Next Article in Journal
Speed Regulator-Based Path Following Control for an Underactuated Hovercraft Considering the Sideslip Phenomenon
Previous Article in Journal
The Numerical Evaluation of Hydrate Saturation in Marine Sediment During the Injection Process of Self-Heat Generating Fluid
Previous Article in Special Issue
Orbital-Scale Modulation of the Middle Miocene Third-Order Eustatic Sequences from the Northern South China Sea
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Reconstruction of South China Sea Deep Water Salinity During the Last Glacial Maximum (LGM)

1
State Key Laboratory of Deep Earth Processes and Resources, Guangzhou Institute of Geochemistry, Chinese Academy of Sciences, Guangzhou 510640, China
2
University of Chinese Academy of Sciences, Beijing 100049, China
3
GEOMAR Helmholtz Centre for Ocean Research Kiel, 24148 Kiel, Germany
*
Author to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2025, 13(9), 1773; https://doi.org/10.3390/jmse13091773
Submission received: 12 August 2025 / Revised: 9 September 2025 / Accepted: 10 September 2025 / Published: 14 September 2025

Abstract

Reconstructing the deep water salinity during the Last Glacial Maximum (LGM, 26.5~19 ka BP), corresponding to Marine Isotope Stage 2, the most recent and coldest period, is crucial for understanding glacial deep ocean circulation variation and its effect on the climate. The South China Sea (SCS) is one of the largest marginal seas in the western Pacific Ocean, where LGM deep water salinity reconstruction remains unexplored. This study employs pore water [Cl] profiles acquired from boreholes of Site U1499 of IODP Expedition 367 and Sites U1431 and U1433 of IODP Expedition 349 to reconstruct the LGM salinity in the deep SCS. Utilizing a one-dimensional diffusion-advection numerical model, the LGM salinity of the deep northern SCS is determined to be 35.68 ± 0.04 g/kg, and that of the deep central SCS is 35.61 ± 0.03 g/kg, revealing an intra-basin salinity gradient of ~0.07 g/kg. LGM salinity gradients within the SCS were reduced relative to modern ones, indicating attenuated deep circulation within the SCS during the LGM. Furthermore, a diminished salinity gradient (Δ = 0.02 g/kg) across the Luzon Strait between the SCS and Pacific and an enhanced vertical stratification between Upper Circumpolar Deep Water (UCDW) and Lower Circumpolar Deep Water (LCDW) collectively support a sluggish deep Pacific circulation during the LGM.

1. Introduction

From 26.5 to 19 ka BP, corresponding to Marine Isotope Stage 2 (MIS 2), the Earth entered the Last Glacial Maximum (LGM), widely regarded as the coldest geological episode in recent Earth history [1,2,3]. Global temperatures were 3–6 °C lower than the present temperature, and ice sheet volumes reached their maximum extent, resulting in a global sea level drop by 120–135 m compared with the current level [4,5,6,7,8,9]. Atmospheric CO2 concentrations during the LGM were 80–90 ppm lower than pre-industrial levels, a magnitude of change comparable with the observed CO2 increase since the Industrial Revolution [10,11,12,13,14]. Consequently, the LGM serves as a benchmark period for global paleoclimate research, proving essential for understanding Quaternary glacial-interglacial cycles and broader geological-scale climate variability and for informing projections of future climate change.
Currently, the oceans cover approximately 71% of Earth’s surface, with the deep ocean (water depth > 2 km) accounting for over 60% of the global ocean volume. This vast volume of deep seawater plays a crucial role in global matter and heat transport through meridional overturning circulation [15,16]. Excluding the carbon stored in rocks, the dissolved inorganic carbon stored in the deep sea accounts for about 90% of the total dissolved inorganic carbon in all oceans, the atmosphere, and the land organic carbon, which is approximately 60 times that of the atmosphere [17]. Studies of glacial-interglacial cycles demonstrate that atmospheric CO2 concentrations have varied in close correspondence with climate changes over the past 730 ka [18]. These findings collectively suggest that alterations in deep ocean circulation must exert significant impacts on atmospheric carbon budgets and global climate variability. To predict future changes in deep ocean circulation and their climatic impacts, it is helpful to conduct quantitative modeling studies of LGM deep ocean circulation.
Deep ocean circulation is primarily governed by seawater density, which is determined by temperature and salinity. Thus, the critical challenge in modeling lies in reconstructing the paleotemperature and paleosalinity of LGM deep waters [19]. Proxies for reconstructing LGM deep-sea paleotemperatures are relatively abundant (e.g., benthic foraminiferal δ18O values, Mg/Ca ratios, and carbonate clumped isotopes) [20,21,22,23,24,25], whereas proxies for paleosalinity remain scarce, making quantitative reconstruction of paleosalinity significantly more challenging. A breakthrough emerged when McDuff [26] identified a pore water chloride concentration ([Cl]) peak at ~25 m depth in sediment cores from Deep Sea Drilling Program (DSDP) Site 576 in the western equatorial Pacific and proposed that this feature might represent a residual signal of LGM deep seawater migration preserved in the sediments, because water was bound in ice during glaciation and the salt was becoming enriched in the remaining seawater. Pore water [Cl] in marine sediments is well known as a conservative tracer, largely retaining its original seawater signature except under influences such as volcanic ash alteration, methane hydrate formation/dissociation, smectite/illite transformation, halite dissolution, phase separation, or evaporation [27]. Furthermore, seawater [Cl] exhibits a linear relationship with salinity [28]. However, pore water salinity can be modified by early diagenetic processes, such as sulfate reduction or mineral dewatering reactions [29]. Therefore, pore water [Cl] may serve as a viable proxy for reconstructing LGM seawater paleosalinity.
When a [Cl] peak (20–35 m) was identified in the pore water depth profile of deep-sea sediments at Ocean Drilling Program (ODP) Site 1063 near the Bermuda Rise in the Atlantic, Adkins and Schrag [30], inspired by McDuff’s hypothesis, first applied a one-dimensional diffusion-advection model to reconstruct the salinity of LGM deep seawater in this region as 35.76 ± 0.04 g/kg, approximately 0.89 g/kg higher than modern deep water salinity. Subsequent studies have extended this approach to key regions across global oceans by analyzing pore water [Cl] depth profiles and applying the same model: In the North Atlantic, simulated LGM deep water salinities at ODP Sites 981 and 1063 were 36.10 ± 0.10 g/kg and 35.83 ± 0.03 g/kg, respectively. In the Southern Ocean, ODP Site 1093 yielded a simulated salinity of 37.08 ± 0.17 g/kg, indicating significantly saltier LGM deep waters in the Southern Ocean compared with the North Atlantic [29]. This contrasts with the modern salinity gradient, where North Atlantic Deep Water (NADW) is saltier than Southern Ocean water [31], implying a reversed deep water salinity gradient between these basins during the LGM. Homola et al. [32] further constrained LGM deep water salinity in the subtropical North Atlantic (35.51–36.21 g/kg) using relationships between pore water density, [Cl], and salinity. Their results, combined with the water depths of these sites, revealed that LGM NADW salinity increased notably with depth, supporting the hypothesis of reduced vertical mixing and intensified stratification in the glacial Atlantic. The enhanced vertical salinity gradient also suggests that the saltier LGM Atlantic bottom water was sourced from high-salinity Antarctic Bottom Water (AABW), whereas the fresher, shallower intermediate/deep water originated from NADW. In contrast, LGM Pacific deep water salinity exhibited relative homogeneity (36.10 ± 0.10 g/kg) [33], with similar values along deep ocean circulation pathways.
Current reconstructions of LGM deep seawater salinity are largely confined to some regions in the North Atlantic, Southwest Pacific, and Northeast Pacific [29,32,33], with a notable lack of studies in the Western Pacific and Southern Ocean and a complete absence of research on marginal seas. The South China Sea (SCS), one of the largest marginal seas in the Western Pacific, serves as a critical conduit for material and energy exchange between the Pacific and Indian Oceans [34]. Existing LGM salinity studies on the SCS have focused primarily on the surface water of northeastern and southwestern continental-oceanic transition zones [35,36,37], whereas salinity data pertaining to the deep SCS remain scarce. This study addresses these research gaps by analyzing pore water [Cl] profiles from the deep ocean (>3700 m) at International Ocean Discovery Program (IODP) Sites U1499, U1431, and U1433 in the SCS. We reconstructed LGM deep water salinity in the SCS using a one-dimensional diffusion-advection model. Specifically, we elaborated in detail on various parameters of the model, including the porosity, effective diffusion coefficient, advection velocity, and initial and boundary conditions of the model. Based on the simulated salinity results, we further analyzed LGM deep water circulation pathways and flow velocities within the SCS. We also analyzed the LGM salinity gradients cross the Luzon Strait by supplementing Upper Circumpolar Deep Water (UCDW) salinity simulation result, providing insights into the strength of western Pacific deep water intrusion into the SCS during the LGM. The reconstruction of SCS deep water salinity not only fills a critical gap in marginal sea research but also supplements the existing database of UCDW salinity during the LGM. This work contributes to a more comprehensive framework for studying glacial-interglacial changes in thermohaline circulation and their implications for global climate dynamics.

2. General Situation of the SCS and Sites Information

The SCS, located at the junction of the Pacific Plate, the Eurasian Plate, and the Indo-Australian Plate, has emerged as a hotspot of global climate change research in recent years due to its unique geographical position and tectonic evolution. The SCS is a semi-enclosed basin, connected to the deep Pacific Ocean solely through the Luzon Strait between Taiwan Island and Luzon Island, and the deepest sill in the strait is approximately 2600 m [38,39,40,41]. Below this depth, there is no direct seawater exchange between the Pacific and the SCS. On the eastern side of the Luzon Strait, within the depth interval spanning approximately 2000 m to the sill’s maximum depth, the lower potential temperature and higher salinity of Pacific deep water resulting in higher density exceeds that of the SCS [41]. This meridional density difference generates a persistent baroclinic pressure gradient oriented westward across the Luzon Strait, providing the primary forcing mechanism for the overflow from the Pacific into the SCS basin. Consequently, this water sinks after it crosses the Luzon Strait, forming a deep water overflow [42,43]. Therefore, SCS deep water is relatively homogenous and exhibits characteristics similar to Pacific water of 2000~2600 m depth [44]. Once inside the basin, this deep seawater circulation within the SCS constitutes a basin-scale cyclonic gyre, characterized by a counterclockwise flow pattern [39].
We used geophysical and pore water [Cl] data collected from Site U1499 of IODP Expeditions 367 and Sites U1431 and U1433 of IODP Expedition 349 to reconstruct the salinity of LGM deep water in the SCS. IODP 367-Site U1499 (18°24.5698′ N, 115°51.5881′ E) is located on Basement Ridge A within the continent–ocean transition zone of the SCS (Figure 1), at a water depth of 3760 m, with a bottom water temperature of 2.5 °C (potential temperature θ ≈ 2.18 °C). The upper 162.4 m of sediment at Hole U1499A, drilled using an advanced piston corer (APC), comprises dark greenish gray bioclast-rich clay (including foraminifers, nannofossils, radiolarians and diatoms) interbedded with thin clayey silt layers, greenish gray clay-rich calcareous ooze, and dark greenish gray nannofossil-rich clay deposited since the early Pleistocene. Sedimentation rates are stratigraphically variable: average of 9.7 cm/ka in the mid-upper Pleistocene interval (0–48.06 m), ~8.4 cm/ka in the lower Pleistocene interval (100.25–162.4 m), and undeterminable within a slumped section (48–100 m) [45]. The temperature profile demonstrates a linear increase with depth, yielding an anomalously high geothermal gradient of 93 °C/km (Table 1) [46]. IODP 349-Site U1431 (15°22.5379′ N, 117°00.0022′ E) is located near the relict spreading ridge, where the youngest crustal magnetic anomalies are observed in the eastern subbasin of the SCS (Figure 1), at a water depth of 4240 m, with a bottom water temperature of 2.5 °C (θ ≈ 2.13 °C). The upper 168.9 m of Pliocene–Pleistocene sediment, drilled by the APC, consist of dark greenish gray clay, silty clay, and greenish gray nannofossil ooze. Borehole temperature increases linearly with depth, corresponding to a geothermal gradient of 14.8 °C/km, and the sedimentation rate is ~5 cm/ka (Table 1) [47]. IODP 349-Site U1433 (12°55.1380′ N, 115°02.8345′ E) is situated in the southwest subbasin near the relict spreading center (Figure 1), with a water depth of 4380 m. The upper 188.3 m of Pleistocene sediment recovered by the APC primarily comprised dark greenish gray clay with interbedded very thin clayey silt layers. The seafloor temperature is about 2.5 °C (θ ≈ 2.11 °C), and the linear geothermal gradient reaches 78 °C/km. The sedimentation rate at this site is relatively high, reaching 20 cm/ka (Table 1) [48]. Despite the lithological variations among these three sites, the clay-rich sediments are characterized by low permeability and high porosity, which are typical for fine-grained marine deposits. Low permeability results from the small particle size and cohesive nature of clays, which is consistent with the analysis of Reece et al. [49] that high clay content corresponds to low permeability. Conversely, high porosity of ~0.8 near the seafloor decreases rapidly to ~0.5 within the upper 100–150 m due to compaction [46,47,48].

3. Methods

3.1. Sampling and Measuring Methods

Sediment pore water samples were extracted using a titanium squeezer system operated via a hydraulic press, which processed 5–10 cm long whole-round sediment core sections [51,52]. At Site U1431, sampling intervals were set at 0.5–1.5 m for the upper 30 m of the sediment column from Hole U1431A and approximately 10 m below 30 m depth from Hole U1431D, giving a total of 52 samples. At Site U1433, intervals were ~5 m above 70 m and ~10 m below 70 m, giving a total of 25 samples from only Hole U1433A. Building on pore water data acquired from IODP 349, IODP 367 implemented enhanced sampling resolution protocols. For Hole U1499A, sampling intervals were refined as follows: ~3 m above 45 m, ~5 m between 45 and 90 m, and ~10 m below 90 m, resulting in a total of 33 samples. Pore water [Cl] was determined through two analytical approaches: (1) argentometric titration using a Metrohm 785 DMP automatic titrator with silver nitrate (AgNO3) solution for Sites U1431 and U1499, and (2) ion chromatography (IC) performed on a Metrohm 850 Professional IC system for Site U1433. Analytical precision, derived from replicate analyses, yielded relative standard deviations of 0.10% (2σ, Site U1431) and 0.12% (2σ, Site U1499) for titrimetric measurements, and 0.9% for chromatographic determinations [51,52]. All analyses were calibrated against the IAPSO Standard Seawater reference material. Pore water salinity values were calculated from measured [Cl] concentrations using the established stoichiometric relationship: S = [Cl] × 1.80655 × 35.45/1000/ρ, where S (g/kg) represents practical salinity, and ρ (g/cm3) is seawater density [32].

3.2. Diffusion-Advection Numerical Model

To accurately reconstruct the change in deep seawater salinity within the SCS since the LGM and to generate depth-dependent salinity curves in pore water of marine sediments of Sites U1431, U1433, and U1499, we employed a one-dimensional diffusion-advection numerical model following Schrag and DePaolo [53]. The governing partial differential equation is expressed as:
φ C t   =   z φ D e f f C z +   v φ C ,
where C represents the salinity of pore fluids, derived from [Cl], φ denotes the porosity of seafloor sediments as dimensionless volume faction, Deff (cm2/s) is the effective diffusion coefficient of [Cl] in pore water, v (cm/ka) represents the advective velocity of downward burial, t (years) signifies time, and z (cm) represents depth (with the seawater-sediment interface located at z = 0, and positive z indicating depth in the downward direction). Cl transport in pore water is governed by diffusion (first term on the right side) and advection (second term), with the model assuming steady-state sedimentary conditions and excluding secondary diagenetic processes such as volcanic ash diagenesis or methane hydrate dissociation. Subsequent sections detail the methods for determining the porosity, the effective diffusion coefficient, and advection velocity, as well as the initial condition and boundary conditions required to solve Equation (1).

3.2.1. Porosity in Sediments

Porosity (φ) in sediments refers to the total void space per volume of the bulk sediments [54]. In upper fine-grained marine sediments, porosity diminishes exponentially with depth because of sediment compaction. Under steady-state compaction conditions (constant sedimentation rate), the empirical exponential decay of porosity with depth is expressed as [55,56]:
φ z   =   φ + φ 0 φ e β z ,
in which φ denotes the seemingly constant porosity achieved at a certain depth, φ 0 represents the porosity at the seawater-sediment interface, and β is an empirical attenuation constant, with units of cm−1.

3.2.2. Effective Diffusion Coefficient

The effective diffusion coefficient (Deff) represents the diffusion coefficient of Cl in pore water of marine sediments. Its magnitude critically influences the morphology of simulated profiles, particularly affecting both the amplitude and depth of concentration peaks. Consequently, precise determination of Deff is essential for ensuring simulation accuracy [57]. The effective diffusion coefficient is mathematically expressed as Deff = D/θ2, where θ denotes the tortuosity parameter of porosity. For fine-grained sediment lithologies, the tortuosity satisfies the empirical relationship θ 2   =   1 2 × l n ( φ ) [58]. The parameter D denotes the molecular diffusion coefficient of seawater Cl, calculated using the following formula [59]:
D D 0   =   μ 0 μ ,
where D0 is the molecular diffusion coefficient of Cl in pure water, expressed as a function of temperature (T, °C): D0 = (9.60 + 0.438 × T) × 10−6 [60]. The variables μ and μ0 represent the viscosity of seawater and pure water, respectively. The seawater viscosity μ is determined by an empirical equation developed by Mattaus [60], which is a function of temperature, salinity (S, g/kg), and pressure (P, bar). When the salinity is set to 0 in this empirical equation, the resulting viscosity corresponds to that of pure water. Pore water temperature can be calculated from the seafloor temperature and the linear geothermal gradient. The current average salinity of deep seawater in the SCS is ~34.62 g/kg [61,62]. Pressure can be determined based on the seafloor depth and simulated depth.

3.2.3. Advection Velocity

Due to compaction from overlying sediments, some physical properties of marine sediments undergo rapid changes within specific depth intervals. For instance, sediment porosity decreases sharply within the upper 100–200 m before plateauing to a stable state [46,47,48]. Under steady-state compaction and the absence of external overpressure, the vertical advection velocity of pore water is controlled by the downward burial of sediment [60]:
v z   =   ω φ φ ( z ) ,
where ω (cm/ka) represents the sediment burial rate (equivalent to the sedimentation rate) at a depth where the porosity remains constant, and hence v   =   ω .

3.2.4. Initial Condition and Boundary Conditions

If φ, Deff, and v are known, Equation (1) is a parabolic partial differential equation in z and t. To obtain the numerical solution of this equation, we must specify (a) the initial condition, (b) the upper boundary condition at the seawater-sediment interface, and (c) the lower boundary condition at the deepest simulated depth. The initial condition establishes the simulation’s starting time, selected under the principle that it should not affect the magnitude of the peak value of the pore water salinity profile in the simulation. In this simulation, 410 ka BP is chosen, corresponding to Marine Isotope Stage 11 (MIS 11) [63]. Because the sea level at 410 ka BP is similar to the current sea level, the salinity of the current bottom seawater in the SCS is taken as the initial condition (Table 1). The upper boundary condition is defined as a time-dependent salinity curve at the seawater-sediment interface, derived by scaling reconstructed sea level variations since 410 ka BP using the empirically established ratio between sea level change and salinity variation since the LGM [4]. Given a sea level rise of ~130 m in the SCS since the LGM, the model iteratively adjusts trial LGM deep water salinity values to generate a suite of salinity-depth profiles. Optimal matching of simulated curves and measured data is achieved through minimizing the sum of squared difference between measured and simulated data (Chi-square, χ2). The lower boundary condition is determined based on the trend in pore water salinity at the lower simulated sediment intervals, typically assigned as the mean of the last few measured data points or the last pore water salinity value. Equation (1) is solved numerically via the finite difference method by MATLAB R2020a’s “pdede” solver, employing a depth resolution of 50 cm. The solver automatically adapts the internal time step to meet convergence criteria, and the temporal output is provided at intervals of 0.5 ka. This discretization ensures computational stability while resolving millennial-scale hydrodynamic processes.

4. Results

4.1. Simulation Results of Porosity

The compaction of sediments is reflected in the depth-dependent variations in physical properties within the uppermost 100 m at Site U1499, and 150 m at Sites U1431 and U1433. Porosity shows an obvious decrease with depth in this uppermost depth range, from 80–90% to 50%, and remains relatively constant in the deeper sediments [46,47,48]. To maintain consistency, the uppermost 200 m of sediments is selected as the simulation depth of porosity analysis for these three sites. The porosity equation is modeled via MATLAB by lsqnonlin function, yielding φ 0 = 0.851 ± 0.031, φ = 0.498 ± 0.016, β = (2.15 ± 0.43) × 10−4 cm−1, and the Chi-square (χ2) is 0.0525 at Site U1499; φ 0 = 0.794 ± 0.033, φ = 0.523 ± 0.046, β = (1.31 ± 0.59) × 10−4 cm−1, and χ2 = 0.304 at Site U1431; and φ 0 = 0.770 ± 0.031, φ = 0.516 ± 0.034, β = (1.31 ± 0.53) × 10−4 cm−1, and χ2 = 0.191 at Site U1433 (Figure 2).

4.2. Simulation Results of Salinity

The simulated depth intervals for salinity profiles are selected based on sediment cores recovered by the APC, with simulated depths of 162.4 m (U1499), 168.9 m (U1431), and 188.3 m (U1433). Because they are unaffected by secondary processes such as volcanic ash alteration or methane hydrate formation/dissociation, these three sites exhibit “standard” pore water salinity-depth profiles within the modeled intervals. “Standard” salinity profiles display a characteristic pattern: salinity increases gradually from lower values at the sediment-water interface, reaching a distinct peak within the upper 20–50 m of the sediment column. This peak corresponds to the LGM residual signal, with variations in peak magnitude and depth-specific positioning (ranging between sites). The LGM salinity output from the diffusion-advection model is largely controlled by this peak value. Below the peak salinity values, a decreasing trend characterized by different slopes in salinity is observed.
This study utilizes the average pore water salinity from the upper 3 m of sediment column, representing modern bottom water salinity, i.e., initial condition (Table 1). The upper boundary condition is defined by salinity evolution curves reconstructed from sea level variations since 410 ka. For the lower boundary condition, the terminal measured salinity value is generally selected, except at Site U1433 where the last salinity datum deviates from the trend observed at the base of the sediment column. Consequently, the penultimate measured salinity value is adopted as the lower boundary condition for this site (Table 1). Integrating porosity equations, effective diffusion coefficients and advection velocities into the one-dimensional diffusion-advection numerical model, we assume a suite of LGM salinity with 0.05 g/kg increment firstly, generating a suite of simulated salinity-depth profiles and calculating the sum of the squared difference between modeled values and measured data (χ2). Then, refining the simulation increment of 0.01 g/kg within the minimal χ2 neighborhood, we obtain the salinity corresponding to the χ2 minimum as the LGM salinity value. Modeled uncertainty is based on the difference in the average of differences between the measured [Cl] ± 1 standard deviation and the modeled [Cl]. Finally, we obtained reconstructed LGM deep water salinity of 35.68 ± 0.04 g/kg, 35.61 ± 0.03 g/kg, and 35.63 ± 0.31 g/kg at U1499, U1431, and U1433, respectively (Figure 3). The substantial uncertainty of Site U1433 is primarily attributable to significant measurement errors in [Cl] analysis by ion chromatography. These LGM salinity values exceed modern deep water salinities by 1.05 ± 0.04 g/kg, 1.23 ± 0.03 g/kg, and 1.26 ± 0.31 g/kg, corresponding to relative increases of 3.03 ± 0.12%, 3.58 ± 0.09%, and 3.67 ± 0.90%, respectively.

5. Discussion

5.1. LGM Deep Water Salinity of the SCS

The deep SCS circulation constitutes a basin-scale cyclonic gyre, characterized by a counterclockwise flow pattern (Figure 1). Modern low-temperature and high-salinity Pacific deep water intrudes into the SCS through the Luzon Strait and moves southwestward along bathymetric contours near the northern continental slope. Influenced by complex submarine topography, this deep current turns southward within the northwest subbasin, flowing along the western boundary of central basin in the SCS. Near the Zhongsha Island, the southward-flowing current bifurcates into two branches. The southward branch continues flowing along the western boundary of the basin, forming a weak subbasin-scale cyclonic circulation within the southwest subbasin, and the eastward branch, namely the primary current, develops a subbasin-scale cyclonic circulation around seamounts in the central SCS modulated by internal waves, mesoscale eddies, and complex topography. Ultimately, these two branches converge at the basin’s southern periphery, forming a very weak northward-flowing current along the eastern boundary of the SCS central basin [64,65,66,67]. Consequently, the salinity of deep SCS decreases progressively along the deep circulation pathway. Deep water salinity at Site U1499 in the northern SCS is ~34.63 g/kg. In contrast, the central SCS, characterized by numerous seamounts and the resultant complex topography, undergoes enhanced vertical mixing of the upper and lower water layers because of topographic upwelling of bottom currents along seamount flanks, resulting in the lowest regional salinity in the deep SCS [61,64]. For example, Sites U1431 and U1433 near seamounts in the central SCS, the deep water salinity are ~34.38 g/kg and 34.37 g/kg, respectively. Consequently, the maximum salinity gradient reaches ~0.26 g/kg from the northern SCS to the central SCS.
During the LGM, the spatial distribution of deep SCS salinity remained analogous to modern patterns, but with an attenuated salinity gradient. Simulated results indicate a maximum salinity gradient of ~0.07 g/kg from the northern SCS to the central SCS. These confirm that the LGM salinity decreased from northern SCS to the central SCS along the modern circulation pathway, indicating the deep water circulation during the LGM is consistent with the present. However, it is noteworthy that the salinity gradient during the LGM was substantially lower than that observed today. Assuming LGM deep water temperature of the SCS more homogenous than modern as the deep open ocean [29,68], then the deep water circulation of SCS was much controlled by salinity. Consequently, the reduced LGM salinity gradient within deep SCS suggests that a much sluggish circulation during the LGM.

5.2. Deep Water Exchange Between the Pacific and SCS During the LGM

Current physical oceanographic studies have established that the deep seawater in the SCS originates from the Southern Ocean. The Southern Ocean deep water (consisting of the UCDW and Lower Circumpolar Deep Water (LCDW)) flows northward along the western boundary of the Pacific Ocean and enters the SCS via the Luzon Strait [69,70,71,72]. The boundary between the LCDW and UCDW in the Pacific is defined at a potential temperature of θ = 1.5 °C, with this boundary situated at an approximate depth of 3000 m [72]. As the maximum sill depth of the Luzon Strait is around 2600 m, the LCDW is precluded from entering the SCS, leaving UCDW as the predominantly constituent of the SCS deep water [38,65]. The Luzon Strait serves as the sole deep-sea conduit connecting the SCS and the western Pacific. Modern hydrographic data show deep water from the western Pacific flows through the Luzon Strait into the SCS at depths exceeding 2000 m [41], with SCS waters at depths greater than 2000 m exhibiting hydrographic properties analogous to Pacific waters at depths ranging from 2000 to 2600 m [42,64,65]. On the eastern flank of the Luzon Strait within the Philippine Sea, deep water salinity averages ~34.65 g/kg at depths 2000–2600 m [73], whereas the interior SCS basin below 2000 m maintains a mean salinity of 34.62 g/kg [64], yielding the modern salinity gradient across the Luzon Strait of ~0.03 g/kg.
During the LGM, the deep water circulation pathway in the Pacific Ocean remained consistent with modern configuration [70,74]. UCDW flowed northward along the western boundary of the Pacific, entering the SCS via the Luzon Strait [71]. Consequently, comparing the salinity differences between the UCDW in the western Pacific and the deep SCS during the LGM, we can elucidate the characteristics of deep water exchange between the western Pacific and SCS. However, up to now, no absolute paleo-salinities of UCDW during LGM are available. This is because low-temperature alteration of young basaltic crust and volcanic detritus within sediments, altering the original pore water [Cl] [75], resulted in no proper data sets for reconstructing LGM salinities of UCDW in western Pacific by one-dimensional diffusion-advection model yet.
Given these constraints, Site 1172 (43°57.5854′ S, 149°55.6961′ E) in the Southern Ocean was selected as an optimal alternative to reconstruct the salinity of the UCDW. Site 1172 is located at a water depth of ~2620 m within UCDW range and represents the upstream source region of the UCDW mass which flows through western Pacific and enters the SCS. The upper 100 m at Site 1172, which primarily consists of white foraminiferal nannofossil ooze and white and light greenish gray nannofossil ooze, is devoid of volcanic ash, ensuring a pristine pore water [Cl] profile unaffected by early diagenetic alteration. The bottom water temperature is 2.46 °C (θ ≈ 2.26 °C), with a linear geothermal gradient of ~46 °C/km and a sedimentation rate of 2–3.2 cm/ka (Table 1) [50]. The simulation initial time is set at 410 ka BP. Due to the absence of pore water [Cl] data in the upper 3 m of the sediment column, the first measured value of 34.71 g/kg is selected as the initial condition, which also falls within the modern average UCDW salinity range of 34.71–34.73 g/kg [76]. The upper boundary condition is defined by a salinity evolution curve reconstructed from sea level change records since 410 ka BP, accounting for a sea level rise of ~130 m since the LGM [4]. The lower boundary condition uses the deepest measured pore water [Cl] value of 34.96 g/kg within the simulated interval. And the porosity equation for the upper 220 m of the sediment column at Site 1172 is modeled as: φ z =   0.574 + ( 0.699 0.574 ) e 0.000241 z . Utilizing this equation and aforementioned conditions, the simulated LGM deep water salinity at Site 1172 is determined to be 35.70 ± 0.14 g/kg (Figure 4).
The simulated salinity result of Site 1172 is close to the LGM deep water salinity at Site U1499 in the SCS, with a smaller salinity gradient of 0.02 g/kg from UCDW to the deep SCS. Combined with more homogeneous deep water temperature during the LGM, salinity dominated potential density variation [29,68]. Consequently, the reduced salinity gradient across the Luzon Strait during the LGM might lead to a weakened potential density gradient, possibly decreasing the overflow flux of western Pacific deep water into the SCS. It is consistent with the interpretation inferred from benthic foraminifera B/Ca ratios and δ13C data [77], radiocarbon activity [78], and δ18O data [79]. The reduced gradients of deep water [CO32−] (reconstructed from benthic foraminifera B/Ca ratios) and δ13C between the Pacific and SCS indicate weakened deep water ventilation and more sluggish Pacific deep water circulation during the glacial relative to the interglacial [77]. Overall, the representative evidence of multiple proxies substantiates that the LGM deep water circulation between the western Pacific and SCS was characterized by sluggish circulation and weakened vertical mixing in deep water, as the evidence of salinity, consistent with attenuated thermohaline forcing under glacial boundary conditions.

5.3. LGM Salinity of the Deep Pacific

At present, salinity of the deep Pacific exhibits a relatively homogeneous distribution. Because of the absence of deep water formation within the Pacific, the majority of deep water originates from the Southern Ocean [71,80,81]. Consequently, the deep Pacific salinity is comparable to that of the Southern Ocean, with an average salinity of ~34.70 g/kg [31,33]. Horizontally, Pacific deep water salinity exhibits minor diminution, with a latitudinal gradient of ~0.02 g/kg observed from the South Pacific to the North Pacific at equivalent depths. Vertically, the salinity of Pacific deep water measures ~34.62 g/kg at a depth of 2000 m and increases with depth gradually, culminating in a salinity of approximately 34.70 g/kg at around 4000 m. The maximum vertical salinity gradient between UCDW and LCDW is estimated to be about 0.08 g/kg [31].
Studies of benthic foraminiferal δ13C, δ18O, and Nd isotopic records [70,74,77,79] demonstrate that the deep water circulation pathways of the Pacific during the LGM remained broadly consistent with the contemporary conditions and exhibited more homogeneous physical and chemical properties of LGM Pacific deep seawater. LGM deep water salinity in the Southwest Pacific, derived from Sites ODP 1123, IODP U1365, and U1370, averaged 36.12 g/kg (Table 2) [29,33], whereas the Northeast Pacific deep water, sampled at Sites ODP 1225, EQP10, and EQP11, yielded a comparable mean salinity of 36.11 g/kg (Table 2) [33]. The LGM salinity gradient of deep Pacific from the Southwest Pacific to the Northeast Pacific is determined to be 0.01 g/kg, smaller than modern latitudinal gradient. These results indicate uniform spatial variation in LGM Pacific deep water salinity along the deep water circulation pathway, further corroborating that the glacial Pacific deep water circulation followed modern-like trajectories and maintained more homogeneous thermohaline properties. The salinity uniformity provides an angle for understanding deep circulation in the Pacific basin across glacial-interglacial cycles.
Notably, these simulation sites referenced above for reconstructing Pacific deep water salinity are situated within the LCDW domain at depths exceeding 3000 m, reconstructing the LGM salinity 35.91–36.25 g/kg for LCDW [33]. In contrast, our simulation results from Sites U1499 and 1172, representing the UCDW salinity characteristics (35.68–35.70 g/kg), reveal a pronounced LGM salinity gradient of 0.21–0.57 g/kg between LCDW and UCDW. This LGM salinity gradient substantially exceeds the modern observed gradient of ~0.08 g/kg [31], indicating significantly enhanced vertical stratification of Pacific deep waters during the LGM. The amplified salinity gradient effectively suppresses vertical mixing of deep water masses and contributes to a more sluggish Pacific deep water circulation. These findings mutually corroborate the “reduced glacial deep water ventilation” conclusion proposed by Wan et al. [77], which was based on the reduced gradients of deep water [CO32−] and δ13C between Pacific and SCS during glacial periods. Collectively, these evidences demonstrate the attenuation of Pacific deep water circulation during the LGM.

6. Conclusions

To fill a knowledge gap regarding LGM salinity data of deep water in the SCS, this study thoroughly elucidates a one-dimensional diffusion-advection model for [Cl] transport in pore water, detailing the calculation methods and applicable conditions for the relevant parameters. By incorporating measured [Cl] data from IODP Sites U1499, U1431, and U1433 in the SCS, we simulated the salinity variations in deep seawater in the SCS during the LGM and compared these findings with the simulation results from the Southern Ocean and Pacific. The main conclusions are as follows:
(1)
The LGM deep water salinity in the northern SCS was 35.68 ± 0.04 g/kg, while in the central SCS, it was 35.61 ± 0.03 g/kg, leading to an intra-basin salinity gradient of ~0.07 g/kg. This gradient value is substantially smaller than the modern gradient of ~0.26 g/kg, demonstrating significantly weakened circulation in the deep SCS during the glacial periods.
(2)
During the LGM, the salinity gradient between UCDW—sourced from the Southern Ocean—and the deep northern SCS was marginally reduced compared to the modern gradient, indicating diminished overflow flux of UCDW into the SCS through the Luzon Strait.
(3)
The LGM salinity gradient between LCDW and UCDW reached 0.21–0.57 g/kg, significantly exceeding the modern gradient of ~0.08 g/kg. Combined with evidence of reduced glacial deep water ventilation in the western Pacific, this enhanced LGM vertical salinity gradient also demonstrates intensified deep ocean stratification and sluggish deep water circulation in the western Pacific during the glacial periods.

Author Contributions

H.W.: methodology, software, interpretation of data, writing—original draft preparation; Y.C.: Conceptualization, methodology, interpretation of data, supervision, writing—review and editing, funding acquisition; M.H.: methodology, software, writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Natural Science Foundation of China (Grant: 91428102).

Data Availability Statement

Data are openly available at https://web.iodp.tamu.edu/OVERVIEW/ (accessed on 6 December 2023).

Acknowledgments

We are grateful for three anonymous reviewers for their insightful comments. We thank the scientific party and technical staff of IODP Expeditions 349 and 367 and Zhen Sun for providing the geological map of the SCS.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Clark, P.U.; Dyke, A.S.; Shakun, J.D.; Carlson, A.E.; Clark, J.; Wohlfarth, B.; Mitrovica, J.X.; Hostetler, S.W.; McCabe, A.M. The Last Glacial Maximum. Science 2009, 325, 710–714. [Google Scholar] [CrossRef]
  2. Lisiecki, L.E.; Raymo, M.E. A Pliocene-Pleistocene stack of 57 globally distributed benthic δ18O records. Paleoceanography 2005, 20, PA1003. [Google Scholar]
  3. Tierney, J.E.; Zhu, J.; King, J.; Malevich, S.B.; Hakim, G.J.; Poulsen, C.J. Glacial cooling and climate sensitivity revisited. Nature 2020, 584, 569–573. [Google Scholar] [CrossRef]
  4. Waelbroeck, C.; Labeyrie, L.; Michel, E.; Duplessy, J.D.; McManus, J.F.; Lambeck, K.; Balbon, E.; Labracherie, M. Sea-level and deep water temperature changes derived from benthic foraminifera isotope records. Quat. Sci. Rev. 2002, 21, 295–305. [Google Scholar] [CrossRef]
  5. Seltzer, A.M.; Ng, J.; Aeschbach, W.; Kipfer, R.; Kulongoski, J.T.; Severinghaus, J.P.; Stute, M. Widespread six degrees Celsius cooling on land during the Last Glacial Maximum. Nature 2021, 593, 228–232. [Google Scholar] [CrossRef] [PubMed]
  6. Annan, J.D.; Hargreaves, J.C.; Mauritsen, T. A new global surface temperature reconstruction for the Last Glacial Maximum. Clim. Past 2022, 18, 1883–1896. [Google Scholar] [CrossRef]
  7. Selzer, A.M.; Davidson, P.W.; Shackleton, S.A.; Nicholson, D.P.; Khatiwala, S. Global Ocean Cooling of 2.3°C During the Last Glacial Maximum. Geophys. Res. Lett. 2024, 51, e2024GL108866. [Google Scholar] [CrossRef]
  8. Lembeck, K.; Rouby, H.; Purcell, A.; Sun, Y.; Sambridge, M. Sea level and global ice volumes from the Last Glacial Maximum to the Holocene. Proc. Natl. Acad. Sci. USA 2014, 111, 15296–15303. [Google Scholar] [CrossRef] [PubMed]
  9. Grant, K.M.; Rohling, E.J.; Bronk Ramsey, C.; Cheng, H.; Edwards, R.L.; Florindo, F.; Heslop, D.; Marra, F.; Roberts, A.P.; Tamisiea, M.E.; et al. Sea-level variability over five glacial cycles. Nat. Commun. 2014, 5, 5076. [Google Scholar] [CrossRef]
  10. Mix, A.C.; Bard, E.; Schneider, R. Environmental processes of the ice age: Land, oceans, glaciers (EPILOG). Quat. Sci. Rev. 2001, 20, 627–657. [Google Scholar] [CrossRef]
  11. Peltier, W.R.; Fairbanks, R.G. Global glacial ice volume and Last Glacial Maximum duration from an extended Barbados sea level record. Quat. Sci. Rev. 2006, 25, 3322–3337. [Google Scholar] [CrossRef]
  12. Petit, J.R.; Jouzel, J.; Raynaud, D.; Barkov, N.I.; Barnola, J.-M.; Basile, I.; Bender, M.; Chappellaz, J.; Davis, M.; Delaygue, G.; et al. Climate and atmospheric history of the past 420, 000 years from the Vostok ice core, Antarctica. Nature 1999, 399, 429–436. [Google Scholar] [CrossRef]
  13. Jouzel, J.; Masson-Delmotte, V.; Cattani, O.; Dreyfus, G.; Falourd, S.; Hoffmann, G.; Minster, B.; Nouet, J.; Barnola, J.-M.; Chappellaz, J.; et al. Orbital and millennial Antarctic climate variability over the past 800, 000 years. Science 2007, 317, 793–796. [Google Scholar] [CrossRef]
  14. Lüthi, D.; Le Floch, M.; Bereiter, B.; Blunier, T.; Barnola, J.-M.; Siegenthaler, U.; Raynaud, D.; Jouzel, J.; Fischer, H.; Kawamura, K.; et al. High-resolution carbon dioxide concentration record 650000–800000 years before present. Nature 2008, 453, 379–382. [Google Scholar] [CrossRef]
  15. Wang, P.; Jian, Z.M.; Liu, Z.F. Interactions between the earth spheres: Deep-sea processes and records (I) research progress and achievement. Adv. Earth Sci. 2006, 21, 331–337. [Google Scholar]
  16. Talley, L.D.; Pickard, G.; Emery, W.J.; Swift, J. Descriptive Physical Oceanography: An Introduction, 6th ed.; Elsevier: Burlingham, MA, USA, 2011; pp. 473–494. [Google Scholar]
  17. Sigman, D.M.; Boyle, E.A. Glacial/interglacial variations in atmospheric carbon dioxide. Nature 2000, 407, 859–869. [Google Scholar] [CrossRef]
  18. Siegenthaler, U.; Stocker, T.F.; Monnin, E.; Lüthi, D.; Schwander, J.; Stauffer, B.; Raynaud, D.; Barnola, J.-M.; Fischer, H.; Masson-Delmotte, V.; et al. Stable carbon cycle-climate relationship during the Late Pleistocene. Science 2005, 310, 1313–1317. [Google Scholar] [CrossRef]
  19. Miller, M.D.; Simons, M.; Adkins, J.F.; Minson, S.E. The information content of pore fluid δ18O and [Cl]. J. Phys. Oceanogr. 2015, 45, 2070–2094. [Google Scholar] [CrossRef]
  20. Marchitto, T.M.; Bryan, S.P.; Curry, W.B.; McCorkle, D.C. Mg/Ca temperature calibration for the benthic foraminifer Cibicidoides pachyderma. Paleoceanography 2007, 22, PA1203. [Google Scholar] [CrossRef]
  21. Bryan, S.P.; Marchitto, T.M. Mg/Ca-temperature proxy in benthic foraminifera: New calibrations from the Florida Straits and a hypothesis regarding Mg/Li. Paleoceanography 2008, 23, PA2220. [Google Scholar] [CrossRef]
  22. Barras, C.; Duplessy, J.-C.; Geslin, E.; Michel, E.; Jorissen, F.J. Calibration of δ18O of cultured benthic foraminiferal calcite as a function of temperature. Biogeosciences 2010, 7, 1349–1356. [Google Scholar] [CrossRef]
  23. Marchitto, T.M.; Curry, W.B.; Lynch-Stieglitz, J.; Bryan, S.P.; Cobb, K.M.; Lund, D.C. Improved oxygen isotope temperature calibrations for cosmopolitan benthic foraminifera. Geochim. Cosmochim. Acta 2014, 130, 1–11. [Google Scholar] [CrossRef]
  24. Meinicke, N.; Ho, S.L.; Hannisdal, B.; Nürnberg, D.; Tripati, A.; Schiebel, R.; Meckler, A.N. A robust calibration of the clumped isotopes to temperature relationship for foraminifers. Geochim. Cosmochim. Acta 2020, 270, 160–183. [Google Scholar] [CrossRef]
  25. Huntington, K.; Petersen, S.V. Frontiers of Carbonate Clumped Isotope Thermometry. Annu. Rev. Earth Planet. Sci. 2023, 51, 611–641. [Google Scholar] [CrossRef]
  26. McDuff, R.E. The chemistry of interstitial waters, Deep Sea Drilling Project Leg 86. In Initial Reports of the Deep Sea Drilling Project; Government Printing Office: Washington, DC, USA, 1984; Volume 86, pp. 675–687. [Google Scholar]
  27. Dählmann, A.; de Lange, G.J. Fluid-sediment interactions at Eastern Mediterrance mud volcanoes: A stable isotope study from ODP Leg 160. Earth Planet. Sci. Lett. 2003, 212, 377–391. [Google Scholar] [CrossRef]
  28. Pilson, M.E.Q. An Introduction to the Chemistry of the Sea, 2nd ed.; Cambridge University Press: London, UK, 2012; pp. 48–52. [Google Scholar]
  29. Adkins, J.F.; McIntyre, K.; Schrag, D.P. The salinity, temperature and δ18O of the Glacial deep ocean. Science 2002, 298, 1769–1773. [Google Scholar] [CrossRef]
  30. Adkins, J.F.; Schrag, D.P. Pore fluid constraints on deep ocean temperature and salinity during the last glacial maximum. Geophys. Res. Lett. 2001, 28, 771–774. [Google Scholar] [CrossRef]
  31. Talley, L.D. Salinity patterns in the ocean. In Encyclopedia of Global Environmental Change; MacCracken, M.C., Perry, J.S., Eds.; John Wiley & Sons, Ltd.: Chichester, UK, 2002; Volume 1, pp. 629–640. [Google Scholar]
  32. Homola, K.; Spivack, A.J.; Murray, R.W.; Pockalny, R.; D’Hondt, S.; Robinson, R. Deep North Atlantic Last Glacial Maximum salinity reconstruction. Paleoceanogr. Paleoclimatol. 2021, 36, e2020PA004088. [Google Scholar] [CrossRef]
  33. Insua, T.L.; Spivack, A.J.; Graham, D.; D’Hondt, S.; Moran, K. Reconstruction of Pacific Ocean bottom water salinity during the Last Glacial Maximum. Geophys. Res. Lett. 2014, 41, 2914–2920. [Google Scholar] [CrossRef]
  34. Qu, T.; Song, Y.; Yamagata, T. An introduction to the South China Sea throughflow: Its dynamics, variability, and application for climate. Dyn. Atmos. Oceans 2009, 47, 3–14. [Google Scholar] [CrossRef]
  35. Steinke, S.; Chiu, H.-Y.; Shen, C.-Z.; Erlenkeuser, H.; Löwemark, L.; Chen, M.-T. On the influence of sea level and monsoon climate on the southern South China Sea freshwater budget over the last 22,000 years. Quat. Sci. Rev. 2006, 25, 1475–1488. [Google Scholar] [CrossRef]
  36. Steinke, S.; Mohtadi, M.; Groeneveld, J.; Lin, L.-C.; Löwemark, L.; Chen, M.-T.; Rendle-Bühring, R. Reconstructing the southern South China Sea upper water column structure since the Last Glacial Maximum: Implications for the East Asian winter monsoon development. Palaeogeography 2010, 25, PA2219. [Google Scholar] [CrossRef]
  37. Chen, Y.; Wang, Y.; Dang, H.; Jian, Z. Hydroclimatic changes in the northeastern South China Sea since the Last Glacial Maximum. Quat. Sci. 2021, 41, 1031–1043. [Google Scholar] [CrossRef]
  38. Jian, Z.; Tian, J.; Huang, W.; Ma, X.; Wan, S. Evolution of the South China Sea basin and the deep circulation. Sci. Technol. Rev. 2020, 38, 52–56. [Google Scholar]
  39. Wang, G.; Tian, J. Origin and development of the deep water in South China Sea. Sci. Technol. Rev. 2020, 38, 21–25. [Google Scholar]
  40. Wang, P. Exploring the deep sea processes in the South China Sea. Sci. Technol. Rev. 2020, 38, 6–20. [Google Scholar]
  41. Zhao, W.; Zhou, C.; Tian, J.; Yang, Q.; Wang, B.; Xie, L.; Qu, T. Deep water circulation in the Luzon Strait. J. Geophys. Res. Oceans 2014, 119, 790–804. [Google Scholar] [CrossRef]
  42. Qu, T.; Girton, J.B.; Whitehead, J.A. Deepwater overflow through Luzon Strait. J. Geophys. Res. 2006, 111, C01002. [Google Scholar] [CrossRef]
  43. Wang, D.; Xiao, J.; Shu, Y.; Xie, Q.; Chen, J.; Wang, Q. Progress on deep circulation and meridional overturning circulation in the South China Sea. Sci. Technol. Rev. 2016, 59, 1827–1833. [Google Scholar] [CrossRef]
  44. Qu, T. Evidence for water exchange between the South China Sea and the Pacific Ocean through the Luzon Strait. Acta Oceanol. Sin. 2002, 21, 175–185. [Google Scholar]
  45. Zhang, Y.; Yi, L.; Ogg, J.O. Pliocene-Pleistocene magneto-cyclostratigraphy of IODP Site U1499 and implications for climate-driven sedimentation in the northern South China Sea. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2019, 527, 118–132. [Google Scholar] [CrossRef]
  46. Sun, Z.; Stock, J.M.; Klaus, A.; Larsen, H.C.; Jian, Z.; Alvarez Zarikian, C.A.; Boaga, J.; Bowden, S.A.; Briais, A.; Chen, Y.; et al. Site U1499. In South China Sea Rifted Margin; Proceedings of the International Ocean Discovery Program; Sun, Z., Jian, Z., Stock, J.M., Larsen, H.C., Klaus, A., Alvarez Zarikian, C.A., the Expedition 367/368 Scientists, Eds.; International Ocean Discovery Program: College Station, TX, USA, 2018; Volume 367/368. [Google Scholar] [CrossRef]
  47. Li, C.-F.; Lin, J.; Kulhanek, D.K.; Williams, T.; Bao, R.; Briais, A.; Brown, E.A.; Chen, Y.; Clift, P.D.; Colwell, F.S.; et al. Site U1431. In Proceedings of the International Ocean Discovery Program, 349: South China Sea Tectonics; Li, C.-F., Lin, J., Kulhanek, D.K., the Expedition 349 Scientists, Eds.; International Ocean Discovery Program: College Station, TX, USA, 2015. [Google Scholar] [CrossRef]
  48. Li, C.-F.; Lin, J.; Kulhanek, D.K.; Williams, T.; Bao, R.; Briais, A.; Brown, E.A.; Chen, Y.; Clift, P.D.; Colwell, F.S.; et al. Site U1433. In Proceedings of the International Ocean Discovery Program, 349: South China Sea Tectonics; Li, C.-F., Lin, J., Kulhanek, D.K., the Expedition 349 Scientists, Eds.; International Ocean Discovery Program: College Station, TX, USA, 2015. [Google Scholar] [CrossRef]
  49. Reece, J.S.; Flemings, P.B.; Dugan, B.; Long, H.; Germaine, J.T. Permeability-porosity relationships of shallow mudstones in the Ursa Basin, northern deepwater Gulf of Mexico. J. Geophys. Res. Solid Earth 2012, 117, B12102. [Google Scholar] [CrossRef]
  50. Exon, N.F.; Kennett, J.P.; Malone, M.J.; Brinkhuis, H.; Chaproniere, G.C.H.; Ennyu, A.; Fothergill, P.; Fuller, M.D.; Grauert, M.; Hill, P.J.; et al. Proceedings of the Ocean Drilling Program, Initial Reports Volume 189, Site 1172; International Ocean Discovery Program: College Station, TX, USA, 2001. [Google Scholar] [CrossRef]
  51. Li, C.-F.; Lin, J.; Kulhanek, D.K.; Williams, T.; Bao, R.; Briais, A.; Brown, E.A.; Chen, Y.; Clift, P.D.; Colwell, F.S.; et al. Methods. In Proceedings of the Integrated Ocean Discovery Program, 349: South China Sea Tectonics; Li, C.-F., Lin, J., Kulhanek, D.K., the Expedition 349 Scientists, Eds.; International Ocean Discovery Program: College Station, TX, USA, 2015. [Google Scholar] [CrossRef]
  52. Sun, Z.; Jian, Z.; Stock, J.M.; Larsen, H.C.; Klaus, A.; Alvarez Zarikian, C.A.; Boaga, J.; Bowden, S.A.; Briais, A.; Chen, Y.; et al. Expedition 367/368 methods. In South China Sea Rifted Margin; Proceedings of the International Ocean Discovery Program; Sun, Z., Jian, Z., Stock, J.M., Larsen, H.C., Klaus, A., Alvarez Zarikian, C.A., the Expedition 367/368 Scientists, Eds.; International Ocean Discovery Program: College Station, TX, USA, 2018; Volume 367/368. [Google Scholar] [CrossRef]
  53. Schrag, D.P.; DePaolo, D.J. Determination of δ18O of seawater in the deep ocean during the Last Glacial Maximum. Paleoceanography 1993, 8, 1–6. [Google Scholar]
  54. Berner, R.A. Early Diagenesis; Princeton University Press: Princeton, NJ, USA, 1980; p. 20. [Google Scholar]
  55. Boudreau, B.P.; Bennett, R.H. New rheological and porosity equations for steady-state compaction. Am. J. Sci. 1999, 299, 517–528. [Google Scholar] [CrossRef]
  56. Wu, L.; Zeng, Z.; Yin, X.; Wang, X.; Chen, S. Study on the porosity of a metalliferous sediment core from the East Pacific Rise 13°N. Mar. Sci. 2012, 36, 81–86. [Google Scholar]
  57. Adkins, J.F.; Schrag, D.P. Reconstructing Last Glacial Maximum bottom water salinities from deep-sea sediment pore fluid profiles. Earth Planet. Sci. Lett. 2003, 216, 109–123. [Google Scholar] [CrossRef]
  58. Boudreau, B.P. The diffusive tortuosity of fine-grained unlithified sediments. Geochim. Cosmochim. Acta 1996, 60, 3139–3142. [Google Scholar]
  59. Li, Y.-H.; Gregory, S. Diffusion of ions in sea water and in deep-sea sediments. Geochim. Cosmochim. Acta 1974, 38, 703–714. [Google Scholar]
  60. Boudreau, B.P. Diagenetic Model and Their Implementation: Modelling Transport and Reactions in Aquatic Sediments; Springer: Berlin, Germany; New York, NY, USA, 1997; pp. 92–115. [Google Scholar]
  61. Tian, J.; Yang, Q.; Zhao, W. Enhanced diapycnal mixing in the South China Sea. J. Phys. Oceanogr. 2009, 39, 3191–3203. [Google Scholar] [CrossRef]
  62. Xie, Q.; Xiao, J.; Wang, D.; Yu, Y. Analysis of deep-layer and bottom circulation in the South China Sea based on eight quasi-global ocean model outputs. Chin. Sci. Bull. 2013, 58, 4000–4011. [Google Scholar] [CrossRef]
  63. Rohling, E.J.; Braun, K.; Grant, K.; Kucera, M.; Roberts, A.P.; Siddall, M.; Trommer, G. Comparison between Holocene and Marine Isotope Stage-11 sea-level histories. Earth Planet. Sci. Lett. 2010, 291, 97–105. [Google Scholar] [CrossRef]
  64. Wang, G.; Xie, S.-P.; Qu, T.; Huang, R.X. Deep South China Sea circulation. Geophys. Res. Lett. 2011, 38, L05601. [Google Scholar] [CrossRef]
  65. Wang, A.; Du, Y.; Peng, S.; Liu, K.; Huang, R.X. Deep water characteristics and circulation in the South China Sea. Deep Sea Res. I Oceanogr. Res. Pap. 2018, 134, 55–63. [Google Scholar] [CrossRef]
  66. Tian, J.; Qu, T. Advances in research on the deep South China Sea circulation. Chin. Sci. Bull. 2012, 57, 1827–1832. [Google Scholar] [CrossRef]
  67. Chang, Y.-T.; Hsu, W.-L.; Tai, J.-H.; Tang, T.Y.; Chang, M.-H.; Chao, S.-Y. Cold deep water in the South China Sea. J. Oceanogr 2010, 66, 183–190. [Google Scholar] [CrossRef]
  68. Malone, M.J.; Martin, J.B.; Schonfeld, J.; Ninnemann, U.S.; Nurnburd, D.; White, T.S. The oxygen isotopic composition and temperature of Southern Ocean bottom waters during the last glacial maximum. Earth Planet. Sci. Lett. 2004, 222, 275–283. [Google Scholar] [CrossRef]
  69. Talley, L.D. Closure of the global overturning circulation through the Indian, Pacific, and Southern Oceans: Schematics and transports. Oceanography 2013, 26, 80–97. [Google Scholar] [CrossRef]
  70. Wu, Q.; Colin, C.; Liu, Z.; Douville, E.; Dubois-Dauphin, Q.; Frank, N. New insights into hydrological exchange between the South China Sea and the Western Pacific Ocean based on the Nd isotopic composition of seawater. Deep-Sea Res. II 2015, 122, 25–40. [Google Scholar] [CrossRef]
  71. Germineaud, C.; Cravatte, S.; Sprintall, J.; Alberty, M.S.; Grenier, M.; Ganachaud, A. Deep Pacific circulation: New insights on pathways through the Solomon Sea. Deep-Sea Res. I 2021, 171, 103510. [Google Scholar] [CrossRef]
  72. Amakawa, H.; Yu, T.-L.; Tazoe, H.; Obata, H.; Gamo, T.; Sano, Y.; Shen, C.-C.; Suzuki, K. Neodymium concentration and isotopic composition distributions in the southwestern Indian Ocean and the Indian sector of the Southern Ocean. Chem. Geol. 2019, 511, 190–223. [Google Scholar] [CrossRef]
  73. Wu, J.; Lao, Q.; Chen, F.; Huang, C.; Zhang, S.; Wang, C.; Zhou, F.; Chen, C.; Zhou, X.; Lu, X. Water mass processes between the South China Sea and the Western Pacific through the Luzon Strait: Insights from hydrogen and oxygen isotopes. J. Geophys. Res. Oceans 2021, 126, e2021JC017484. [Google Scholar] [CrossRef]
  74. Duplessy, J.-C.; Labeyrie, L.; Waelbroeck, C. Constraints on the ocean oxygen isotopic enrichment between the Last Glacial Maximum and the Holocene: Paleoceanographic implications. Quat. Sci. Rev. 2002, 21, 315–330. [Google Scholar] [CrossRef]
  75. Rosenthal, Y.; Holbourn, A.E.; Kulhanek, D.K.; the Expedition 363 Scientists. Western Pacific Warm Pool. In Proceedings of the International Ocean Discovery Program; International Ocean Discovery Program: College Station, TX, USA, 2018; Volume 363. [Google Scholar]
  76. Oris, A.H.; Whitworth, T., III. Hydrographic Atlas of the World Ocean Circulation Experiment (WOCE) Volume 1: Southern Ocean; Sparrow, M., Chapman, P., Gould, J., Eds.; International WOCE Project Office: Southampton, UK, 2005. [Google Scholar]
  77. Wan, S.; Jian, Z.; Gong, X.; Dang, H.; Wu, J.; Qiao, P. Deep water [CO32-] and circulation in the South China Sea over the last glacial cycle. Quat. Sci. Rev. 2020, 243, 106499. [Google Scholar] [CrossRef]
  78. Wan, S.; Jian, Z. Deep water exchanges between the South China Sea and the Pacific since the last glacial period. Paleoceanography 2014, 29, 1162–1178. [Google Scholar] [CrossRef]
  79. Wan, S.; Jian, Z.; Dang, H. Deep hydrography of the South China Sea and deep water circulation in the Pacific since the Last Glacial Maximum. Geochem. Geophys. Geosyst. 2018, 19, 1447–1463. [Google Scholar] [CrossRef]
  80. Kawabe, M.; Fujio, S. Pacific Ocean circulation based on observation. J. Oceanogr. 2010, 66, 389–403. [Google Scholar] [CrossRef]
  81. Liu, Y.; Guan, Y.; Lin, Y. One question of the oceanic thermohaline circulation: Whether deepwater formation in the north Pacific. Adv. Earth Sci. 2006, 21, 1185–1192. [Google Scholar]
Figure 1. General schematic diagram of modern deep circulation with some tectonic elements in the SCS. Red asterisks are study sites in the SCS. Yellow lines represent the schematic pathway of deep circulation in the SCS.
Figure 1. General schematic diagram of modern deep circulation with some tectonic elements in the SCS. Red asterisks are study sites in the SCS. Yellow lines represent the schematic pathway of deep circulation in the SCS.
Jmse 13 01773 g001
Figure 2. Porosity profiles of marine sediments changing with depth at Sites U1499, U1431, and U1433, respectively. Black lines are fitted porosity curves, and red dots are measured porosity data.
Figure 2. Porosity profiles of marine sediments changing with depth at Sites U1499, U1431, and U1433, respectively. Black lines are fitted porosity curves, and red dots are measured porosity data.
Jmse 13 01773 g002
Figure 3. Salinity profiles of pore water changing with depth below the seafloor for Sites U1499, U1431, and U1433 in the SCS, respectively. SLGM is modeled LGM salinity value. Optimized modeled salinity curves (black bold lines), salinity curve error (gray dashed lines) produced by SLGM ± 0.1 g/kg for each site, and measured salinity based on chloride concentration (red dots).
Figure 3. Salinity profiles of pore water changing with depth below the seafloor for Sites U1499, U1431, and U1433 in the SCS, respectively. SLGM is modeled LGM salinity value. Optimized modeled salinity curves (black bold lines), salinity curve error (gray dashed lines) produced by SLGM ± 0.1 g/kg for each site, and measured salinity based on chloride concentration (red dots).
Jmse 13 01773 g003
Figure 4. Salinity profiles of pore water changing with depth below the seafloor for Site 1172 in the Southern Ocean. SLGM is the modeled LGM salinity value. Optimized modeled salinity curve (black bold line), salinity curve error (gray dashed lines) produced by SLGM ± 0.1 g/kg, and measured salinity based on chloride concentration (red dots).
Figure 4. Salinity profiles of pore water changing with depth below the seafloor for Site 1172 in the Southern Ocean. SLGM is the modeled LGM salinity value. Optimized modeled salinity curve (black bold line), salinity curve error (gray dashed lines) produced by SLGM ± 0.1 g/kg, and measured salinity based on chloride concentration (red dots).
Jmse 13 01773 g004
Table 1. Site descriptions and model parameters in the numerical simulations of IODP Leg 349—Site U1431 [47], Site U1433 [48], IODP Leg 367—Site U1499 [46], and ODP Leg 189—Site 1172 [50].
Table 1. Site descriptions and model parameters in the numerical simulations of IODP Leg 349—Site U1431 [47], Site U1433 [48], IODP Leg 367—Site U1499 [46], and ODP Leg 189—Site 1172 [50].
SiteIODP U1499IODP U1431IODP U1433ODP 1172 1
Site
description
Latitude18°24.5698′ N15°22.5379′ N12°55.1380′ N43°57.5854′ S
Longitude115°51.5881′ E117°00.0022′ E115°02.8345′ E149°55.6961′ E
Water depth (m)3760424043802620
Bottom water temperature (°C)2.5
(θ ≈ 2.18 °C)
2.5
(θ ≈ 2.13 °C)
2.5
(θ ≈ 2.11 °C)
2.46
(θ ≈ 2.26 °C)
Geothermal gradient (°C/km)9314.87846
Sedimentation rate (cm/ka)8.45203.2
Simulation
parameters
Simulation depth (m)162.4168.9188.3100
Simulation time (ka)410410410410
Porosity (φ0, φ,
β/cm−1)
0.851, 0.498,
2.15 × 10−4
0.794, 0.523,
1.31 × 10−4
0.770, 0.516,
1.31 × 10−4
0.699, 0.574,
2.41 × 10−4
Upper boundary condition (x = 0)sea level
(410 ka~0)
sea level
(410 ka~0)
sea level
(410 ka~0)
sea level
(410 ka~0)
Lower boundary condition (g/kg)34.8934.3934.2534.96
Initial condition (g/kg)34.6334.3834.3734.71
Simulation
results
LGM salinity (g/kg)35.68 ± 0.0435.61 ± 0.0335.63 ± 0.3135.70 ± 0.14
LGM salinity difference with modern salinity (g/kg)1.05 ± 0.041.23 ± 0.031.26 ± 0.310.99 ± 0.14
Relative difference (%) 23.03 ± 0.123.58 ± 0.093.67 ± 0.902.85 ± 0.40
1 Site 1172 was used for comparison with other three sites in the SCS and is located next to Tasmania 2 Relative difference is the ratio of the difference between the LGM and modern salinity to the modern salinity.
Table 2. Basic information and reconstructed results of sites in the Southern Ocean, Pacific Ocean and SCS, and the relative difference in LGM deep water salinity compared with modern salinity.
Table 2. Basic information and reconstructed results of sites in the Southern Ocean, Pacific Ocean and SCS, and the relative difference in LGM deep water salinity compared with modern salinity.
SiteLatitudeLongitudeWater Depth (m)Present Salinity (g/kg)LGM Salinity (g/kg)Relative Difference (%)
IODP U149918°24.57′ N115°51.59′ E376034.6335.68 ± 0.043.03 ± 0.12
IODP U143115°22.54′ N117°00.00′ E424034.3835.61 ± 0.033.58 ± 0.09
IODP U143312°55.14′ N115°02.83′ E438034.3735.63 ± 0.313.67 ± 0.90
ODP 117243°57.59′ S149°55.70′ E262034.7135.70 ± 0.142.85 ± 0.40
IODP U1370 [33]41°51.12′ S153°6.36′ W507434.7135.91 ± 0.093.54 ± 0.3
ODP 1123 [29]41°47.16′ S171°29.94′ W329034.7336.19 ± 0.074.2 ± 0.2
IODP U1365 [33]23°51.06′ S165°38.64′ W569534.7036.25 ± 0.094.53 ± 0.3
ODP 1225 [33]2°46.26′ N110°34.26′ W376034.6936.07 ± 0.094.01 ± 0.3
EQP 10 [33]20°40.98′ N143°21.42′ W541234.7036.21 ± 0.184.41 ± 0.5
EQP 11 [33]30°21.30′ N157°52.26′ W581334.6936.06 ± 0.183.96 ± 0.5
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, H.; Chen, Y.; Haeckel, M. Reconstruction of South China Sea Deep Water Salinity During the Last Glacial Maximum (LGM). J. Mar. Sci. Eng. 2025, 13, 1773. https://doi.org/10.3390/jmse13091773

AMA Style

Wang H, Chen Y, Haeckel M. Reconstruction of South China Sea Deep Water Salinity During the Last Glacial Maximum (LGM). Journal of Marine Science and Engineering. 2025; 13(9):1773. https://doi.org/10.3390/jmse13091773

Chicago/Turabian Style

Wang, Haolan, Yifeng Chen, and Matthias Haeckel. 2025. "Reconstruction of South China Sea Deep Water Salinity During the Last Glacial Maximum (LGM)" Journal of Marine Science and Engineering 13, no. 9: 1773. https://doi.org/10.3390/jmse13091773

APA Style

Wang, H., Chen, Y., & Haeckel, M. (2025). Reconstruction of South China Sea Deep Water Salinity During the Last Glacial Maximum (LGM). Journal of Marine Science and Engineering, 13(9), 1773. https://doi.org/10.3390/jmse13091773

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

Article Metrics

Back to TopTop