spectral analysis shows that the theoretical orbital cycles are the 1,000 ka ultralong, 405 ka long, and 125 and 95 ka short eccentricity cycles; 53.0, 40.4, and 39.4 ka obliquity cycles; and 23.5, 22.2, and 18.9 ka precession cycles (Figure 3b). The main theoretical orbital cycles of the Early Miocene are the 405 and 95 ka eccentricity, 40.4 ka obliquity, and 23.5 ka precession cycles. The proportion of the main theoretical orbital cycles was used as a benchmark for the analysis of the Milankovitch cycles.

Figure 3 
                  Theoretical orbital cycles for the period of 15.97–23.03 Ma: (a) ETP for the period of 15.97–23.03 Ma in the La2010d orbital solution and (b) MTM power spectrum of the ETP in (a).
Figure 3

Theoretical orbital cycles for the period of 15.97–23.03 Ma: (a) ETP for the period of 15.97–23.03 Ma in the La2010d orbital solution and (b) MTM power spectrum of the ETP in (a).

4.2 Determination of Milankovitch cycles

Milankovitch cycles were determined by comparing the thicknesses of the dominant stratigraphic cycles with theoretical orbital cycles. If the ratio of the dominant stratigraphic cycle thickness is the same as the ratio of the theoretical orbital cycles, it indicates that the stratigraphic cycles are driven by astronomical orbital cycles, and the Milankovitch cycles are preserved in the strata [33,34]. Based on the strata thickness and paleontological ages, the average sedimentation rates of the Zhujiang Formation in Wells Y1, Y2, and Y3 were estimated to be 4.7, 5.8, and 5.4 cm/ka, respectively, which were used as the constraint conditions for determining the Milankovitch cycles.

The depth of the Zhujiang Formation in Well Y1 is 2869.1–3202.8 m. There are significant spectral peaks at 17.94, 7.94, 4.20, 3.28, 2.11, 1.99, 1.78, 1.70, 1.55, and 1.05 m. The ratios of 17.94, 4.20, 1.78, and 1.05 m are fairly consistent with the ratios of the 405, 95, 40.4, and 23.5 ka cycles (Figure 4a, Table 1). The wavelet spectrum shows that the thicknesses of the dominant stratigraphic cycles are consistent with the results of spectral analysis (Figure 4b), suggesting that the deposition of the Zhujiang Formation is controlled by astronomical orbital cycles. Thus, based on the estimated average sedimentation rate of 4.7 cm/ka from biostratigraphic data, the 17.94 and 4.20 m cycles correspond to the 405 and 95 ka eccentricity cycles, the 1.78 m cycle correspond to the 40.4 ka obliquity cycle, and the 1.05 m cycle correspond to the 23.5 ka precession cycle. From the spectral diagram, the amplitude of the 17.94 m cycle is the largest, indicating that the Zhujiang Formation was primarily controlled by the 405 ka long eccentricity cycle.

Figure 4 
                  Spectral and wavelet analyses of the GR series from the Zhujiang Formation in the Baiyun Sag: (a and b) MTM power and wavelet spectra of the GR series from 2869.1 to 3202.8 m in Well Y1, (c and d) from 2553.12 to 2,963 m in Well Y2, and (e and f) from 2,745 to 3,128 m in Well Y3.
Figure 4

Spectral and wavelet analyses of the GR series from the Zhujiang Formation in the Baiyun Sag: (a and b) MTM power and wavelet spectra of the GR series from 2869.1 to 3202.8 m in Well Y1, (c and d) from 2553.12 to 2,963 m in Well Y2, and (e and f) from 2,745 to 3,128 m in Well Y3.

Table 1

Stratigraphic cycle thicknesses and corresponding astronomical orbital cycles

Well Depth (m) Frequency Cycle thickness (m) Cycle thickness ratio Orbital cycles (ka) Orbital cycles ratio
Y1 2869.1–3202.8 0.05575 17.94 17.09 405 17.23
0.2380 4.20 4.00 95 4.04
0.5629 1.78 1.70 40.4 1.72
0.9520 1.05 1.00 23.5 1.00
Y2 2553.12–2,963 0.0400 25.00 16.03 405 17.23
0.1215 8.23 5.28 125 5.32
0.1600 6.25 4.01 95 4.04
0.3961 2.52 1.62 40.4 1.72
0.6410 1.56 1.00 23.5 1.00
Y3 2,745–3,128 0.0455 21.98 17.87 405 17.23
0.1996 5.01 4.07 95 4.04
0.4549 2.20 1.79 40.4 1.72
0.8147 1.23 1.00 23.5 1.00

The results of spectral and wavelet analyses of the GR data from the Zhujiang Formation in Well Y2 show that the ratios of the dominant cycle thicknesses of 25.0, 8.23, 6.25, 2.52, and 1.56 m are consistent with the ratios of the orbital cycles of 405, 125, 95, 40.4, and 23.5 ka (Figure 4c and d, Table 1). Thus, based on the estimated average sedimentation rate of 5.8 cm/ka, the 25.0, 8.23, 6.25, 2.52, and 1.56 m cycles correspond to the 405, 125, 95, 40.4, and 23.5 ka orbital cycles, respectively.

Spectral and wavelet analyses of the GR data from the Zhujiang Formation in Well Y3 show that the dominant stratigraphic cycle thicknesses are 21.98, 5.01, 4.07, 3.09, 2.20, and 1.23 m, and the ratios of 21.98, 5.01, 2.20, and 1.23 m are similar to the ratios of the orbital cycles of 405, 95, 40.4, and 23.5 ka (Figure 4e and f, Table 1). Based on the estimated average sedimentation rate of 5.4 cm/ka from biostratigraphic data, the stratigraphic cycles of 21.98, 5.01, 2.20, and 1.23 m are controlled by the 405 ka long eccentricity, 95 ka short eccentricity, 40.4 ka obliquity, and 23.5 ka precession cycles. The signal of the 21.98 m cycle is the strongest.

A comparison of Wells Y1, Y2, and Y3 shows that the Zhujiang Formation has preserved sedimentary cycles controlled by the 405 ka long eccentricity, 95 ka short eccentricity, 40.4 ka obliquity, and 23.5 ka precession cycles (Table 1), and the signals of the stratigraphic cycles controlled by the 405 ka long eccentricity cycles are the strongest. Therefore, the Zhujiang Formation deposition in the Baiyun Sag was controlled by Earth’s orbital cycles.

4.3 The floating ATS

The 405 ka long eccentricity cycle has been relatively stable in the Cenozoic [35,36], and they are the main cyclic factors controlling the Zhujiang Formation in the Baiyun Sag. Therefore, the 405 ka long eccentricity cycle was selected as the target curve for astronomical tuning to construct the floating ATS. Eccentricity affects the insolation of the entire Earth, and ice ages develop during the orbital eccentricity minima [37]. High GR values correspond to layers with high shale and organic contents, reflecting warm and humid climatic conditions. Thus, high GR values corresponded to high eccentricity values when performing astronomical tuning.

We bandpass filtered the 17.94 m cycles of GR data from the Zhujiang Formation in Well Y1 using a Gaussian bandpass filter (passband: 0.05575 ± 0.0035) and recognized seventeen 17.94 m cycles, expressed as E1–E17 (Figure 5a). Tuning the 17.94 m cycles to the 405 ka orbital cycles, a 7.13 Ma long floating ATS was constructed (Figure 5b). Therefore, the average sedimentation rate of the Zhujiang Formation in Well Y1 was 4.68 cm/ka. The MTM spectral analysis of the 405 ka-tuned GR time series of the Zhujiang Formation shows peaks with confidence greater than 95% at 405, 173, 95, 53, 47, 41, 39.4, 23.5, and 22.2 ka, and the 405 ka cycle peak is the maximum (Figure 6a), which is consistent with the theoretical orbital cycles, indicating that the floating ATS is reliable.

Figure 5 
                  The floating ATS of the Zhujiang Formation in the Baiyun Sag. Preprocessed GR series for: (a) 2869.1–3202.8 m in Well Y1 with 17.94 m filtered output curves (passband: 0.05575 ± 0.0035), (c) 2553.12–2,963 m in Well Y2 with 25.0 m filtered output curves (passband: 0.04 ± 0.003), (e) 2,745–3,128 m in Well Y3 with 21.98 m filtered output curves (passband: 0.0455 ± 0.01), and (b, d, f) 405 ka-tuned GR series of Wells Y1, Y2, and Y3 with 405 ka filtered output curves (passband: 0.002469 ± 0.0001).
Figure 5

The floating ATS of the Zhujiang Formation in the Baiyun Sag. Preprocessed GR series for: (a) 2869.1–3202.8 m in Well Y1 with 17.94 m filtered output curves (passband: 0.05575 ± 0.0035), (c) 2553.12–2,963 m in Well Y2 with 25.0 m filtered output curves (passband: 0.04 ± 0.003), (e) 2,745–3,128 m in Well Y3 with 21.98 m filtered output curves (passband: 0.0455 ± 0.01), and (b, d, f) 405 ka-tuned GR series of Wells Y1, Y2, and Y3 with 405 ka filtered output curves (passband: 0.002469 ± 0.0001).

Figure 6 
                  Spectral analysis of the 405 ka-tuned GR series. MTM power spectrum of the 405 ka-tuned GR series in (a) Figure 5b, (b) Figure 5d, and (c) Figure 5f.
Figure 6

Spectral analysis of the 405 ka-tuned GR series. MTM power spectrum of the 405 ka-tuned GR series in (a) Figure 5b, (b) Figure 5d, and (c) Figure 5f.

Similarly, a 6.93 Ma long ATS was constructed for the Zhujiang Formation in Well Y2 based on tuning the 25.0 m cycles to the 405 ka orbital cycles (Figure 5c and d), with an average sedimentation rate of 5.91 cm/ka. A 7.18 Ma long ATS was constructed for the Zhujiang Formation in Well Y3 based on tuning the 21.98 m cycles to the 405 ka orbital cycles (Figure 5e and f), with an average sedimentation rate of 5.33 cm/ka. The MTM spectral analysis of the 405 ka-tuned GR time series of the Zhujiang Formation in these two wells shows significant peaks at 405, 125, ∼100, 40.4, 39.4, 23.5, and 22.2 ka, which are consistent with the theoretical orbital cycles (Figure 6b and c).

5 Discussion

5.1 The Zhujiang Formation age model

Based on the floating ATS, the sedimentary durations of the Zhujiang Formation in Wells Y1, Y2, and Y3 were 7.13, 6.93, and 7.18 Ma, respectively. These results are consistent with the biostratigraphic data. The maximum difference is 250 ka, which is a reasonable uncertainty and less than the error of one 405 ka orbital cycle for astronomical calibration of the geologic time scale. This shows that the floating ATS of the three wells can be used for high-precision strata division and correlation in the Baiyun Sag.

The 405 ka-tuned GR time series provides a 7.18 Ma long floating ATS for the Zhujiang Formation in Well Y3. Considering the base of the Zhujiang Formation at 3,128 m in Well Y3 as the Oligocene/Miocene boundary and anchored at 23.03 Ma based on the GTS2012 (Figure 7), the ATS age at the top of the Zhujiang Formation is 15.85 Ma, which is similar to the biostratigraphic age of 15.87 and 0.12 Ma later than the GTS2012 assigned age of 15.97 Ma for the boundary of the Early and Middle Miocene [32]. By comparing the seventeen 21.98 m cycles (E1–E17) of the GR series with the 41–57th 405 ka long eccentricity cycles (E40541–E40557) of the La2010d astronomical solution, a corresponding relationship was revealed (Figure 7), indicating that the ATS has high credibility.

Figure 7 
                  The ATS and chronostratigraphic framework of the Zhujiang Formation in the Baiyun Sag. (From left to right, the graph on the right side of the figure shows the 405 ka-tuned GR series of the Zhujiang Formation in Well Y3; the 405 ka filtered output curves of the 405 ka-tuned GR series [passband: 0.002469 ± 0.0001]; the calculated sedimentation rate of the Zhujiang Formation in Well Y3; and the La2010d solution eccentricity curve from 15.85 to 23.03 Ma and the 405 ka filtered output curve.)
Figure 7

The ATS and chronostratigraphic framework of the Zhujiang Formation in the Baiyun Sag. (From left to right, the graph on the right side of the figure shows the 405 ka-tuned GR series of the Zhujiang Formation in Well Y3; the 405 ka filtered output curves of the 405 ka-tuned GR series [passband: 0.002469 ± 0.0001]; the calculated sedimentation rate of the Zhujiang Formation in Well Y3; and the La2010d solution eccentricity curve from 15.85 to 23.03 Ma and the 405 ka filtered output curve.)

5.2 Establishment of chronostratigraphic framework

ATS is continuous and has high precision; thus, it can be used to accurately date geological and biological events. The ATS ages for the tops of third-order sequences SQ1–SQ5 are 22.6, 22.1, 18.56, 16.89, and 15.85 Ma, respectively (Figure 7).

The abundance of planktonic foraminifera and calcareous nannofossils in the Zhujiang Formation of the Baiyun Sag was high. The first occurrence (FO) surface of planktonic foraminifera Globorotalia sicanus and the last occurrence (LO) surface of planktonic foraminifera Catapsydrax dissimilis, Globoquadrina binaiensis, and G. ciperoensis were found in the Zhujiang Formation in Well Y3 at depths of 2,788, 2,838, 2,898, and 3,124 m, respectively. According to the ATS, the four biological events are dated at 16.53, 17.39, 18.56, and 22.97 Ma, respectively (Figure 7). The LO surfaces of calcareous nannofossils Sphenolithus belemnos, Triquetrorhabdulus carinatus, and Disctyococcites bisectus have been found in the Zhujiang Formation in Well Y3 at depths of 2,826, 2,928, and 3,124, respectively. According to the ATS, the three biological events are dated at 17.18, 19.08, and 22.97 Ma, respectively (Figure 7). A comparison of the ATS ages of these seven biological events with those of their predecessors reveals that there is a difference of 0.06–0.72 Ma [18,3840].

5.3 Division of high-frequency cycles

The high-frequency cycles are controlled by climate and sea-level changes caused by the Milankovitch cycles. The fourth-, fifth-, and sixth-order cycles were controlled by long eccentricity (405 ka), short eccentricity (∼100 ka), and obliquity cycles (∼40 ka), respectively. Selecting different Milankovitch cycles as the standard for the division of high-frequency cycles can realize isochronicity and improve the accuracy of the stratigraphic division and correlation.

Based on the above spectral analysis, the center frequency and bandwidth of the 405 ka long eccentricity cycle in Well Y3 are 0.0455 and 0.01, respectively; those of the 95 ka short eccentricity cycle are 0.1996 and 0.04, respectively; and those of the 40.4 ka obliquity cycles are 0.4549 and 0.05, respectively. Milankovitch curves of different frequency bands were obtained using Gaussian bandpass filtering of the preprocessed GR data. These Milankovitch curves were used as reference curves for the division of the fourth-, fifth-, and sixth-order cycles. Combined with the core and logging data, we divided the Zhujiang Formation of Well Y3 into 17 fourth-, 76 fifth-, and 174 sixth-order cycles (Figure 8).

Figure 8 
                  High-frequency cycles identification and division of the Zhujiang Formation in Well Y3 (405, 95, and 40.4 ka filtered curves passband: 0.0455 ± 0.01, 0.1996 ± 0.04, and 0.4549 ± 0.05).
Figure 8

High-frequency cycles identification and division of the Zhujiang Formation in Well Y3 (405, 95, and 40.4 ka filtered curves passband: 0.0455 ± 0.01, 0.1996 ± 0.04, and 0.4549 ± 0.05).

6 Conclusions

In this study, by conducting spectral and wavelet analyses to conduct cyclostratigraphic research, a floating ATS was constructed based on 405 ka orbital eccentricity cycle tuning of the GR series in the Zhujiang Formation. Based on the ATS, sequence stratigraphy and biological events were accurately dated, and a high-precision chronostratigraphic framework of the Zhujiang Formation in the Baiyun Sag was established. The following conclusions were drawn:

  1. The Zhujiang Formation in the Baiyun Sag records Milankovitch cycles. The stratigraphic cycles controlled by the 405 and 95 ka eccentricity cycles, 40.4 ka obliquity cycles, and 23.5 ka precession cycles are identified, and the signal of the stratigraphic cycles controlled by the 405 ka long eccentricity cycles is the strongest.

  2. Based on the ATS, the precise durations of the Zhujiang Formation in Wells Y1, Y2, and Y3 are 7.13, 6.93, and 7.18 Ma, respectively. The average sedimentation rates of the Zhujiang Formation in Wells Y1, Y2, and Y3 are 4.68, 5.91, and 5.33 cm/ka, respectively.

  3. The ATS ages of the top of sequences SQ1–SQ5 are 22.6, 22.1, 18.56, 16.89, and 15.85 Ma, respectively. The FO ATS ages of the planktonic foraminifera G. sicanus and those of the LO of the planktonic foraminifera C. dissimilis, G. binaiensis, and G. ciperoensis in Well Y3 are 16.53, 17.39, 18.56, and 22.97 Ma, respectively. The LO ATS ages of calcareous nannofossils S. belemnos, T. carinatus, and D. bisectus in Well Y3 are 17.18, 19.08, and 22.97 Ma, respectively.

  4. Based on the Milankovitch theory, the Zhujiang Formation in the Baiyun Sag can be divided into 17 fourth-, 76 fifth-, and 174 sixth-order cycles.

Acknowledgments

The authors are grateful to the anonymous reviewers for their valuable comments, and the editors for their careful revisions. All these helped to improve the manuscript significantly.

  1. Funding information: This research was supported by the National Natural Science Foundation of China (Grant No. 41472098) and the National 13th Five-Year Plan’s Major Science and Technology Program of China (Grant No. 2017ZX05032-002-002).

  2. Author contributions: Shangfeng Zhang contributed to the conception of the study and approved the final version of the article for publication, Ping He analyzed the data and wrote the manuscript, Enze Xu helped perform the analysis with constructive discussions, and Chunxia Zhu contributed to the data collection.

  3. Conflict of interest: The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this article.

References

[1] Berger A, Loutre MF, Laskar J. Stability of the astronomical frequencies over the earth’s history for paleoclimate studies. Sci N Ser. 1992;255(5044):560–6.10.1126/science.255.5044.560Search in Google Scholar PubMed

[2] Strasser A, Hilgen FJ, Heckel PH. Cyclostratigraphy – concepts, definitions, and applications. Newsl Stratigr. 2006;42(2):75–114.10.1127/0078-0421/2006/0042-0075Search in Google Scholar

[3] Wu HC, Zhang SH, Feng QL, Fang NQ, Yang TS, Li HY. Theoretical basis, research advancement and prospects of cyclostratigraphy. Earth Sci. 2011;36(3):409–28.Search in Google Scholar

[4] Yan JP, Yan Y, Peng J, Feng CZ, Geng B, Li XW. Log identification of astronomical cycle in lacustrine facies mud shale and its application in the Lower 3rd Member of Shahejie Formation in Zhanhua Sag. Well Log. 2017;41(6):701–7.Search in Google Scholar

[5] Wu HC, Zhong YY, Fang Q, Yang TS, Li HY, Zhang SH. Paleozoic cyclostratigraphy and astronomical time scale. Bull Mineral Petrol Geochem. 2017;36(5):750–70.Search in Google Scholar

[6] Fang Q, Wu HC, Hinnov LA, Jing XC, Wang XL, Yang TS, et al. Astronomical cycles of Middle Permian Maokou Formation in South China and their implications for sequence stratigraphy and paleoclimate. Palaeogeogr Palaeoclimatol Palaeoecol. 2017;474:130–9.10.1016/j.palaeo.2016.07.037Search in Google Scholar

[7] Svensen HH, Hammer Corfu F. Astronomically forced cyclicity in the Upper Ordovician and U-Pb ages of interlayered tephra, Oslo Region, Norway. Palaeogeogr Palaeoclimatol Palaeoecol. 2015;418:150–9.10.1016/j.palaeo.2014.11.001Search in Google Scholar

[8] Liu GH, Zhang SH, Wu HC. Progress and challenges in Precambrian cyclostratigraphy research. J Stratigraphy. 2020;44(3):239–49.Search in Google Scholar

[9] Abels HA, Abdul Aziz H, Krijgsman W, Smeets Sander JB, Hilgen FJ. Long-period eccentricity control on sedimentary sequences in the continental Madrid Basin (middle Miocene, Spain). Earth Planet Sci Lett. 2010;289(1–2):220–31.10.1016/j.epsl.2009.11.011Search in Google Scholar

[10] Wu HC, Zhang SH, Feng QL, Jiang GQ, Li HY, Yang TS. Milankovitch and sub-Milankovitch cycles of the Early Triassic Daye Formation, South China and their geochronological and paleoclimatic implications. Gondwana Res. 2012;22(2):748–59.10.1016/j.gr.2011.12.003Search in Google Scholar

[11] Huang CJ. The current status of cyclostratigraphy and astrochronology in the Mesozoic. Earth Sci Front. 2014;21(2):48–66.Search in Google Scholar

[12] Liu ZH, Huang CJ, Algeo TJ, Liu HM, Hao YQ, Du XB, et al. High-resolution astrochronological record for the Paleocene–Oligocene (66–23 Ma) from the rapidly subsiding Bohai Bay Basin, northeastern China. Palaeogeogr Palaeoclimatol Palaeoecol. 2018;510:78–92.10.1016/j.palaeo.2017.10.030Search in Google Scholar

[13] Wu HC, Zhang SH, Hinnov LA, Jiang GQ, Yang TS, Li HY, et al. Cyclostratigraphy and orbital tuning of the terrestrial upper Santonian-Cambrian Danian in Songliao Basin, northeastern China. Earth Planet Sci Lett. 2014;407:82–95.10.1016/j.epsl.2014.09.038Search in Google Scholar

[14] Hinnov LA. Cyclostratigraphy and its revolutionizing applications in the earth and planetary sciences. Geol Soc Am Bull. 2013;125(11–12):1703–34.10.1130/B30934.1Search in Google Scholar

[15] Xu K, Chen HH, Huang CJ, Ogg JG, Zhu JX, Lin SQ, et al. Astronomical time scale of the Paleogene lacustrine paleoclimate record from the Nanxiang Basin, central China. Palaeogeogr Palaeoclimatol Palaeoecol. 2019;532:1–19.10.1016/j.palaeo.2019.109253Search in Google Scholar

[16] Wang M, Chen HH, Huang CJ, Kemp DB, Xu TW, Zhang HA, et al. Astronomical forcing and sedimentary noise modeling of lake-level changes in the Paleogene Dongpu Depression of North China. Earth Planet Sci Lett. 2020;535:1–13.10.1016/j.epsl.2020.116116Search in Google Scholar

[17] Qin GQ. Late Cenozoic sequence stratigraphy and sea-level changes in Pearl River Mouth Basin, South China Sea. China Offshore Oil Gas. 2002;16(1):1–10.Search in Google Scholar

[18] Li QY, Zheng F, Liu CL. Stratigraphic events across the Oligocene/Miocene boundary. Mar Geol Quat Geol. 2007;27(5):57–64.Search in Google Scholar

[19] Li SS, Xie JY, Liu XY, He WJ, Zhu YH. Biostratigraphic research on the Paleogene–Neogene boundary in the western Pearl River Mouth Basin. Acta Micropalaeontol Sin. 2010;27(3):221–30.Search in Google Scholar

[20] Zhu YH, Chen F. Review of the Neogene calcareous nannofossil research of the Pearl River Mouth Basin, South China. Acta Micropalaeontol Sin. 2007;24(1):76–81.Search in Google Scholar

[21] Zhao F, Zhu YH, Ma RF. Early Miocene to Pleistocene calcareous nannofossil stratigraphy and palaeoenvironment of Well PY30 in the Pearl River Mouth Basin, Northern South China Sea, China. Acta Micropalaeontol Sin. 2017;34(3):332–42.Search in Google Scholar

[22] Lin CS, Shi HS, Li H, He M, Zhang ZT, Gong Y, et al. Sequence architecture, depositional evolution and controlling processes of continental slope in Pearl River Mouth Basin, Northern South China Sea. Earth Sci. 2018;43(10):3407–22.Search in Google Scholar

[23] Liu Y, Wu HC, Zhang SH, Fang NQ, Jiang ZL, Yang TS. Cyclostratigraphy research on the Hanjiang–Wanshan Formations in Zhuyi Depression, Pearl River Mouth Basin. Earth Sci. 2012;37(3):411–23.Search in Google Scholar

[24] Tian SF, Chen ZQ, Zha M. Astronomical time scale of middle Miocene Hanjiang formation in Pearl River Mouth Basin, South China Sea. J China Univ Pet. 2012;36(1):27–32.Search in Google Scholar

[25] Tian SF, Chen ZQ, Gao CH, Zha M. Astronomical tuning of the early Middle Miocene Hanjiang Formation in the Pearl River Mouth Basin, South China Sea. Acta Geologica Sin (Engl Ed). 2013;87(1):48–58.10.1111/1755-6724.12029Search in Google Scholar

[26] Shi HS, He M, Zhang LL, Yu QH, Pang X, Zhong ZH, et al. Hydrocarbon geology, accumulation pattern and the next exploration strategy in the eastern Pearl River Mouth Basin. China Offshore Oil Gas. 2014;26(3):11–22.Search in Google Scholar

[27] Hou MC, Deng M, Shi HS, Liu J, Wang RL, Zeng Y, et al. The process of growth, development and extinction of the Early Miocene carbonate rocks and controlling factors in the Pearl River Mouth Basin. Acta Petrologica Sin. 2017;33(4):1257–71.Search in Google Scholar

[28] Radwan AE. Modeling the depositional environment of the sandstone reservoir in the Middle Miocene Sidri Member, Badri Field, Gulf of Suez Basin, Egypt: integration of gamma-ray log patterns and petrographic characteristics of lithology. Nat Resour Res. 2021;30(1):431–49.10.1007/s11053-020-09757-6Search in Google Scholar

[29] Yuan R, Zhu R, Xie SW, Hu W, Zhou FJ, Yu Y. Utilizing maximum entropy spectral analysis (MESA) to identify Milankovitch cycles in Lower Member of Miocene Zhujiang Formation in north slope of Baiyun Sag, Pearl River Mouth Basin, South China Sea. Open Geosci. 2019;11(1):877–87.10.1515/geo-2019-0068Search in Google Scholar

[30] Liu ZC, Chen Y, Yuan LW. The paleoclimate changes of Qaidam Basin during the last 2.85 Ma recorded by Gamma-ray logging. Sci China Ser D. 2001;44(2):133–45.10.1007/BF02879656Search in Google Scholar

[31] Laskar J, Fienga A, Gastineau M, Manche H. La2010: a new orbital solution for the long-term motion of the Earth. Astron Astrophys. 2011;532:A89. 10.1051/0004-6361/201116836.Search in Google Scholar

[32] Hilgen FJ, Lourens LJ, Van Dam JA. The Neogene period. In: Gradstein FM, Ogg JG, Schmitz MD, Ogg GM, editors. The Geologic Time Scale 2012. Oxford, UK: Elsevier; 2012. p. 923–78.10.1016/B978-0-444-59425-9.00029-9Search in Google Scholar

[33] Hinnov LA. New perspectives on orbitally forced stratigraphy. Annu Rev Earth Planet Sci. 2000;28:419–75.10.1146/annurev.earth.28.1.419Search in Google Scholar

[34] Weedon GP. Time-series analysis and cyclostratigraphy: examining stratigraphic records of environmental cycles. Cambridge: Cambridge University Press; 2003. p. 217–20.10.1017/CBO9780511535482Search in Google Scholar

[35] Li QY, Tian J, Wang PX. Recognizing the stratigraphic and paleoclimatic significance of eccentricity cycles. Earth Sci. 2005;30(5):519–28.Search in Google Scholar

[36] Huang CJ, Hinnov L. Astronomically forced climate evolution in a saline lake record of the middle Eocene to Oligocene, Jianghan Basin, China. Earth Planet Sci Lett. 2019;528:1–13.10.1016/j.epsl.2019.115846Search in Google Scholar

[37] Gong YM, Du YS, Tong JN, Zhang KX, Feng QL, Xie SC, et al. Cyclostratigraphy: the third milestone of stratigraphy in understanding time. Earth Sci. 2008;33(4):443–57.Search in Google Scholar

[38] Berggren WA, Kent DV, Swisher CC, Aubry MP. A revised Cenozoic geochronology and chronostratigraphy. In: Berggren WA, Kent DV, Aubry MP, Hardenbol J, editors. Geochronology, time scales and global stratigraphic correlation. Oklahoma, USA: SEPM Society for Sedimentary Geology. Vol. 54; 1995. p. 129–212.10.2110/pec.95.04.0129Search in Google Scholar

[39] Pearson PN, Chaisson WP. Late Paleocene to Middle Miocene planktonic foraminifer biostratigraphy of the Ceara Rise. In: Shackleton NJ, Curry WB, Richter C, Bralower TJ, editors. Proceedings of the Ocean Drilling Program, Scientific Results. Vol. 154; 1997. p. 33–68.10.2973/odp.proc.sr.154.106.1997Search in Google Scholar

[40] Mao XL, Xu SL, Liu XY. Late Cenozoic high resolution biostratigraphy and its bearing on sea-level fluctuation in the western Pearl River Mouth Basin. Mar Geol Quaternary Geol. 2019;39(3):40–50.Search in Google Scholar

Received: 2021-06-28
Revised: 2022-10-24
Accepted: 2022-10-26
Published Online: 2022-12-02

© 2022 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 10.12.2023 from https://www.degruyter.com/document/doi/10.1515/geo-2022-0434/html
Scroll to top button