Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Journal of Asian Earth Sciences 213 (2021) 104740

Contents lists available at ScienceDirect

Journal of Asian Earth Sciences


journal homepage: www.elsevier.com/locate/jseaes

Using n-alkanes as a proxy to reconstruct sea-level changes in Thale Noi,


the west coast of the Gulf of Thailand
Assuma Sainakum a, Piyada Jittangprasert b, Penjai Sompongchaiyakul c,
Akkaneewut Jirapinyakul a, *
a
Morphology of Earth Surface and Advanced Geohazards in Southeast Asia Research Unit (MESA), Department of Geology, Faculty of Science, Chulalongkorn
University, Bangkok 10330, Thailand
b
Department of Chemistry, Faculty of Science, Srinakharinwirot University, Sukhumvit 23, Bangkok 10110, Thailand
c
Department of Marine Science, Faculty of Science, Chulalongkorn University, Phayathai Rd., Patumwan, Bangkok 10330, Thailand

A R T I C L E I N F O A B S T R A C T

Keywords: The n-alkanes content in the sedimentary sequence from Thale Noi, on the west coast of the Gulf of Thailand, was
n-Alkanes analyzed and used to reconstruct the past sea-level fluctuations. The results were further compared with those
Sea-level changes based on sediment characteristics, loss on ignition, grain size distribution, and pollen analysis from the same core
Mangrove
to investigate the efficiency of n-alkanes as a paleoenvironmental proxy. The analysis of n-alkanes indicated the
The Gulf of Thailand
likely periods of regression and transgression from c. 8300–7950 cal. a BP and from c 7950–7665 cal. a BP that
agrees well with the reconstructed sea-level changes from the sediment characteristics and pollen records. The
steady abundance of mangrove pollen taxa in the sediment core points to a standstill in the sea-level from
7650–6870 cal. a BP. However, the results from the n-alkanes analysis further suggested a sea-level fall at c.
7650–7610 cal. a BP and at c. 7400–6870 cal. a BP, and a sea-level rise at c. 7610–7400 cal. a BP. Furthermore,
these sea-level changes were intervened by short intervals of regression centered at c.7590 and 7470 cal. a BP.

1. Introduction investigate the efficiency of n-alkanes to reconstruct sea-level changes in


the Gulf of Thailand using the sedimentary sequence TLN-CP5.
n-Alkanes have been widely used as proxy for environmental Core TLN-CP5 was taken from Thale Noi, on the west coast of the
reconstruction, due to their source specificity and high resistance to the Gulf of Thailand, during January 2015 (Fig. 1). The glacial isostatic and
early diagenetic alteration (e.g., Meyers and Ishiwatari, 1993; Ankit tectonic adjustments on the sea-level changes in the Gulf of Thailand are
et al., 2017; Zhang et al., 2020). The long-chain odd carbon n-alkanes potentially insignificant, since it is located in a tropical climate and
(C27, C29, C31, C33, and C35) are well-known as source indicators of tectonically stable region (e.g., Tjia, 1992; Horton et al., 2005; Culver
terrestrial organic matter (OM). The submerged and floating aquatic et al., 2015). Therefore, the Gulf of Thailand is a crucially important
plants are the main producers of the mid-chain odd carbon n-alkanes area for the study of sea-level fluctuations (Voris, 2000; Horton et al.,
(C21, C23, and C25) (Ficken et al., 2000), while the short-chain odd 2005; Hanebuth et al., 2011). The sedimentary characteristics, loss on
carbon n-alkanes (C15, C17, and C19) generally originated from algae and ignition, grain size, and pollen analysis in the core TLN-CP5 were used to
phytoplankton (Cranwell et al., 1987; Zhang et al., 2020). Consequently, reconstruct sea-level changes in Thale Noi from 8300 to 7000 cal. a BP
the profile of preserved n-alkanes in sediments is a geochemical record (Chabangborn et al., 2020).
that can be used to assess the sources of OM, which will reflect the According to the sedimentary characteristics and sparsity of pollen,
environmental changes from the surrounding catchment area (Zhang Thale Noi and its surrounding area was likely to have been flooded
et al., 2020). However, the multiple sources of OM in sediments make it before c. 8300 cal. a BP. The abundance of Acrostichum, a back mangrove
difficult for reliable environmental interpretation, and so the validation pollen taxa, suggests that the sea-level subsequently dropped from 8300
of any n-alkane-based analysis for environmental assessment by the use to 8050 cal. a BP. The recolonized mangrove swamp forest reflecting this
of other proxies is necessary. The objective of this study was to transgression can be reconstructed from the replacement with

* Corresponding author.
E-mail address: akkaneewut@gmail.com (A. Jirapinyakul).

https://doi.org/10.1016/j.jseaes.2021.104740
Received 7 July 2020; Received in revised form 7 March 2021; Accepted 8 March 2021
Available online 26 March 2021
1367-9120/© 2021 Elsevier Ltd. All rights reserved.
A. Sainakum et al. Journal of Asian Earth Sciences 213 (2021) 104740

Rhizophora in the mangrove pollen taxa, at 8050 cal. a BP. The continued winter monsoon (Amatayakul and Chomtha, 2017). The mean annual
sea-level rise possibly extended from 8050 to 7650 cal. a BP, and led to a precipitation and temperature in this area are about 2000 mm and 27.9
hiatus due to the coastal erosion. The steady abundance of Rhizophora ◦
C (Amatayakul and Chomtha, 2017). The temperature difference be­
pollen points to a standstill in the sea-level from 7650 to 7000 cal. a BP. tween winter and summer is about 3 ◦ C (Amatayakul and Chomtha,
The abundance of associate mangrove pollen assemblage, especially 2017).
Poaceae, is well consistent with the recent conditions of the Thale Noi,
which is a shallow freshwater swamp (Sompongchaiyakul and Sir­ 3. Methods
inawin, 2007).
Eight sediment cores were taken at intervals of approximately 1 km
2. Regional setting in north-south and east-west transects from Thale Noi, Pattalung Prov­
ince, in January 2015 using a modified Russian corer (7.5 cm diameter,
Thale Noi (7.79◦ N, 100.15◦ E) is located in Pattalung Province, near 1 m length). Sediment cores were taken with a 50-cm overlap at each
the west coast of the Gulf of Thailand (Fig. 1). It is a shallow freshwater coring site in order to achieve a continuous sequence. The sediment
lake of 1.2–2 m water depth (Sompongchaiyakul and Sirinawin, 2007; cores were wrapped in plastic and placed in PVC tubes for transport to
Pradit et al., 2010). The lake is surrounding by a peat swamp forest, the Department of Geology, Chulalongkorn University. The sediment
mainly Melaleuca cajuputi, dense sedge and grass near the lake edge, and cores were described in detail, correlated, constructed composite depth,
the scattered open water with floating/submerged plants further into sub-sampled, and stored in a refrigerator until analyses. Detailed sedi­
the lake (Storrer, 1977; Sompongchaiyakul and Sirinawin, 2007; Pradit mentary descriptions of each core and lithostratigraphic correlation
et al., 2010) (Fig. 2). The water inflow into the lake is from the north, have been presented in Chabangborn et al. (2020).
east, and west canals (Fig. 1). Thale Noi covers an area of c. 25 km2, Since TLN-CP5 was the most extended sequence, and was comprised
which is surrounded by a flat plain of c. 0–2 m above mean sea-level of all the sedimentary units, it was selected for the further analysis of the
(Chaimanee et al., 1986). Limestone mountain and outcrops of clastic loss on ignition, grain size distribution, and pollen analysis (Chabang­
sedimentary rocks are found on the western part of the lake and scat­ born et al., 2020). To validate the environmental reconstruction, TLN-
tered in north of Thale Noi, respectively, (Chaimanee et al., 1986). The CP5 was again selected for the analysis of the n-alkanes. Consecutive
eastern area of the lake is a flat plain and a strand plain next to the 10-cm samples were freeze-dried at the Department of Marine Science,
coastline. The southernmost part of Thale Noi is connected to Songkhla Chulalongkorn University. The dried samples were subsequently ground
Lake, a shallow coastal lagoon, by the Khlong (canal) Nang Riam, and homogenized to increase their available surface area for the lipid
Khlong Ban Klang, and Klong Yuan (Storrer, 1977; Sompongchaiyakul extraction.
and Sirinawin, 2007; Pradit et al., 2010) (Fig. 1b). Thale Noi was The prepared samples were left in 20 ml of a 9:1 (v/v) mixture of
possibly a component of Songkhla Lake that was separated from the sea dichloromethane: methanol, and then placed in an ultrasonic extractor
by sand spits extending along he west coast of the Gulf of Thailand from for 10 min. This procedure was repeated three times and the solvent
Nakorn Si Thammarat to Song Khla Provinces in a north-south direction phases (containing the extracted total lipids) were pooled and evapo­
(Storrer, 1977). rated to approximately 2 ml in a rotary evaporator before being trans­
The climate of Thailand is mainly controlled by the southwest ferred to a glass vial and left to evaporate all of the solvent under room
(summer) and northeast (winter) monsoons (Wohlfarth et al., 2012; temperature. The non-polar fraction was further separated from the total
Chawchai et al., 2013). The summer monsoon transports humid air lipid by silica gel column chromatography, using silica gel at ten-fold the
masses from the Indian Ocean and elevates the amount of precipitation weight of the total lipid weight. The total lipid was transferred to the
in Thailand from May to October (Wohlfarth et al., 2012; Chawchai silica gel column by an auto pipette, and the non-polar (n-alkane)
et al., 2013). In contrast, the northeast monsoon, which carries the cold fraction was eluted using three- to four-column volumes of pure hexane.
and dried air masses from central Asia, prevails from November to The fractions were evaporated by a rotary evaporator before transferring
February and decreases the rainfall amount in Thailand. However, the to a glass vial. Finally, a 1:1 (v/v) hexane: ethyl acetate was added to the
monthly precipitation amount in the study area is approximately 100 n-alkane fraction and sent to Scientific and Technological Research
mm from January to September, and then remarkably increases from Equipment Center (STREC), Chulalongkorn University for further anal­
200 to 500 mm during October–December, which is caused by the ysis by Gas Chromatography-Mass Spectrometer (GC–MS).
transportation of humid air masses from the Gulf of Thailand by the The relative abundances of long-chain n-alkanes, which ranged from

Fig. 1. (a) Location of the study area on the west coast of the Gulf of Thailand and (b) the location of coring points (red circle) discussed in the text. (For inter­
pretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

2
A. Sainakum et al. Journal of Asian Earth Sciences 213 (2021) 104740

Fig. 2. Thale Noi (TLN) is surrounding by dense sedge and grass near the lake edge (a and b) and the scattered open water with floating/submerged plants further
into the lake (c).

C14–35, were calculated from the peak area of the GC–MS chromato­
C25 +C31 + C33
grams. The C26–35 alkanes were here classified as long-chain n-alkanes TAR = . (2)
C15 +C17 + C19
(Cranwell et al., 1987; Meyers, 2003; Ankit et al., 2017; Zhang et al.,
2020), while mid- and short-chain n-alkanes were C20–25 and C14–19, The proxy ratio (Paq) is the discriminate between submerged,
respectively, (Cranwell et al., 1987; Ficken et al., 2000; Meyers, 2003, emergent, and terrestrial input, based on the relative proportion of mid-
Ankit et al., 2017; Zhang et al., 2020). The n-alkanes distribution was and long-chain n-alkanes (Ficken et al., 2000; Ankit et al., 2017; Torres
considered with their indices developed to identify the likely sources of et al., 2014; Zhang et al., 2020), and was derived from Eq. (3),
the OM.
C23 + C25
The average chain length (ACL) and terrestrial/aquatic ratio (TAR) Paq = . (3)
C23 + C25 + C29 + C31
were used to estimate the relative variation in the n-alkanes input from
terrestrial or aquatic sources (Ankit et al., 2017; Zhang et al., 2020). The The degradations were assessed using the carbon preference index of
ACL and TAR were calculated from Eqs. (1) and (2), respectively; the short- and long-chain n-alkanes (CPIS and CPIL) (Grimalt and
∑ Albaigés, 1987; Zhang et al., 2020), which were calculated by Eqs. (4)
(Cn × n)
ACLT = ∑ (1) and (5), respectively;
(Cn )
C15 + C17 + C19
CPI S = , (4)
where Cn is the relative abundance of each n-alkane and n is the carbon C16 + C18 + C20
number, and

3
A. Sainakum et al. Journal of Asian Earth Sciences 213 (2021) 104740

(∑ ∑ )
C25− 31 C25− 31 prevalence of the short-chain homologues are under debate, most of the
CPI L = 0.5 × ∑ odd + ∑ odd (5)
even C24− 30 even C26− 33
studies we reviewed indicated that microorganisms, such as bacteria,
algae, fungi, and yeasts, are the main contributor for these n-alkanes (e.
A CPIS value of >1 that is rich in odd-numbered short-chain n-al­ g., Lu and Meyers, 2009; Aloulou et al., 2010; Grimalt and Albaigés,
kanes, especially C17, are directly produced by phytoplankton, such as 1987; Kim et al., 2018; Duan et al., 2019; Zhang et al., 2020). The odd
diatoms, green algae, dinoflagellates, and cyanobacteria (Zhang et al., carbon-numbered long-chain n-alkanes are characteristic components of
2020). Although the sources of CPIS ≤ 1 with an even carbon-numbered

Fig. 3. Carbon number distributions of n-alkanes in the sedimentary sequence TLN-CP5.

4
A. Sainakum et al. Journal of Asian Earth Sciences 213 (2021) 104740

epicuticular plant waxes in terrestrial plants, including mangrove leaves abundances of the long-chained n-alkanes ranged from 5 to 12.5%, with
(Ankit et al., 2017). However, even carbon-numbered long-chain n-al­ the most abundant components being C28, C25, and C29 for the samples
kanes are produced by microorganisms and recycling of the OM (Mey­ at 3.8, 3.6, and 3.5 m DBWS, respectively, (Fig. 3).
ers, 2003; Chevalier et al., 2015). The ACL, CPIL, and TAR values increased from 24–26, 1.5–1.6, and
The chronology of core TLN-CP5 was based on the 14C dating of 10 1.0–2.5, respectively, while CPIS and Paq decreased from 1.5–0.8 and
sequential samples of plant macro-remains. These dating results were 0.5–0.3, respectively, in this sedimentary unit (Fig. 4).
calibrated with InCal13 (Reimer et al., 2013). An age-depth model was
further constructed by Bacon, the Bayesian statistic base routine, from 4.2. Unit II (c. 3.4 m DBWS)
3.87 to 2.17 m DBWS, and by interpolation from 2.17 to 2.05 m DBWS
(Blaauw and Christen, 2011). The age-depth model construction for this Unit II was characterized by a higher abundance of short-chain n-
core has been discussed in detail before (Chabangborn et al., 2020). The alkanes than in units I and III. The proportion of C16 and C17 n-alkanes
sedimentary sequence TLN-CP5 demonstrated the environmental increased and reached c. 12.5 and 14.5%, respectively, whilst C18 and
changes from 8300 cal. a BP to recent. C19 were detected at approximately 7.5%, and C21, C25, C31, and C35 all
varied from 5 to 8% each (Fig. 3). However, the other homologues were
4. Results generally at less than 5%.
The ACL, CPIL, CPIS, Paq, and TAR values were c. 24, 2, 1, 0.4, and 1,
Core TLN-CP5 is composed of five sedimentary units and 20 sedi­ respectively, (Fig. 4), which were slightly different from those at the top
mentary layers (detailed in Chabangborn et al., 2020), as shown in of unit I.
Fig. 4. The lowermost unit A (3.87–3.68 m DBWS) is a stiff, light grey
clay, which gradually changes to the dark grey clay of unit B 4.3. Unit III (c. 3.3–3.0 m DBWS)
(3.68–3.335 m DBWS). The interstratifications of the dark grey clay with
detrital OM and peat layers can be investigated at 3.505–3.44 and The long-chain n-alkanes, especially C31, in this unit were signifi­
3.44–3.405 m DBWS, respectively. Unit B is overlaid by the peat layer of cantly more abundant compared to other units (Fig. 3). In contrast, the
unit C (3.335–2.985 m DBWS), in which a thin layer of peaty clay has short-chain homologues were found at only approximately 5%. The
been found from 3.11 to 3.085 m DBWS. A sharp boundary marked the relative abundance of long-chain n-alkanes gradually increased from 5%
division between units C and D (2.985–2.17 m DBWS), the latter being for C25 to 14, 24, and 29% in C31 in the samples from 3.3, 3.2, and 3.1 m
the interbedding of the dark grey clay with detrital OM and dark grey DBWS, respectively, (Fig. 3). However, the distribution of n-alkanes in
clay layers. A 21 cm thick of the grey clay layer has been found in unit D the sample from 3.0 m DBWS were different from the other samples in
from 2.48 to 2.275 m DBWS. Unit D gradually changes to the uppermost this unit, with the highest proportion being 12.5% for C27 (Fig. 3).
unit E (2.17–1.50 m DBWS), which is the gyttja interbedded with gyttja Despite that the ACL, CPIL, and TAR values gradually increased and
clay. reached their maximum at 28.8, 3.5, and 9.2, respectively, at c. 3.1 m
However, the sedimentary sequence TLN-CP5 was here divided into DBWS, they declined at the top of unit III (Fig. 4). The variations in the
five units (I–V), according to the n-alkanes distribution (Fig. 3). CPIS and Paq values differed from the other n-alkane indices, which
decreased from 2.4 to 0.8 and 0.2 to 0.1 between 3.3 and 3.1 m DBWS,
4.1. Unit I (c. 3.8–3.5 m DBWS) respectively, and suddenly increased at 3.0 m DBWS (Fig. 4).

Unit I was characterized by the bimodal distribution of n-alkanes, 4.4. Unit IV (c. 2.9–2.0 m DBWS)
with C16 and C17 being the most abundant short-chain homologues and
varied from c. 8–12.5% in all samples (Fig. 3). In this unit, C18, C19, and In contrast to unit III, the short-chain n-alkanes, especially C17 and
C20 homologues, however, were indeterminable. The relative C16, were clearly dominant in unit IV (Fig. 3), except for samples from

Fig. 4. Lithostratigraphy, and profiles of ACL, CPIL, CPIS, Paq, and TAR in the sedimentary sequence TLN-CP5 discussed in the text.

5
A. Sainakum et al. Journal of Asian Earth Sciences 213 (2021) 104740

2.7 and 2.3 m DBWS that demonstrated a bimodal distribution of n-al­ from various sources in the intertidal environments. The primary sour­
kanes (Fig. 3). The C16, C17, and C18 n-alkanes ranged from c. 10–12.5%, ces of the long- and short-chain n-alkanes were therefore the epicutic­
12.5–20%, and 15–22.5% in samples from 2.9–2.8, 2.6–2.4, and 2.2–2.0 ular plant waxes in mangrove leaves (Mead et al., 2005; Ankit et al.,
m DBWS, respectively. However, C14, C15, and C19 were generally at less 2017), and phytoplankton (Cranwell et al., 1987) and their degradation
than 5% in all the samples. The most dominant long-chain n-alkane was by microbial activities (Hedges and Prahl, 1993), respectively. The
C31 at c. 10% from 2.9–2.7 and from 2.2–2.0 m DBWS (Fig. 3). Although distribution of n-alkanes in the OM in this environment can be classified
C29, C31, and C33 were the dominant long-chain n-alkanes from 2.6–2.5 into the three types of the predominance of the long-chain (short-chain)
m and 2.3 m DBWS, their relative abundances were only c. 5–7.5%. The n-alkanes reflecting the increased (decreased) terrestrial input and the
relative abundances of the other long-chain homologues in these sam­ bimodal distribution indicating the terrestrial and marine inputs (He
ples, including those in the sample at 2.4 m DBWS, were generally less et al., 2016).
than 5% (Fig. 3). The bimodal distribution of n-alkanes in unit I (Fig. 3) demonstrated
The ACL, CPIL, CPIS, Paq, and TAR values showed slight changes of that the organic input was produced by both terrestrial and marine
approximately 23, 0.5, 0.6, 0.3, and 1.0, respectively, (Fig. 4). sources from c. 8250–8125 cal. a BP (He et al., 2016). In regards to the
high abundance of C17 with a CPIS > 1, phytoplankton were the po­
4.5. Unit V (c. 1.9–1.55 m DBWS) tential contributor of the short-chain homologues from 8250 to 8175
cal. a BP (Zhang et al., 2020) (Figs. 3 and 3). Then, C16 with a CPIS < 1
The sample from 1.9 m DBWS was characterized by a remarkably subsequently replaced the C17 prevalence, which points to an increased
high abundance of the long-chained n-alkanes, while the short-chain n- microbial degradation at 8125 cal. a BP (Kim et al., 2017; Zhang et al.,
alkanes were found at less than 5% (Fig. 3). The C14–35 homologues 2020) (Figs. 3 and 3). Since the CPIL was approximately 1, the bacterial
showed insignificant differences at c. 7.5%. However, a bimodal distri­ recycling of OM was a potential source of the long-chain n-alkanes at
bution of n-alkanes was recognized from 1.7 to 1.55 m DBWS, where 8250 cal. a BP (Meyers, 2003; Hu et al., 2009; Chevalier et al., 2015;
C17, and C25, C27, C29, and C31 were at c. 10% and the highest abundance Zhang et al., 2020) (Fig. 4).
level for the short- and long-chained n-alkanes, respectively, (Fig. 3). The increased CPIL (~1–1.5) suggested a decreased reworking or an
The ACL and TAR decreased from 26–24 and 6–1, respectively, while increased contribution of terrestrial organic input between 8250 and
CPIL, CPIS, and Paq increased from 0.25–2, 0.6–2.4, and 0.2–0.4, 8125 cal. a BP (Fig. 4). The elevated terrestrial input was supported by
respectively, in this unit (Fig. 4). the increased ACL and TAR, and also the decreased Paq (Ficken et al.,
2000; Mead et al., 2005; Ankit et al., 2017; Zhang et al., 2020) (Fig. 4).
5. Discussion Since the degradation of long-chain n-alkanes in the water column is less
than their short-chain homologues (e.g. Meyers, 1997), the inverse
5.1. Chronology relation between CPIL and CPIS can be explained by the lowering of the
sea-level in this marginal marine environment from 8250 to 8125 cal. a
The sequential dating suggested the gradual deposition of sediments BP.. This interpretation agrees well in time with the reconstructed
at c. 2.3 mm/a between 3.68 and 2.985 m DBWS (c. 8300–7920 cal. a regression based on the sediment characteristic of the stiff light grey clay
BP) (Chabangborn et al., 2020). At c. 2.985 m DBWS, a hiatus was ex­ and the appearance of mangrove pollen taxa in the bottom and top of
pected in regards to the sharp boundary between sedimentary units C unit I (Chabangborn et al., 2020).
and D, and the relatively significant difference in the date between the The preference of the short-chain n-alkanes and the decline in TAR in
close-together samples (c. 7950 cal. a BP at 3.015 m DBWS, and c. 7620 unit II suggests a lower contribution of terrestrial organic input than in
cal. a BP at 2.94 m DBWS) (Fig. 4). Above the hiatus, the sedimentation the period before at 8100 cal. a BP (He et al., 2016; Ankit et al., 2017;
rate increased to approximately 3.3 mm/a between 2.985 and 2.17 m Zhang et al., 2020) (Figs. 3 and 4). This scenario agrees well with the
DBWS (c. 7660–7420 cal. a BP) and declined to 0.3 mm/a between 2.17 increase in submerged plants based on the elevated Paq of ~ 0.4 (Ficken
and 2.05 m DBWS (c 7420–7020 cal. a BP). The decrease in sedimen­ et al., 2000; Mead et al., 2005) (Fig. 4). Alternatively, the dominance of
tation rate at 2.17 m DBWS is consistent with the boundary between long- and short-chain odd carbon-number alkanes on the increases in
sedimentary units D and E, gradual increase in loss on ignition and grain CPIL and CPIS, together with the abundance of C17, could indicate the
size distribution (Chabangborn et al., 2020). Since the radiocarbon replacement of the eutrophic conditions caused by the continued
dating of the sample at c. 1.61 m DBWS was modern, the presence of the decrease in the sea-level at 8100 cal. a BP (Meyers, 1997; Lu and Meyers,
upper hiatus was estimated at 2.05 m DBWS, which is the boundary 2009; He et al., 2016; Zhang et al., 2020) (Fig. 4). The fall in sea-level
between the sedimentary layer (Chabangborn et al., 2020). allowed the back mangrove plants, especially Acrostichum aureum, to
However, this new investigation demonstrated that the predomi­ thrive (Chabangborn et al., 2020).
nance of short-chain n-alkanes in the samples from 2.1 and 2.0 m DBWS The distinctively abundant long-chain n-alkanes, and the gradual
was different from the distribution in the sample at 1.9 m DBWS (Fig. 3). increase to a maximum value of the ACL, CPIL, and TAR in unit III
This result, and the lack of samples for radiocarbon dating between 2.06 indicated a high terrestrial organic input from 8050 to 7950 cal. a BP
and 1.6 m DBWS, made us assume that the upper hiatus was at (He et al., 2016; Ankit et al., 2017; Zhang et al., 2020) (Fig. 4). This
approximately 1.95 m DBWS instead of 2.05 m DBWS. Since the dating interpretation corresponds with the gradual decline in Paq from 0.4 to
result of the upper part of core TLN-CP5 is modern, the environmental 0.1 during this interval (Ficken et al., 2000; Mead et al., 2005). More­
reconstruction here was based on the new chronology focused on from c. over, the CPIS reached its maximum value at 8050 cal. a BP before
8250 to 6870 cal. a BP. gradually decreasing from 8050 to 7950 cal. a BP, which potentially
suggested a shift from eutrophic conditions to subaerial/aerial envi­
5.2. Environmental reconstruction based on n-alkanes analysis ronments and so reflect the continued lowering of the sea level. In
contrast, the sea-level rise was reconstructed by the abundance of Rhi­
Although to identify sources of OM based on only the n-alkanes zophora pollens, indicating a recolonizing of the mangrove swamp forest
analysis directly is difficult, the pollen analysis and other supporting (Chabangborn et al., 2020).
information for core TLN-CP5 from Chabangborn et al. (2020) indicate The lead/lag response of proxies to the environmental changes is a
that sea-level fluctuations played a crucial role in the environmental potential cause of the discrepancies in the reconstructed sea-level
changes between a mangrove swamp and back mangrove forests in changes based on n-alkanes and pollen (Wohlfarth et al., 2012).
Thale Noi from 8300 to 7000 cal. a BP. This information allows us to Although the high contribution of terrestrial input can be assessed, the
assume that the OM in this sedimentary sequence was mainly derived sea-level rise began according to the predominance of the mid-chain n-

6
A. Sainakum et al. Journal of Asian Earth Sciences 213 (2021) 104740

alkanes and an abrupt increase in Paq at 7950 cal. a BP (Fig. 3). This and 7470 cal. a BP intervene these sea-level fluctuations.
transgression is a potential cause of the hiatus at 2.985 m DBWS and was
extends from c. 7950 to 7665 cal. a BP (Chabangborn et al., 2020) CRediT authorship contribution statement
(Fig. 4).
In contrast, the preferential short-chain homologues found in unit IV Assuma Sainakum: Methodology, Validation, Formal analysis,
reflect the predominance of marine input from c. 7650 to 6870 cal. a BP Investigation, Writing - original draft, Writing - review & editing,
(He et al., 2016) (Fig. 3). These results, and the domination of C16 and Visualization. Piyada Jittangprasert: Methodology, Validation, Formal
C18 with a CPIS of <1, further indicate that microbial activities in the analysis, Writing - review & editing, Resources. Penjai Sompong­
water column were the main contributor of the short-chain n-alkanes chaiyakul: Conceptualization, Writing - review & editing, Resources.
(Kim et al., 2018; Zhang et al., 2020) (Figs. 3 and 4). The dominance of Akkaneewut Jirapinyakul: Conceptualization, Formal analysis,
mangrove pollen taxa, especially Rhizophora, also demonstrated the Investigation, Writing - original draft, Writing - review & editing,
prolonged flood of the mangrove swamp, which reflects the standstill of Visualization, Supervision, Project administration, Funding acquisition.
the sea-level from c. 7650 to 6870 cal. a BP. However, the n-alkanes
analysis suggests a fluctuation in the sea-level during this interval
(Chabangborn et al., 2020). Declaration of Competing Interest
The decline in the CPIL and CPIS reflects the progressive degradation
of the long- and short-chain n-alkanes (Fig. 4). Furthermore, the The authors declare that they have no known competing financial
decrease in Paq and the relatively high TAR indicate the emergent interests or personal relationships that could have appeared to influence
aquatic macrophyte and high terrestrial organic input, respectively, the work reported in this paper.
(Ficken et al., 2000; Mead et al., 2005) (Fig. 4). These results point to a
sea-level drop from c. 7650 to 7610 cal. a BP. Although the slight Acknowledgements
changes in the TAR, CPIL, and CPIS suggest a steady sea-level, the var­
iations in Paq between 0.3 and 0.4 indicate a sea-level rise from c. 7575 This research was financed through the Ratchadapisek Sompoch
to 7400 cal. a BP and drop from 7400 to 6870 cal. a BP (Ficken et al., Endowment Fund 2018 (CU-GR_61_011_23_004) and Science Super III
2000; Mead et al., 2005; Ankit et al., 2017) (Fig. 4). The inverse rela­ (Department)-009, Chulalongkorn University. The authors are grateful
tionship between CPIL and CPIS supports the later regression from 7400 to the Department of National Parks, Wildlife and Plant Conservation for
to 6870 cal. a BP. In addition, the bimodal distribution of the n-alkanes collecting samples, the Departments of Geology and Marine Science,
indicate the increased terrestrial organic input and the subsequently Faculty of Science, Chulalongkorn University for providing geochemical
short intervals of regression center at c.7590 and 7470 cal. a BP that laboratory and other facilities, Dr. Piyaphong Chenrai, Dr. Paramita
intervened in the standstill of sea-level from c. 7650 to 6870 cal. a BP Punwong and Mr. Niran Chaimanee for helpful discussions, Dr. Robert
(He et al., 2016) (Fig. 4). Butcher for proofreading and manuscript preparation. The anonymous
In the wetland, the OM is mainly originated from the surrounding reviewers are thanked for their useful and constructive comments.
plants, while the pollen is an integrated signal of regional vegetation
(Meyers and Teranes, 2001; Zheng et al., 2009; Huang et al., 2013).
References
Consequently, the differences in sea-level reconstructions based on n-
alkanes and pollen between 7650 and 6870 cal. a BP were possibly Aloulou, F., Kallel, M., Dammak, M., Elleuch, B., Saliot, A., 2010. Even-numbered n-
caused by their transportation pathways (Zheng et al., 2009; Huang alkanes/n-alkenes predominance in surface sediments of Gabes Gulf in Tunisia.
et al., 2013). Environ. Earth Sci. 61, 1–10.
Amatayakul, P., Chomtha, T., 2017. Agricultural meteorology to know for Phatthalung,
The sea-level rose and reached its highstand at c. 6000 a BP, which is
Agrometeorological Division, Meteorological Development Bureau, Thai
a potential cause of the upper hiatus at 1.95 m DBWS (~6870 cal. a BP) Meteorological Department, Thailand, p. 140.
(e.g. Sinsakul, 1992; Choowong et al., 2004). The pollen records above Ankit, Y., Mishra, P.K., Kumar, P., Jha, D.K., Kumar, V.V., Ambili, V., Anoop, A., 2017.
Molecular distribution and carbon isotope of n-alkanes from Ashtamudi Estuary,
the hiatus suggest a replacement of the freshwater swamp forest during
South India: Assessment of organic matter sources and paleoclimatic implications.
the modern period (Chabangborn et al., 2020). The n-alkanes distribu­ Mar. Chem. 196, 62–70.
tion demonstrates the remarkable changes in OM sources from the Blaauw, M., Christen, J.A., 2011. Flexible paleoclimate age-depth models using an
predominance of terrestrial input at 1.9 m DBWS to the terrestrial/ autoregressive gamma process. Bayesian Analy. 6, 457–474.
Chabangborn, A., Punwong, P., Phountong, K., Nudnara, W., Yoojam, N., Sainakum, A.,
aquatic input between 1.9 and 1.55 m DBWS (Fig. 3). The increase in Paq Won-In, K., Sompongchaiyakul, P., 2020. Environmental changes on the west coast
and decrease in TAR suggest a water level rise from 1.9 to 1.55 m DBWS of the Gulf of Thailand during the 8.2 ka event. Quat. Int. 536, 103–113.
(Ficken et al., 2000; Zhang et al., 2020) (Fig. 4). However, the increase Chaimanee, N., Teeyapan, S., Teerarungsigul, N., 1986. Geology of Amphoe Cha-uat and
Amphoe Ranot, Bureau of Geological Survey, Department of Mineral Resources,
in CPIL and CPIS suggest high catchment runoffs and lake productivities Bangkok, Thailand.
(Fig. 4). Chawchai, S., Chabangborn, A., Kylander, M., Löwemark, L., Mörth, C.M., Blaauw, M.,
Klubseang, W., Reimer, P.J., Fritz, S.C., Wohlfarth, B., 2013. Lake Kumphawapi – an
archive of Holocene palaeoenvironmental and palaeoclimatic changes in northeast
6. Conclusions Thailand. Quat. Sci. Rev. 68, 59–75.
Chevalier, N., Savoye, N., Dubois, S., Lama, M.L., David, V., Lecroart, P., le Ménach, K.,
Sediment cores from Thale Noi, which is in the western part of the Budzinski, H., 2015. Precise indices based on n-alkane distribution for quantifying
sources of sedimentary organic matter in coastal systems. Org Geochem. 88, 69–77.
GOT, were analyzed using n-alkanes. The results were compared with Choowong, M., Ugai, H., Charoentitirat, T., Charusiri, P., Daorerk, V., Songmuang, R.,
other records from the same core to investigate the efficiency of n-al­ Ladachart, R., 2004. Holocene biostratigraphical records in coastal deposit from Sam
kanes as a paleoenvironmental proxy. Overall, the reconstructed sea- Roi Yod National Park, Prachuap Khiri Khan, Western Thailand. National History J.
Chulalongkorn Univ. 4, 1–18.
level fluctuations based on n-alkanes agree well with those from sedi­ Cranwell, P.A., Eglinton, G., Robinson, N., 1987. Lipids of aquatic organisms as potential
ment characteristics, loss on ignition, grain size distribution, and pollen contributors to lacustrine sediments—II. Org Geochem. 11, 513–527.
analysis. They suggest a sea-level fall from c. 8300–7950 cal. a BP, and a Culver, S.J., Leorri, E., Mallinson, D.J., Corbett, D.R., Shazili, N.A.M., 2015. Recent
coastal evolution and sea-level rise, Setiu Wetland, Peninsular Malaysia.
gradual rise in sea-level at c 7920–7665 cal. a BP. However, the stand­
Palaeogeogr. Palaeoclimatol. Palaeoecol. 417, 406–421.
still of sea level derived from pollen records differ from the recon­ Duan, L., Song, J., Yuan, H., Li, X., Peng, Q., 2019. Occurrence and origins of biomarker
structed sea-level fluctuations based on n-alkanes from 7650 to 6870 cal. aliphatic hydrocarbons and their indications in surface sediments of the East China
a BP. The n-alkane analysis suggests a regression at c. 7650–7610 cal. a Sea. Ecotoxicol. Environ. Saf. 167, 259–268.
Ficken, K.J., Li, B., Swain, D.L., Eglinton, G., 2000. An n-alkane proxy for the
BP and at c. 7400–6870 cal. a BP, and a transgression at c. 7610–7400 sedimentary input of submerged/floating freshwater aquatic macrophytes. Org
cal. a BP. Moreover, the short intervals of regression centered at c.7590 Geochem. 31, 745–749.

7
A. Sainakum et al. Journal of Asian Earth Sciences 213 (2021) 104740

Grimalt, J., Albaigés, J., 1987. Sources and occurrence of C12–C22 n-alkane distributions Meyers, P.A., Teranes, J.L., 2001. Sediment Organic Matter, in: W.M., L., J.P., S. (Eds.),
with even carbon-number preference in sedimentary environments. Geochim. Tracking Environmental Change Using Lake Sediments. Developments in
Cosmochim. Acta 51, 1379–1384. Paleoenvironmental Research. Springer, Dordrecht.
Hanebuth, T.J.J., Voris, H.K., Yokoyama, Y., Saito, Y., Okuno, J.I., 2011. Formation and Pradit, S., Wattayakorn, G., Angsupanich, S., Baeyens, W., Leermakers, M., 2010.
fate of sedimentary depocentres on Southeast Asia’s Sunda Shelf over the past sea- Distribution of Trace Elements in Sediments and Biota of Songkhla Lake, Southern
level cycle and biogeographic implications. Earth Sci. Rev. 104, 92–110. Thailand. Water Air Soil Pollut. 206, 155–174.
He, J., Zhang, S., Zhang, X., Qian, Y., He, H., Wu, H., 2016. Composition and Distribution Reimer, P.J., Bard, E., Bayliss, A., Beck, J.W., Blackwell, P.G., Bronk Ramsey, C., Buck, C.
Characteristics and Geochemical Significance of n-Alkanes in Core Sediments in the E., Cheng, H., Edwards, R.L., Friedrich, M., Grootes, P.M., Guilderson, T.P.,
Northern Part of the South Yellow Sea. J. Chem. 2016, 1–6. Haflidason, H., Hajdas, I., Hatté, C., Heaton, T.J., Hoffmann, D.L., Hogg, A.G.,
Hedges, J.I., Prahl, F.G., 1993. Early Diagenesis: Consequences for Applications of Hughen, K.A., Kaiser, K.F., Kromer, B., Manning, S.W., Niu, M., Reimer, R.W.,
Molecular Biomarkers. In: Engel, M.H., Macko, S.A. (Eds.), Organic Geochemistry: Richards, D.A., Scott, E.M., Southon, J.R., Staff, R.A., Turney, C.S.M., van der Plicht,
Principles and Applications. Springer, US, Boston, MA, pp. 237–253. J., 2013. IntCal13 and Marine13 Radiocarbon Age Calibration Curves 0–50,000
Horton, B.P., Gibbard, P.L., Mine, G.M., Morley, R.J., Purintavaragul, C., Stargardt, J.M., Years cal BP. 2013, 19.
2005. Holocene sea levels and palaeoenvironments, Malay-Thai Peninsula, southeast Sinsakul, S., 1992. Evidence of quarternary sea level changes in the coastal areas of
Asia. The Holocene 15, 1199–1213. Thailand: a review. J. SE Asian Earth Sci. 7, 23–37.
Hu, J., Peng, P.A., Chivas, A.R., 2009. Molecular biomarker evidence of origins and Sompongchaiyakul, P., Sirinawin, W., 2007. Arsenic, chromium and mercury in surface
transport of organic matter in sediments of the Pearl River estuary and adjacent sediment of Songkhla Lake system. Thailand Asian J. Water, Environ. Pollut. 4,
South China Sea. Appl. Geochem. 24, 1666–1676. 17–24.
Huang, X., Xue, J., Wang, X., Meyers, P.A., Huang, J., Xie, S., 2013. Paleoclimate Storrer, P.J., 1977. A study of the waterflowl at Thale Noi waterfowl reserve area.
influence on early diagenesis of plant triterpenes in the Dajiuhu peatland, central Natural History Bull. Siam Soc. 26, 317–338.
China. Geochim. Cosmochim. Acta 123, 106–119. Tjia, H.D., 1992. Holocene sea-level changes in the Malay-Thai Peninsula, a tectonically
Kim, J.-H., Lee, D.-H., Yoon, S.-H., Jeong, K.-S., Choi, B., Shin, K.-H., 2017. Contribution stable environment. Bull. Geol. Soc. Malaysia 31, 157–176.
of petroleum-derived organic carbon to sedimentary organic carbon pool in the Torres, T., Ortiz, J.E., Martín-Sánchez, D., Arribas, I., Moreno, L., Ballesteros, B.,
eastern Yellow Sea (the northwestern Pacific). Chemosphere 168, 1389–1399. Blázquez, A., Domínguez, J.A., Rodríguez Estrella, T., Martini, I.P., Wanless, H.R.,
Kim, D., Kim, J.-H., Kim, M.-S., Ra, K., Shin, K.-H., 2018. Assessing environmental 2014. The long Pleistocene record from the Pego-Oliva marshland (Alicante-
changes in Lake Shihwa, South Korea, based on distributions and stable carbon Valencia, Spain), Sedimentary Coastal Zones from High to Low Latitudes:
isotopic compositions of n-alkanes. Environ. Pollut. 240, 105–115. Similarities and Differences. Geol. Soc. London 388.
Lu, Y., Meyers, P.A., 2009. Sediment lipid biomarkers as recorders of the contamination Voris, H.K., 2000. Maps of Pleistocene sea levels in Southeast Asia: shorelines, river
and cultural eutrophication of Lake Erie, 1909–2003. Org Geochem. 40, 912–921. systems and time durations. J. Biogeogr. 27, 1153–1167.
Mead, R., Xu, Y., Chong, J., Jaffé, R., 2005. Sediment and soil organic matter source Wohlfarth, B., Klubseang, W., Inthongkaew, S., Fritz, S.C., Blaauw, M., Reimer, P.J.,
assessment as revealed by the molecular distribution and carbon isotopic Chabangborn, A., Löwemark, L., Chawchai, S., 2012. Holocene environmental
composition of n-alkanes. Org Geochem. 36, 363–370. changes in northeast Thailand as reconstructed from a tropical wetland. Global
Meyers, P.A., 1997. Organic geochemical proxies of paleoceanographic, paleolimnologic, Planet. Change 92–93, 148–161.
and paleoclimatic processes. Org Geochem. 27, 213–250. Zhang, Y., Yu, J., Su, Y., Du, Y., Liu, Z., 2020. A comparison of n-alkane contents in
Meyers, P.A., 2003. Applications of organic geochemistry to paleolimnological sediments of five lakes from contrasting environments. Org Geochem. 139, 103943.
reconstructions: a summary of examples from the Laurentian Great Lakes. Org Zheng, Y., Zhou, W., Xie, S., Yu, X., 2009. A comparative study of n-alkane biomarker
Geochem. 34, 261–289. and pollen records: an example from southern China. Chin. Sci. Bull. 54, 1065–1072.
Meyers, P.A., Ishiwatari, R., 1993. Lacustrine organic geochemistry—an overview of
indicators of organic matter sources and diagenesis in lake sediments. Org Geochem.
20, 867–900.

You might also like