• Journal of Geographical Sciences
  • Vol. 30, Issue 12, 2002 (2020)
Tuoyu LI1, Jifeng ZHANG2、*, Yongqiu WU3, Shisong DU3, Duowen MO4, Yinan LIAO4, Zhitong CHEN2, Jianbao LIU2, and Qing LI5
Author Affiliations
  • 1Capital Normal University, Beijing 100048, China
  • 2Key Laboratory of Alpine Ecology (LAE), Institute of Tibetan Plateau Research, CAS, Beijing 100101, China
  • 3MOE Engineering Center of Desertification and Blown-sand Control, Beijing Normal University, Beijing 100875, China
  • 4Laboratory for Earth Surface Processes, Ministry of Education, College of Urban and Environmental Sciences, Peking University, Beijing 100871, China
  • 5Institute of Geographical Sciences, Hebei Engineering Research Center of Geographic Information Application, Hebei Academy of Sciences, Shijiazhuang 050011, China
  • show less
    DOI: 10.1007/s11442-020-1824-6 Cite this Article
    Tuoyu LI, Jifeng ZHANG, Yongqiu WU, Shisong DU, Duowen MO, Yinan LIAO, Zhitong CHEN, Jianbao LIU, Qing LI. Holocene aeolian activities linked to Indian summer monsoon in the middle reaches of the Yarlung Zangbo River[J]. Journal of Geographical Sciences, 2020, 30(12): 2002 Copy Citation Text show less

    Abstract

    Widespread aeolian deposits on the Tibetan Plateau (TP) have provided valuable palaeoclimatic information. However, the primary factors (e.g., climate factors, human activity, and vegetation cover) controlling aeolian deposition remain elusive. In this paper, we use a dataset that comprises new and published ages of Holocene aeolian sand and loess in the middle reaches of the Yarlung Zangbo River to identify the primary controlling factors and palaeoclimatic implications of aeolian deposition. Several intervals of enhanced aeolian accumulation centered at 8.5-7.8, 6.4-5.8, 4.5-4.0, 3.1-1.8, and 0.9 ka are identified, generally consistent with regional low rainfall events and weak Indian summer monsoon (ISM). This suggests that regional wetness, dominated by the ISM, may play a key role in modulating dust emissions and aeolian deposition on centennial timescales. Our results show that on centennial- to millennial-scales, ISM activity can be reconstructed by non-continuous aeolian deposits in the monsoon dominated TP.

    1 Introduction

    The Tibetan Plateau (TP), with a mean elevation greater than 4000 m above sea level, plays a key role in modulating Asian and global climate systems (e.g., An et al., 2001; Molnar et al., 2010). Aeolian sediments are very widespread on the TP and have been used to reconstruct local and regional climatic and environmental changes (Lehmkuhl, 2000, 2014; Sun et al., 2007; Kaiser et al., 2009; Lai et al., 2009; Stauch et al., 2012, 2018; Stauch, 2015, 2016; Zhang et al., 2015; Feng et al., 2016; Dong et al., 2017). The TP contains aeolian sediments, which are mainly distributed on dry basins, wide river valleys, lakeshores, and other sites (Dong et al., 2017). These sediments always originate from local materials, including glacial outwash material (Sun et al., 2007; Zhang et al., 2015), fluvial/alluvial deposits, and exposed lake sediments from shrinking lakes (Lehmkuhl et al., 2000).

    Although aeolian sand and loess deposits serve as an important archive for observing topographic and climatic events on the TP, the specific palaeoclimatic interpretations they reflect remain elusive. Generally, the formation of aeolian sediments requires three prerequisites: (1) availability of sediments for dust formation, (2) suitable wind for dust transportation, and (3) a suitable site for dust accumulation (Pye, 1995). Usually, the accumulation of aeolian sediments on the TP is interpreted as reflecting dry environments, while the formation of palaeosols is used as an indicator of wet climatic conditions (e.g., Lu et al., 2011; Chen et al., 2016). However, a number of recent studies have shown that the strongest aeolian accumulations occur during the relatively wet late Glacial and early Holocene, suggesting that suitable trapping conditions with sufficient vegetation cover may play a key role in the formation of aeolian sediments (Sun et al., 2007; Stauch, 2015; Qiang et al., 2016).

    Local dust storms, sand dunes, and well-preserved aeolian sand and loess deposits are widespread in the southern TP (Li et al., 1999; Sun et al., 2007; Lai et al., 2009; Shen et al., 2012; Pan et al., 2014; Zhang et al., 2015; Li et al., 2016; Dong et al., 2017; Ling et al., 2019, 2020). Sun et al., (2007) dated loess deposits from several sites in the Yarlung Zangbo River (YZR) valley, southern TP, and found that aeolian deposits accumulated only after the last deglaciation and not during glacial periods. The authors suggested that the lack of glacial loess deposits was due to the minimal vegetation cover or the erosion of loess during deglaciation.

    On the other hand, Lai et al. (2009) reported several ages of aeolian deposits from the Kyichu River valley, southern TP, throughout most of the last 100 ka, which is much older than what was previously thought. The authors worked with a dataset that comprises 24 ages of aeolian deposits and found episodic aeolian deposition at approximately 3, 8, 16-21, 33, and 79-83 ka, most of which can be explained through synchronous global arid events. Overall, a number of studies have reported late Quaternary aeolian deposits in the southern TP; however, only sporadic Holocene aeolian sediments have been reported, and their possible linkages to large-scale atmospheric circulations, such as the Indian summer monsoon (ISM) and the westerlies, remain unknown.

    Previous studies have shown that the TP is predominantly controlled by the interplay between the ISM and the westerlies (Bohner, 2006; An et al., 2012; Zhu et al., 2015). Modern precipitation δ18O observations have shown that the ISM moisture can reach 34°-35°N in the southern part of TP (e.g., Yao et al., 2013), while the northern part of the TP is primarily dominated by westerlies. In the middle reaches of the YZR, annual precipitation mainly occurs in summer under the influence of the ISM, while the effects of mid-latitude westerlies on regional precipitation are limited. Several palaeo-moisture records based on lake sediments in the region showed several regional wet and dry intervals associated with ISM variability during the Holocene (Zhu et al., 2002; Bird et al., 2014; Conroy et al., 2017). In this paper, we use a dataset that comprises six new and 46 prior published ages of Holocene aeolian sand and loess in the middle reaches of the YZR to identify aeolian dynamics on centennial to millennial timescales, and their association with large-scale atmospheric circulation (e.g., the ISM).

    2 Material and methods

    The YZR, with a length of 2,057 km, is the longest river in the southern TP; it lies between the Gangdise and Nyainqentanglha mountains and the Himalayas orogeny (Figure 1). In summer, the ISM intrudes through the YZR valley, delivering most of this region’s annual precipitation, while in winter, the climate is cold and dry under the impact of mid-latitude westerlies (Zhang et al., 2017). Some reaches of the YZR are characterized by broad valleys that provided suitable sites for the formation of aeolian deposits, which were as large as 2736 km² in 2008 (Shen et al., 2012). Optically stimulated luminescence (OSL), thermoluminescence (TL), and radiocarbon (14C) methods have been used to date aeolian sand and loess in the YZR basin and its surrounding areas (Sun et al., 2007; Li et al., 2010; Zhang et al., 2015; Li et al., 2016). Due to rapid bleaching by sunlight, luminescence methods are the most appropriate technique to date aeolian sediments (Singhvi et al., 2001; Wintle and Murray, 2006). Occasionally, the radiocarbon ages of plant residues, charcoal, and total organic matter in aeolian deposits are also used to measure the ages of dust depositions.

    Location of the study area (a) and distribution of aeolian sediment profiles in the middle reaches of the YZR (b). The boundary of westerlies and the ISM (dashed line; Chen et al., 2010) are also shown.

    Figure 1.Location of the study area (a) and distribution of aeolian sediment profiles in the middle reaches of the YZR (b). The boundary of westerlies and the ISM (dashed line; <xref ref-type="bibr" rid="b10">Chen <italic>et al</italic>., 2010</xref>) are also shown.

    In this study, we sampled five aeolian profiles, named DRX, RM, WL, SK, and DPZ (Figures 1 and 2), in the YZR valley and obtained four OSL ages (Table 1) and two radiocarbon ages (Table 2). These aeolian profiles are located on the fluvial terraces or the foothills of mountains along the middle reaches of the YZR. Massive suspended sediments deposited in the braided river channels or alluvial fans are reworked by local near-surface winds in dry seasons, and they accumulate on the foothills of mountains along the river to form dunes and loess (Sun et al., 2007; Zhang et al., 2015). The DRX profile (29°21′59.2″N, 91°08′58.0″E) is located on a fluvial terrace and can be divided into four sedimentary units: 0-50 cm, fluvial sand; 50-60 cm, gravel unit; 60-90 cm, aeolian sand unit (an OSL age sample collected at 68-72 cm); 90-120 cm, gravel unit. The RM profile (29°21′10.0″N, 88°27′41.5″E) is located on a foothill and can be divided into three sedimentary units: 0-65 cm, aeolian loess unit with plant roots (an OSL age sample collected at 28-32 cm); 65-130 cm, palaeosol unit; 130-290 cm, aeolian loess unit (an OSL age sample collected at 208-212 cm). The WL (29°8′40.3″N, 93°40′35.9″), SK (29°17′58.1″N, 91°24′41.3″), and DPZ (29°17′0.4″N, 91°38′54.7″E) profiles are also located on foothills and consist mainly of aeolian sand with a thickness of approximately 440 cm. A charcoal sample from the aeolian sand of SK and an organic matter sample from the aeolian sand of WL were used for 14C dating of SK and WL, respectively (Table 2). An OSL sample was also collected from the aeolian sand of DPZ (Table 1).

    Sediment logs and ages of aeolian sand and loess in the middle reaches of the YZR

    Figure 2.Sediment logs and ages of aeolian sand and loess in the middle reaches of the YZR

    SampleDepth (cm)U (ppm)Th (ppm)K (%)Dose rate (Gy•ka‒1)De/GyOSL age (ka)
    DRX-OSL68-721.56±0.313.95±0.72.41±0.043.73±0.272.93±0.300.79±0.10
    RM-OSL-128-322.64±0.03917.8±0.0252.46±0.0264.36±0.327.99±0.401.83±0.29
    RM-OSL-3208-2123.09±0.03819.2±0.0242.06±0.0304.17±0.3119.35±0.984.64±0.36
    DPZ-OSL-3436-4404.16±0.03418.2±0.0252.36±0.0284.76±0.358.37±0.421.76±0.18

    Table 1.

    OSL dating results of aeolian sediments in the YZR basin

    SampleDepth/cmDating materialConventional 14C age (BP, 2σ)Calibrated14C age (BP, 2σ)
    SK-14C-1435-439Charcoal2880±253005±37
    WL-14C-150-55Total organic matter710±20670±6

    Table 2.

    Radiocarbon dating results of aeolian sediments in the YZR basin

    Pure quartz was extracted in a dark room with a dimmed red light for OSL dating. Optically stimulated luminescence measurements were carried out on an automated Risø TL/OSL-DA-20-C/D reader, and De values were calculated with the single aliquot regenerative-dose (SAR) protocol (Murray and Wintle 2000). Uranium (238U), thorium (232Th), and potassium (40K) concentrations were measured using inductively coupled plasma mass spectrometry (ICP-MS). The resulting dose rate for each sample was estimated based on depth, altitude, and geomagnetic latitude. Cosmic ray contributions were calculated based on Prescoott and Hutton (1994). The water content of each sample was set to 15%±7%. The measurement accuracy of the equivalent dose using the SAR protocol was approximately 2%.The OSL samples were measured at the Nanjing Normal University and the Qinghai Normal University. The materials used for the two radiocarbon dating samples were charcoal and total organic matter, and the radiocarbon dating was processed at Peking University. IntCal13 was used to calibrate all radiocarbon data were calibrated for the calendar year (Reimer et al., 2013).

    To show the regional patterns of aeolian activity, we also considered prior published ages of aeolian profiles in the middle reaches of the YZR. Ultimately, we obtained a dataset that comprises ages of aeolian sand and loess in the middle reaches of the YZR, including 52 ages from 28 profiles (Table 3). These profiles are mainly distributed in the broad valley of the YZR, with a small amount distributed in the YZR tributary valley (Figure 1). Cumulative probability density functions have been widely used for OSL age (e.g., Singhvi et al., 2001; Lai et al., 2009; Stauch, 2015) and radiocarbon age (e.g., Hoffmann et al., 2008) distribution analysis. Each OSL and radiocarbon age was assumed to be a dust deposition event, and respective probability density functions were calculated using the mean value and standard deviation of each age. Cumulative probability density functions were calculated at 50-year intervals: for each time point, with an interval of 50 years, we summed up the probability values of all the ages to obtain a cumulative probability distribution record. Peaks of the summed probability density distributions were thought to represent intervals of heavy dust deposition.

    SectionDepth (cm)Dating methodDating materialAge (ka/cal ka BP)Latitude (°N)Longitude (°E)Altitude (m asl)Source
    TB1350OSLAeolian loess2.70±0.2029.316789.55003800Sun et al., 2007
    TB7380OSLAeolian loess11.00±1.2029.316788.91673920Sun et al., 2007
    DAR1300-325 14CCharcoal3.15±0.08Kaiser et al., 2009
    STA150OSLAeolian sand2.90±0.2029.633191.09783660Kaiser et al., 2009
    STA1180OSLAeolian sand4.10±0.4029.633191.09783667Kaiser et al., 2009
    STA1280OSLAeolian sand6.70±0.5029.633191.09783667Kaiser et al., 2009
    QUX 1280OSLAeolian sand8.50±0.7029.355390.72343603Kaiser et al., 2009
    QUX 2325-330 14CCharcoal7.78±0.0729.365990.75563536Kaiser et al., 2009
    Section 4867-73OSLAeolian loess8.80±3.9029.733389.81674571Lehmkuhl et al., 2000
    Section 4947-53OSLAeolian loess7.80±1.2029.766789.85004835Lehmkuhl et al., 2000
    LXD170OSLAeolian loess7.90±0.9029.327589.53863797Hu et al., 2018
    LXD98OSLAeolian loess3.20±0.3029.327589.53863797Hu et al., 2018
    Xigaze 14COrganic matter0.92±0.0229.305788.86883811Hu et al., 2018
    TDD87OSLAeolian loess2.60±0.3029.337290.32363687Hu et al., 2018
    TDD195OSLAeolian loess2.90±0.3029.337290.32363687Hu et al., 2018
    TDD285OSLAeolian loess5.00±0.5029.337290.32363687Hu et al., 2018
    JB260TLAeolian sand8.56±0.6529.396989.35003890Li et al., 2010
    QS430TLAeolian loess8.85±0.5329.390090.75784000Li et al., 2010
    GM34014COrganic matter6.20±0.31Li et al., 2010
    GM531TLAeolian sand8.30±0.30Li et al., 2010
    Cha'er6514COrganic mattera2.23±0.1029.389589.28233856Zheng et al., 2009
    Cha'er235TLAeolian sand8.56±0.6529.389589.28233856Zheng et al., 2009
    ZD158OSLAeolian loess5.90±0.2029.246691.71203561Zheng, 2009
    ZD628OSLAeolian sand8.50±0.6029.246691.71203561Zheng, 2009
    CGG168OSLAeolian sand1.82±0.1629.365391.14913652Li et al., 2020
    CGG287OSLAeolian sand8.43±0.6629.365391.14913652Li et al., 2020
    YJP10.4OSLSandy loess1.90±0.1029.455694.46932943Ling et al., 2020
    YJP10.9OSLSandy loess3.90±0.3029.455694.46932943Ling et al., 2020
    YJP11.4OSLSandy loess4.40±0.3029.455694.46932943Ling et al., 2020
    YJP11.9OSLSandy loess4.30±0.3029.455694.46932943Ling et al., 2020
    YJP12.5OSLSandy loess5.10±0.4029.455694.46932943Ling et al., 2020
    YJP13OSLSandy loess3.20±0.2029.455694.46932943Ling et al., 2020
    YJP13.6OSLSandy loess8.30±0.6029.455694.46932943Ling et al., 2020
    YJP21.7OSLSandy loess110±0.9029.455694.46932943Ling et al., 2020
    MLP6.5OSLAeolian sand4.50±0.3029.118993.77813004Ling et al., 2020
    MLP10OSLAeolian sand6.20±0.5029.118993.77813004Ling et al., 2020
    LXP1.3OSLSandy loess4.90±0.4029.066892.79933172Ling et al., 2020
    LXP2OSLSandy loess6.50±0.5029.066892.79933172Ling et al., 2020
    SRP0.7OSLAeolian sand0.40±0.1029.261791.98733553Ling et al., 2020
    SectionDepth (cm)Dating methodDating materialAge (ka/cal ka BP)Latitude (°N)Longitude (°E)Altitude (m asl)Source
    SRP1.4OSLAeolian sand0.80±0.1029.261791.98733553Ling et al., 2020
    SRP2.1OSLAeolian sand1.00±0.1029.261791.98733553Ling et al., 2020
    SRP2.8OSLAeolian sand1.10±0.1029.261791.98733553Ling et al., 2020
    SRP3.5OSLAeolian sand1.00±0.1029.261791.98733553Ling et al., 2020
    SRP4.2OSLAeolian sand1.20±0.1029.261791.98733553Ling et al., 2020
    SRP4.9OSLAeolian sand4.10±0.4029.261791.98733553Ling et al., 2020
    LCP2.9OSLSandy loess9.20±0.8029.387289.32543815Ling et al., 2020
    DRX70OSLAeolian sand0.79±0.1029.366491.14943656This study
    RM30OSLAeolian loess1.83±0.2929.352888.46153876This study
    RM210OSLAeolian loess4.64±0.3629.352888.46153876This study
    WL52 14COrganic matter0.67±0.0129.144593.67663092This study
    SK437 14CCharcoal3.01±0.0429.299591.41153557This study
    DPZ438OSLAeolian sand1.76±0.1829.283591.64853584This study

    Table 3.

    A dataset of synthesized ages of Holocene aeolian sediments in the middle reaches of the YZR

    3 Results and discussion

    The cumulative probability density curve of the ages of Holocene aeolian sediments in the middle reaches of the YZR, as shown in Figure 3a, shows the strongest deposition during the late Holocene and weak aeolian activity during the early Holocene. This contrasts with aeolian activity reports on the northeastern TP, in which several recent studies have reported increased dust accumulation during relatively wet late Glacial and early Holocene periods, highlighting the key role of vegetation cover as a dust trap in aeolian sediment accumulation (Stauch, 2015). Our synthesized curve is also punctuated by several strong dust deposition events (Figure 3a) centered at 8.5-7.8, 6.4-5.8, 4.5-4.0, 3.1-1.8, and 0.9 ka. Two of these events are similar to those reported in Lai et al. (2009) that analyzed a YZR tributary.

    Comparison of the probability density curve for Holocene aeolian sediment ages in the middle reaches of the YZR (a) with other palaeoclimatic records: (b) summer insolation at 30°N (Berger and Loutre, 1991); (c) a lake-level record reconstructed by the PC1 grain size of the Lake Paru Co (Bird et al., 2014); (d) isotopic divergence between C23 and C31 n-alkanes (ΔδD) in Hongyuan peat (Seki et al., 2011); (e) a stalagmite δ18O record from southern Oman (Fleitmann et al., 2003).

    Figure 3.Comparison of the probability density curve for Holocene aeolian sediment ages in the middle reaches of the YZR (a) with other palaeoclimatic records: (b) summer insolation at 30°N (<xref ref-type="bibr" rid="b3">Berger and Loutre, 1991</xref>); (c) a lake-level record reconstructed by the PC1 grain size of the Lake Paru Co (<xref ref-type="bibr" rid="b4">Bird <italic>et al</italic>., 2014</xref>); (d) isotopic divergence between C<sub>23</sub> and C<sub>31</sub> n-alkanes (ΔδD) in Hongyuan peat (<xref ref-type="bibr" rid="b47">Seki <italic>et al</italic>., 2011</xref>); (e) a stalagmite δ<sup>18</sup>O record from southern Oman (<xref ref-type="bibr" rid="b15">Fleitmann <italic>et al</italic>., 2003</xref>).

    Several events of increased dust deposition in the middle reaches of the YZR, such as the event around 8 ka, have not been reported on the northeastern TP (e.g., Stauch, 2015), suggesting different mechanisms for aeolian sedimentation in the southern and northeastern TP. First, this difference can be related to the different atmospheric circulation backgrounds of the two regions. The northeastern TP is mainly influenced by the East Asian summer monsoon, while the southern TP is mainly impacted by the ISM. Second, this difference can be explained by the different mechanisms for the supply of dust to the two regions. The much larger area of desertic land on the northeastern TP may provide sufficient supplies of dust during the early Holocene. Thus, improved vegetation cover, as a sediment trap, may contribute to an enhanced aeolian deposition during the early Holocene. In contrast, aeolian sediments in the YZR valley are supplied by local dust, which is mainly generated from small areas of outwash sediment and desertic land in the low-level floodplains (Sun et al., 2007; Zhang et al., 2015). Strong ISM activity in response to high summer insolation (Figure 3b) and increased regional monsoon rainfall during the early Holocene (Fleitmann et al., 2003; Gupta et al., 2003) may improve vegetation cover in the source area and significantly inhibit dust supply and aeolian activity in the middle reaches of the YZR. Finally, the confined valley landscape of the YZR, which differs significantly from the open terrain in northern TP, may partly explain the differences in aeolian activity between the two regions.

    To gain a better understanding of the controlling factors of aeolian sedimentation in the middle reaches of the YZR, we compared our records with regional palaeoclimatic records. Reconstructed aeolian activities identified in this study are generally consistent with regional palaeohydrological records on the TP. Monsoon rainfall record collected from Paru Co, which is located in the southern TP, shows that several intervals of increased dust deposition mostly happen in the dry seasons (Bird et al., 2014) (Figure 3c). In this study, the strong aeolian sediment accumulation events approximately 8.5-7.8, 6.4-5.8, and 3.1-1.8 ka are consistent with several dry season intervals identified by a rainfall record based on leaf wax hydrogen isotopic evidence from Hongyuan peat (Seki et al., 2011) (Figure 3d). A pollen record from Lake Tangra Yumco, in central TP, has identified several weak ISM events at 8.2-7.4, 5.5, 4.7-3.7, and 3.2 ka (Ma et al., 2019), which are largely consistent with several dust deposition events identified in this study, given the dating uncertainty of the different archives. Moreover, the drought event around 8.5-7.8 ka is also found in other lacustrine records in the western (Gasse et al., 1996; Hou et al., 2017), central (Zhu et al., 2008; Doberschutz et al., 2014; Ma et al., 2014), and eastern (Hong et al., 2003; Kramer et al., 2010) TP. Our research has made it possible to record a long duration of increased dust deposition at 3.1-1.8 ka, which is consistent with the expected significantly reduced monsoon rainfall events across the TP (Zhu et al., 2008; Liu et al., 2009; Xie et al., 2009; Kramer et al., 2010; Zhao et al., 2011; Hou et al., 2017; Shi et al., 2017).

    Modern precipitation isotope observations have shown that moistures transported by the ISM can invade the TP through the Brahmaputra/YZR valley and reach as far as the Tanggula Mountains (Tian et al., 2001, 2007). To identify the main climatic controlling factors for dust deposition in the southern TP, we compared our results with several records of ISM activity in Asia that have been reconstructed by marine sediment and stalagmite. Our reconstructed aeolian activity intensity in the YZR basin is perfectly consistent with the ISM variability reconstructed by a stalagmite δ18O record from the Qunf Cave in southern Oman (Fleitmann et al., 2003) (Figure 3e) and a G. bulloides percentage record from the Arabian Sea (Gupta et al., 2003). Several intervals of strong aeolian accumulation found in this study, subject to dating error, coincide with pauses of weak summer monsoons despite the two weak monsoon events during the early Holocene. The strongest dust sedimentation at 3.1-1.8 ka is associated with a period of hiatus of stalagmite growth in the Qunf Cave (Fleitmann et al., 2003) and a period with the most positive δ18O values in the Dongge Cave (Wang et al., 2005), suggesting that severe drought may have caused this long-lasting increased dust deposition. Generally, the similarity of our results with above ISM records suggests that the regional rainfall in the southern TP, controlled by ISM circulations, may play a key role in regulating regional aeolian activity during the Holocene. Several previous studies have also found a strong correlation between aeolian activity on the TP and summer monsoon intensity (Thompson et al., 2000, Qiang et al., 2014). An ice core record from Dasuopu, in the Himalayas (Thompson et al., 2000), showed that dust storm frequency has increased over the last 200 years, consistent with reduced ISM rainfall (Chu et al., 2011). In general, weaker summer monsoon winds and reduced summer rainfall can prolong wind erosion periods and also lead to degraded vegetation cover, which can facilitate aeolian erosion and dust accumulation.

    Different views exist on the main factors that control aeolian deposition on the TP. The formation of aeolian sediments requires/depends on the availability of dust sources, wind transportation, and suitable environment for dust deposition (Pye, 1995). Previous studies on the northeastern TP (Lu et al., 2011; Chen et al., 2016) have interpreted the accumulation of aeolian sand and loess as an indicator of a relatively dry environment that ensures an adequate dust supply. However, Stauch (2015) found that increased accumulation of dust occurred during the late Glacial and early Holocene across the TP, suggesting that vegetation cover, as a dust trap, may play a key role in the formation of these deposits. Sun et al. (2007), based on OSL ages of aeolian loess in the YZR valley, found that in the southern TP loess deposits had a basal age of 13-11 ka and concluded that the absence of full glacial loess is mainly due to minimal vegetation cover. In contrast, Lai et al. (2009) found older aeolian deposits in a YZR tributary and defined several intervals of increased sediment accumulation at approximately 16-21, 33, and 79-83 ka. The two aeolian sediment accumulation events at approximately 3 and 8 ka, which are identified in this study, have been explained to be associated with two drought events in response to consistent global climate changes. Based on a relatively larger dataset, we find five intervals of increased aeolian sediment accumulation in the middle reaches of the YZR during the Holocene, which tend to be associated with weak ISM events. The coincidence of dust deposition events with dry periods suggests that the predominant driver of aeolian sedimentation may be the climatic controlled dust emission, rather than vegetation cover. Recently, several studies have found footprints of permanent human occupation in central (Meyer et al., 2017; Zhang et al., 2018) and the northeastern (Chen et al., 2015, 2019) TP during the Holocene and earlier periods. As a result, the probable impact of human activity on aeolian activity during the Holocene cannot be ruled out. Due to a dearth of information on the extent of human activity in ancient southern TP, it is difficult to ascertain the extent to which human activity may have affected Holocene aeolian sand accumulations.

    4 Conclusions

    In this paper, we constructed a probability density curve of the ages of aeolian sand and loess in the middle reaches of the YZR, reflecting regional aeolian sediment accumulation intensity. We achieved this using a dataset that comprises new and prior published ages of Holocene aeolian sand and loess. In contrast to related studies on the northeastern TP, we found weak aeolian activities during the early Holocene in this region, and this may be associated with intensified ISM rainfall as a result of high summer insolation. Our results showed several intervals of increased dust accumulation centered at 8.5-7.8, 6.4-5.8, 4.5-4.0, 3.1-1.8, and 0.9 ka. These events are consistent with regional low rainfall events and weak ISM events. Regional monsoon rainfall/wetness, dominated by ISM activity, may play a key role in modulating dust emissions and aeolian deposition in the middle reaches of the YZR at centennial timescales. Overall, this study shows that aeolian deposits in the monsoon dominated TP can provide valuable information on past ISM activities. In the future, more effects of ISM activity would be identified with the availability of more data on the ages of aeolian sediments.

    References

    [1] AnZ, Colman SM, ZhouW. Interplay between the Westerlies and Asian monsoon recorded in Lake Qinghai sediments since 32 ka. Scientific Reports, 2, 619(2012).

    [2] AnZ, Kutzbach JE, Prell WL et al. Evolution of Asian monsoons and phased uplift of the Himalaya-Tibetan Plateau since Late Miocene times. Nature, 411, 62-66(2001).

    [3] BergerA, Loutre MF. Insolation values for the climate of the last 10 million years. Quaternary Science Reviews, 10, 297-317(1991).

    [4] Bird BW, Polisar PJ, LeiY. A Tibetan lake sediment record of Holocene Indian summer monsoon variability. Earth and Planetary Science Letters, 399, 92-102(2014).

    [5] B?hnerJ. General climatic controls and topoclimatic variations in Central and High Asia. Boreas, 35, 279-295(2006).

    [6] ChenF, WelkerF, ShenC. A late Middle Pleistocene Denisovan mandible from the Tibetan Plateau. Nature, 569, 409-412(2019).

    [7] ChenF, WuD, ChenJ. Holocene moisture and East Asian summer monsoon evolution in the northeastern Tibetan Plateau recorded by Lake Qinghai and its environs: A review of conflicting proxies. Quaternary Science Reviews, 154, 111-129(2016).

    [8] Chen FH, Dong GH, Zhang DJ et al. Agriculture facilitated permanent human occupation of the Tibetan Plateau after 3600 B.P. Science, 347, 248-250(2015).

    [9] ChuG, SunQ, YangK et al. Evidence for decreasing South Asian summer monsoon in the past 160 years from varved sediment in Lake Xinluhai, Tibetan Plateau. Journal of Geophysical Research, 116, 1-11(2011).

    [10] ChenF, ChenJ, Holmes JA et al. Moisture changes over the last millennium in arid Central Asia: A review, synthesis and comparison with monsoon region. Quaternary Science Reviews, 29, 1055-1068(2010).

    [11] Conroy JL, Hudson AM, Overpeck JT et al. The primacy of multidecadal to centennial variability over late-Holocene forced change of the Asian monsoon on the southern Tibetan Plateau. Earth and Planetary Science Letters, 458, 337-348(2017).

    [12] DoberschützS, FrenzelP, HaberzettlT et al. Monsoonal forcing of Holocene paleoenvironmental change on the central Tibetan Plateau inferred using a sediment record from Lake Nam Co (Xizang, China). Journal of Paleolimnology, 51, 253-266(2014).

    [13] DongZ, HuG, QianG et al. High-altitude aeolian research on the Tibetan Plateau. Reviews of Geophysics, 55, 864-901(2017).

    [14] Feng JL, Hu HP, ChenF. An eolian deposit-buried soil sequence in an alpine soil on the northern Tibetan Plateau: Implications for climate change and carbon sequestration. Geoderma, 266, 14-24(2016).

    [15] FleitmannD, Burns SJ, MudelseeM et al. Holocene forcing of the Indian monsoon recorded in a stalagmite from southern Oman. Science, 300, 1737-1739(2003).

    [16] GasseF, Fontes JC, Van CampoE et al. Holocene environmental changes in Bangong Co basin (Western Tibet). Part 4: Discussion and conclusions. Palaeogeography, Palaeoclimatology, Palaeoecology, 120, 79-92(1996).

    [17] Gupta AK, Anderson DM, Overpeck JT. Abrupt changes in the Asian southwest monsoon during the Holocene and their links to the North Atlantic Ocean. Nature, 421, 354-357(2003).

    [18] HoffmannT, LangA, DikauR. Holocene river activity: Analysing 14C-dated fluvial and colluvial sediments from Germany. Quaternary Science Reviews, 27, 2031-2040(2008).

    [19] HongY, HongB, Lin QH et al. Correlation between Indian Ocean summer monsoon and North Atlantic climate during the Holocene. Earth and Planetary Science Letters, 211, 371-380(2003).

    [20] HouJ, D'Andrea WJ, WangM et al. Influence of the Indian monsoon and the subtropical jet on climate change on the Tibetan Plateau since the late Pleistocene. Quaternary Science Reviews, 163, 84-94(2017).

    [21] HuH, FengJ, ChenF. Sedimentary records of a palaeo-lake in the middle Yarlung Tsangpo: Implications for terrace genesis and outburst flooding. Quaternary Science Reviews, 192, 135-148(2018).

    [22] Hudson AM, Olsen JW, QuadeJ. Radiocarbon dating of interdune paleo-wetland deposits to constrain the age of Mid-to-Late Holocene microlithic artifacts from the Zhongba site, southwestern Qinghai-Tibet Plateau. Geoarchaeology, 29, 33-46(2014).

    [23] KaiserK, LaiZ, SchneiderB et al. Stratigraphy and palaeoenvironmental implications of Pleistocene and Holocene aeolian sediments in the Lhasa area, southern Tibet (China). Palaeogeography, Palaeoclimatology, Palaeoecology, 271, 329-342(2009).

    [24] KramerA, HerzschuhU, MischkeS et al. Holocene treeline shifts and monsoon variability in the Hengduan Mountains (southeastern Tibetan Plateau), implications from palynological investigations. Palaeogeography, Palaeoclimatology, Palaeoecology, 286, 23-41(2010).

    [25] LaiZ, KaiserK, BrücknerH. Luminescence-dated aeolian deposits of late Quaternary age in the southern Tibetan Plateau and their implications for landscape history. Quaternary Research, 72, 421-430(2009).

    [26] LehmkuhlF, KlingeM, Rees-JonesJ et al. Late Quaternary aeolian sedimentation in central and south-eastern Tibet. Quaternary International, 68-71, 117-132(2000).

    [27] LehmkuhlF, SchulteP, ZhaoH et al. Timing and spatial distribution of loess and loess-like sediments in the mountain areas of the northeastern Tibetan Plateau. Catena, 117, 23-33(2014).

    [28] LiS, DongG, ShenJ et al. Formation mechanism and development pattern of aeolian sand landform in Yarlung Zangbo River valley. Science in China Series D: Earth Sciences, 42, 272-284(1999).

    [29] LiS, YangP, DongY et al. Desertification and Its Control in Tibet(2010).

    [30] LiT, RenX, LiaoY. Paleoenvironment analysis of the middle reaches of Yarlung Zangbo River and Changguogou site. Quaternary Sciences, 40, 547-555(2020).

    [31] LiT, WuY, DuS. Geochemical characterization of a Holocene aeolian profile in the Zhongba area (southern Tibet, China) and its paleoclimatic implications. Aeolian Research, 20, 169-175(2016).

    [32] LingZ, JinJ, WuD et al. Aeolian sediments and their paleoenvironmental implication in the Yarlung Zangbo catchment (southern Tibet, China) since MIS3. Acta Geographica Sinica, 74, 2385-2400(2019).

    [33] LingZ, YangS, WangX et al. Spatial-temporal differentiation of eolian sediments in the Yarlung Tsangpo catchment, Tibetan Plateau, and response to global climate change since the Last Glaciation. Geomorphology, 357, 107104(2020).

    [34] LiuX, DongH, YangX. Late Holocene forcing of the Asian winter and summer monsoon as evidenced by proxy records from the northern Qinghai-Tibetan Plateau. Earth and Planetary Science Letters, 280, 276-284(2009).

    [35] LuH, ZhaoC, MasonJ. Holocene climatic changes revealed by aeolian deposits from the Qinghai Lake area (northeastern Qinghai-Tibetan Plateau) and possible forcing mechanisms. Holocene, 21, 297-304(2011).

    [36] MaQ, ZhuL, LüX. Pollen-inferred Holocene vegetation and climate histories in Taro Co, southwestern Tibetan Plateau. Chinese Science Bulletin, 59, 4101-4114(2014).

    [37] MaQ, ZhuL, LüX. Late glacial and Holocene vegetation and climate variations at Lake Tangra Yumco, central Tibetan Plateau. Global and Planetary Change, 174, 16-25(2019).

    [38] Meyer MC, AldenderferM, WangZ. Permanent human occupation of the central Tibetan Plateau in the early Holocene. Science, 355, 64-67(2017).

    [39] MolnarP, Boos WR, Battisti DS. Orographic controls on climate and paleoclimate of Asia: Thermal and mechanical roles for the Tibetan Plateau. Annual Review of Earth and Planetary Sciences, 38, 77-102(2010).

    [40] Murray AS, Wintle AG. Application of the single-aliquot regenerative-dose protocol to the 375°C quartz TL signal. Radiation Measurements, 32, 579-583(2000).

    [41] PanM, WuY, ZhengY. Holocene aeolian activity in the Dinggye area (southern Tibet, China). Aeolian Research, 12, 19-27(2014).

    [42] Prescott JR, Hutton JT. Cosmic ray contribution to dose rates for luminescence and ESR dating: Large depths and long-term time variations. Radiation Measurements, 23, 497-500(1994).

    [43] PyeK. The nature, origin and accumulation of loess. Quaternary Science Reviews, 14, 653-667(1995).

    [44] QiangM, JinY, LiuX. Late Pleistocene and Holocene aeolian sedimentation in Gonghe Basin, northeastern Qinghai-Tibetan Plateau: Variability, processes, and climatic implications. Quaternary Science Reviews, 132, 57-73(2016).

    [45] QiangM, LiuY, JinY. Holocene record of eolian activity from Genggahai Lake, northeastern Qinghai-Tibetan Plateau, China. Geophysical Research Letters, 41, 589-595(2014).

    [46] Reimer PJ, BardE, BaylissA. IntCal13 and Marine13 radiocarbon age calibration curves 0-50,000 years cal BP. Radiocarbon, 55, 1869-1887(2013).

    [47] SekiO, Meyers PA, YamamotoS. Plant-wax hydrogen isotopic evidence for postglacial variations in delivery of precipitation in the monsoon domain of China. Geology, 39, 875-878(2011).

    [48] ShenW, LiH, SunM. Dynamics of aeolian sandy land in the Yarlung Zangbo River basin of Tibet, China from 1975 to 2008. Global and Planetary Change, 86/87, 37-44(2012).

    [49] ShiX, KirbyE, Furlong KP. Rapid and punctuated Late Holocene recession of Siling Co, central Tibet. Quaternary Science Reviews, 172, 15-31(2017).

    [50] SinghviA, BluszczA, BatemanM. Luminescence dating of loess-palaeosol sequences and coversands: Methodological aspects and palaeoclimatic implications. Earth-Science Reviews, 54, 193-211(2001).

    [51] StauchG. Geomorphological and palaeoclimate dynamics recorded by the formation of aeolian archives on the Tibetan Plateau. Earth-Science Reviews, 150, 393-408(2015).

    [52] StauchG. Multi-decadal periods of enhanced aeolian activity on the north-eastern Tibet Plateau during the last 2 ka. Quaternary Science Reviews, 149, 91-101(2016).

    [53] StauchG, IjmkerJ, P?tschS. Aeolian sediments on the north-eastern Tibetan Plateau. Quaternary Science Reviews, 57, 71-84(2012).

    [54] StauchG, LaiZ, LehmkuhlF. Environmental changes during the late Pleistocene and the Holocene in the Gonghe Basin, north-eastern Tibetan Plateau. Palaeogeography, Palaeoclimatology, Palaeoecology, 509, 144-155(2018).

    [55] SunJ, LiS, Muhs DR. Loess sedimentation in Tibet: Provenance, processes, and link with Quaternary glaciations. Quaternary Science Reviews, 26, 2265-2280(2007).

    [56] Thompson LG, YaoT, MosleythompsonE. A high-resolution millennial record of the South Asian monsoon from Himalayan ice cores. Science, 289, 1916-1919(2000).

    [57] TianL, YaoT, MacCluneK. Stable isotopic variations in west China: A consideration of moisture sources. Journal of Geophysical Research: Atmospheres, 112, 1-12(2007).

    [58] Tian LD, Masson-DelmotteV, StievenardM. Tibetan Plateau summer monsoon northward extent revealed by measurements of water stable isotopes. Journal of Geophysical Research Atmospheres, 106, 28081-28088(2001).

    [59] WangY, ChengH, Edwards RL. The Holocene Asian monsoon: Links to solar changes and North Atlantic climate. Science, 308, 854-857(2005).

    [60] Wintle AG, Murray AS. A review of quartz optically stimulated luminescence characteristics and their relevance in single-aliquot regeneration dating protocols. Radiation Measurements, 41, 369-391(2006).

    [61] XieM, ZhuL, PengP et al. Ostracod assemblages and their environmental significance from the lake core of the Nam Co on the Tibetan Plateau 8.4 kaBP. Journal of Geographical Sciences, 19, 387-402(2009).

    [62] YaoT, Masson-DelmotteV, GaoJ et al. A review of climatic controls on δ18O in precipitation over the Tibetan Plateau: Observations and simulations. Reviews of Geophysics, 51, 525-548(2013).

    [63] ZhangJ, Feng JL, HuG. Holocene proglacial loess in the Ranwu valley, southeastern Tibet, and its paleoclimatic implications. Quaternary International, 372, 9-22(2015).

    [64] Zhang JF, XuB, TurnerF et al. Long-term glacier melt fluctuations over the past 2500 yr in monsoonal High Asia revealed by radiocarbon-dated lacustrine pollen concentrates. Geology, 45, 359-362(2017).

    [65] ZhangX, Ha BB, WangS. The earliest human occupation of the high-altitude Tibetan Plateau 40 thousand to 30 thousand years ago. Science, 362, 1049-1051(2018).

    [66] ZhaoY, YuZ, ZhaoW. Holocene vegetation and climate histories in the eastern Tibetan Plateau: Controls by insolation-driven temperature or monsoon-derived precipitation changes?. Quaternary Science Reviews, 30, 1173-1184(2011).

    [67] ZhengY. The response of Holocene aeolian activities to climate change in the typical area of Qinghai-Tibet Plateau: Take the broad valley area in the middle reaches of the Yarlung Zangbo River in south Tibet and the Gonghe basin in Qinghai as examples. Beijing: Beijing Capital Normal University(2009).

    [68] ZhengY, WuY, LiS. Grain-size characteristics of sediments formed since 8600 yr BP in middle reaches of Yarlung Zangbo River in Tibet and their paleoenvironmental significance. Chinese Geographical Science, 19, 113-119(2009).

    [69] ZhuL, LiY, LiB. The ostracod assemblages and their environmental significance in the Chen Co area, southern Tibet in recent 1400 years. Journal of Geographical Sciences, 12, 451-459(2002).

    [70] ZhuL, LüX, WangJ. Climate change on the Tibetan Plateau in response to shifting atmospheric circulation since the LGM. Scientific Reports, 5, 13318(2015).

    [71] ZhuL, WuY, WangJ. Environmental changes since 8.4 ka reflected in the lacustrine core sediments from Nam Co, central Tibetan Plateau, China. The Holocene, 18, 831-839(2008).

    Tuoyu LI, Jifeng ZHANG, Yongqiu WU, Shisong DU, Duowen MO, Yinan LIAO, Zhitong CHEN, Jianbao LIU, Qing LI. Holocene aeolian activities linked to Indian summer monsoon in the middle reaches of the Yarlung Zangbo River[J]. Journal of Geographical Sciences, 2020, 30(12): 2002
    Download Citation