1. Introduction
Marine oil spills refer to the unintended release of petroleum products into the marine environment due to accidents during exploration, transportation, or storage. These incidents result in the rapid discharge of large quantities of crude oil or other petroleum products, causing severe marine pollution [
1,
2]. The sea area from Haikou to Danzhou features dense shipping routes and complex navigational conditions, increasing the likelihood of oil spill incidents. Additionally, advancements in offshore oil and gas exploration and production technologies have facilitated the expansion of Hainan’s marine petroleum industry [
3]. According to the 2022 Hainan Marine Economy Statistical Bulletin, the offshore oil and gas industry’s output value increased by 217% compared to 2015, with 23 operational offshore drilling platforms and annual oil tanker traffic exceeding 4800 voyages. The intensification of production and transportation activities significantly elevates oil spill risks. As a major province for marine aquaculture and tourism, Hainan possesses valuable coastal ecosystems, including 1320 hectares of mangroves, with
Pinctada maxima (silverlip pearl oysters) accounting for 95% of the national population, and 78 pristine beaches along the Haikou–Danzhou coastline. Oil spills in this region could catastrophically impact marine ecosystems, threatening biodiversity, aquaculture, and coastal tourism. Historical data indicate that a 2016 drilling platform leakage in the Beibu Gulf caused direct economic losses of CNY 730 million and a 38% decline in zooplankton biodiversity, underscoring the urgency of oil spill prevention and control. Currently, Hainan’s early warning systems and decision support capabilities for marine ecological disasters remain underdeveloped, necessitating enhanced oil spill forecasting and numerical simulation capabilities. Numerical simulations of oil spill drift and dispersion are critical for understanding spilled oil dynamics [
4]. Predicting trajectories, oil-covered area variations, and environmental impacts provides essential information for ecological protection. Strengthening forecasting capabilities can mitigate adverse effects on marine ecosystems, aquaculture, and tourism industries.
The oil particle model, established using the Lagrangian particle method and hydrodynamic flow fields, exhibits high accuracy and wide applicability [
5]. Marine oil spill research began in the 1960s with Fay theory-based oil film expansion models [
6], followed by advection–diffusion models and oil particle models [
7]. The oil particle model discretizes oil slicks into independent particles, each representing a specific oil volume. These particles move under hydrodynamic forces (e.g., tidal currents and wind fields) [
8], with individual parameter calculations integrated to determine spatiotemporal oil slick distribution [
9]. Researchers globally have applied oil particle models to forecast spills in critical regions. For instance, Yang Hong et al. [
10] developed a model for the Yangtze River Estuary using FVCOM and Lagrangian particle tracking to assess pollution risks in nature reserves under varying wind conditions. Similarly, Sakar [
11] modeled New York Bay using GNOME and ADIOS to evaluate spill risks and analyze tidal wind influences. Sun Yan et al. [
12] studied Zhanjiang Bay with Delft3D Flow and PART models, examining tidal phase and wind effects on spill behavior. While oil particle models are mature tools for spill prediction [
13], they face environmental adaptability limitations: (1) inadequate representation of complex seabed topography, (2) low resolution for tidal fronts, and (3) restricted applicability of structured grids in reef-dense areas [
14]. Compared to traditional structured grid models, FVCOM employs unstructured triangular grids and finite volume methods, enabling flexible adaptation to Hainan’s complex coastline. Its three-dimensional baroclinic approach accurately simulates tidal characteristics at the Qiongzhou Strait’s western entrance.
The revised Marine Environmental Protection Law of China (2023 Edition, Article 47) mandates oil spill emergency forecasting systems for key marine areas. However, Hainan’s disaster warning system has two key shortcomings: (1) a lack of localized high-precision oil slick parameter databases, leading to simulation deviations when using Bohai Sea-calibrated parameters, and (2) insufficient coupling mechanisms for dynamic interactions (e.g., South China Sea monsoons and shelf waves).
This study employs the Lagrangian particle method for its dual advantages: (1) discretization effectively simulates oil spill fragmentation–recombination processes, enhancing accuracy, and (2) random walk algorithms improve turbulent diffusion modeling. To address these challenges, innovations include the following:
- (1)
Establishing a three-dimensional hydrodynamic particle coupled model based on FVCOM, enabling refined representation of complex coastlines through unstructured grids;
- (2)
Conducting offshore drifter and dye experiments to supplement previous oil spill models that lacked validation components;
- (3)
Designing a multi-scenario oil spill simulation matrix to quantitatively analyze the coupling effects of tidal phases, monsoon intensity, and oil spill weathering processes.
This work fills gaps in Hainan’s oil spill forecasting models and provides scientific support for protecting the fragile Haikou–Danzhou marine ecosystem from spill incidents.
2. Materials and Methods
The modeling process of this study consists of four stages:
- (1)
The Data Preparation Stage. This involves integrating multi-source data, including wind fields, tides, and topography, to establish a South China Sea oil property database.
- (2)
Hydrodynamic Model Construction. This involves developing an unstructured grid hydrodynamic model based on FVCOM to resolve the three-dimensional tidal current field.
- (3)
Oil Spill Diffusion Model Development. This involves using the Lagrangian particle method to construct an oil spill diffusion model, incorporating advection–diffusion processes, emulsification–volatilization–dissolution weathering mechanisms, and shoreline adsorption algorithms.
- (4)
Model Optimization and Application. This involves conducting double-blind experiments with drifters and tracers to optimize model parameters, ultimately enabling multi-scenario oil spill risk simulation and emergency response plan generation.
The developed oil spill model accommodates multiple petroleum product types. Its modular core design allows for the input of diverse oil property parameters (e.g., density, viscosity, and volatility) and the adjustment of weathering processes via interface controls. The framework also reserves extensible interfaces to support the future integration of degradation models for emerging oil types, such as biofuels.
The flowchart of the research framework is shown in
Figure 1.
2.1. Hydrodynamic Model
The FVCOM ocean hydrodynamic model is based on unstructured grids and discretized using the finite volume method [
15]. Compared to structured grid models (e.g., ROMS and Delft3D), FVCOM’s triangular mesh accurately resolves the complex coastline of northwestern Hainan Island. This study employs FVCOM to establish a hydrodynamic model for the Haikou–Danzhou sea area and simulates the tidal current field. The following governing equations include the momentum equation and continuity equation [
16]:
where
t is time;
x,
y, and
z represent east, north, and vertical coordinate axes, respectively;
u,
v, and
w are velocity components in the
x,
y, and
z directions;
P is atmospheric pressure;
f is the Coriolis parameter;
g is gravitational acceleration;
Km is the vertical rotation viscosity coefficient; and
Fu and
Fv represent the diffusion terms of horizontal momentum and vertical momentum.
The computational grid of the FVCOM hydrodynamic model is a triangular mesh generated using the Surface-water Modeling System (SMS 10.1) software. To enhance accuracy in the study area, the domain employs nested grids of varying resolutions, as shown in
Figure 2. The coarse grid shoreline data are sourced from the National Oceanic and Atmospheric Administration (NOAA) coastline database, with bathymetry derived from NOAA’s ETOPO global terrain model. The coarse grid spans 105.663–113.542° E and 15.989–21.913° N, comprising 46,901 nodes and 91,067 triangular elements. Due to the steep bathymetric gradient in the Qiongzhou Strait, grid resolution is refined in this region. High-precision fine grid coastline data were processed using ArcGIS 10.8 software (Esri) with Landsat 8 remote sensing imagery. Fine grid bathymetry is extracted from electronic nautical charts, covering 108.923–110.881° E and 19.503–20.441° N, with 12,710 nodes and 22,903 triangular elements. The Haikou–Danzhou coastal zone, the study’s focus area, features further node densification.
Tidal forcing utilizes harmonic constants for 8 principal constituents (M2, S2, N2, K2, K1, O1, P1, and Q1), 2 long-period constituents (Mf and Mm), and 3 shallow-water constituents (M4, MS4, and MN4) from Oregon State University’s TPXO global tidal model, interpolated to the coarse grid open boundary. Sea surface height and current data from the Copernicus Marine Environment Monitoring Service (CMEMS) are similarly interpolated. Tidal elevation calibration combines boundary harmonic analysis with CMEMS sea surface height. The coarse grid model employs a 20 s time step and cold-start initialization (zero initial velocity and elevation). After stabilization, coarse grid outputs (velocity and elevation) drive the fine grid through boundary interpolation. The fine grid model also uses cold-start initialization, with initial fields interpolated from the coarse grid.
The oil particle model is established based on the Lagrangian particle tracking method and random walk diffusion theory [
17]. The oil film is discretized into a number of particles with Lagrangian characteristics, and the influence of random walks caused by tides, wind fields, and turbulence on the drift of oil particles on the sea surface is considered [
18]. At the same time, in the process of oil particles drifting with the flow, the expansion, emulsification, dissolution, volatilization, and shore arrival processes [
19] are simulated.
2.2. Oil Particle Model
In the oil particle model, uncertainties in input parameters affect prediction accuracy through coupled physical processes. Wind speed deviations directly alter surface oil film drift trajectories, with impacts modulated by the relative orientation of wind and tidal current fields. When the vector angle between these forces exceeds 45°, error propagation exhibits pronounced nonlinearity. Tidal phase uncertainty, particularly in dynamic zones like tidal fronts, is critical: temporal deviations induce advection phase shifts, causing spatial–temporal misalignment of oil film distributions. These errors are amplified during current direction reversal periods.
Among oil property parameters, the emulsification rate demonstrates the highest sensitivity. Minor variations in this rate trigger significant morphological changes in oil films via viscosity expansion positive feedback. Additionally, the temperature-dependent volatilization rate heightens prediction sensitivity to sea surface temperature data accuracy. Combined parameter uncertainties generate sublinear cumulative effects in long-term simulations and may induce phase jumps in prediction errors through oil film thickness-dependent weathering thresholds. While probabilistic ensemble forecasting and data assimilation are widely used to constrain such uncertainties, localized parameter sensitivity experiments remain essential for quantifying their impacts. Existing studies provide substantial insights into oil spill sensitivity analysis, which this work builds upon [
20]. This study establishes an oil particle model based on Lagrangian particle tracking and random walk diffusion theory. The oil film is discretized into Lagrangian-representative particles, accounting for stochastic drift effects from tides, wind fields, and turbulence. During advection, the model simulates oil spill expansion, emulsification, dissolution, volatilization, and shoreline adsorption processes.
2.2.1. Expansion
According to Fay’s theory, the early stage of oil film development can be divided into three stages: the stage of gravity and inertial force, the stage of gravity and viscous force, and the stage of surface tension and viscous force [
21].
where
L is the equivalent diameter centered on the particle;
K1,
K2, and
K3 are empirical coefficients, which are 2.28, 2.90, and 3.20, respectively;
g is the acceleration due to gravity, which is 9.81 kg/m
3;
V is the volume of the oil particle;
, where
is the density of the oil and
is the density of seawater;
, where
,
, and
are the surface tension coefficients between seawater and air, oil and air, and oil and seawater, respectively [
7];
t is time; and
is the kinematic viscosity of seawater.
2.2.2. Drifting
The advection process of particles is affected by tidal currents, circulation, wind fields, etc., and is the main solution of the Lagrangian particle tracking method. The advection velocity of particles is mainly controlled by the flow field and wind field and can be calculated using Equation (8) [
22]:
where
U and
V are the flow velocities of particles in the
x and
y directions, respectively;
Uw and
Vw are the wind speeds in the
x and
y directions 10 m above the sea surface;
Uc and
Vc are the flow velocities of the ocean current in the
x and
y directions; and
is the wind drift coefficient. According to experience, the flow velocity of wind-driven current ranges from 1% to 6% of the wind speed. In this paper,
.
The Lagrangian pursuit method is mainly used to solve a nonlinear differential equation [
16]:
where
x is the position of the particle;
v is the velocity of the particle.
The discrete integration of Equation (10) yields the following:
where
tn is the time when the particle moves to
xn.
The solution using the 4th-order Runge–Kutta method is as follows:
where
;
;
;
;
is the time step; and
xn+1 is the particle position at the next moment.
By substituting Equation (8) into Equation (11), we can obtain the position of the particle in the x and y axis directions at the next moment.
2.2.3. Diffusion
Turbulent diffusion is defined as the mass transfer caused by turbulent pulsation [
23]. The turbulent diffusion process of particles is random, so the random walk method is often used to describe the turbulent diffusion process of particles to simulate the grid-scale turbulence changes in the velocity field. The turbulent diffusion of oil particles can be expressed as follows:
where
Um and
Vm are the turbulent diffusion velocities of particles in the
x-axis and
y-axis directions;
R is a random number between [−1, 1];
Dx and
Dy are the horizontal dispersion coefficients in the
x- and
y-axis directions, both of which are taken as 0.5 m/s; and
is the time step of turbulent diffusion.
2.2.4. Evaporation
Evaporation is the most important mass loss process in the weathering of oil spills. In the first few days, the evaporation rate of light oil can reach 75%, and the evaporation rate of medium oil can reach 40% [
24]. The evaporation rate of oil spills is calculated as follows:
where
t is the time since the oil spill occurred;
KE is the evaporation index, which is related to the mass migration coefficient
KM:
,
, where
A is the oil spill area;
VM is the molar volume;
atm-m
3/(K-mol) is the gas constant;
T is the oil film surface temperature; and
V0 is the oil spill volume. When
, the initial volatile gas pressure
P0 is calculated as follows:
where
T0 is the initial boiling point;
TE·
C is a constant. According to the calculation formula of the American Petroleum Organization,
where
API is
API petroleum gravity.
2.2.5. Dissolution
Although the effect of dissolution on the amount of oil during the weathering process of oil spills is relatively small and the amount of dissolution will not exceed 1% of the total oil volume, the harm caused to marine life by oil being dissolved in seawater is huge [
22], so it is very necessary to calculate the dissolution process of oil spills. The calculation formula for the dissolution rate is as follows:
where
K is the solubility mass transfer coefficient, calculated as
, where
e is determined by the properties of the oil, with values of 1.4 for alkanes, 2.2 for aromatic hydrocarbons, and 1.8 for refined oils.
S is the solubility of oil in water, determined by the type of oil, calculated as follows:
where
S0 is the initial solubility of oil;
is the attenuation coefficient. According to the research of Lu et al. [
25], the initial solubility of crude oil is 8915 mg/m
2, and the attenuation coefficient is 2.38 day
−1.
2.2.6. Emulsification
The emulsification of oil spills refers to the process in which oil spills gradually form stable viscous emulsions under the action of agitation on the sea surface. This emulsion will affect the evaporation and dissolution of oil spills, making the oil spills difficult to clean up [
7]. Generally speaking, the emulsification process of oil spills can be characterized by the water content
YW:
where
KA is generally taken as
;
, where
is the maximum moisture content, which is 0.85 in this paper.
2.2.7. Density and Surface Tension Changes
The evaporation and emulsification processes will affect the density of the remaining oil spill. Combining the two, the final change in density can be expressed as follows [
26]:
where
is the initial density of the oil spill;
is the density of seawater.
The formula for calculating the change in surface tension of oil is
where
is the initial surface tension of the oil film;
is the residual oil volume;
is the emulsified oil volume.
2.2.8. Concentration
The amount of oil in each grid cell is calculated by establishing a structural grid and counting the mass of particles in each grid cell. The amount of oil per square meter is then divided by the area of the grid cell. The calculation formula is
where
C is the concentration of the grid unit;
Mj is the mass of the number of
j particles;
N is the number of particles in the grid unit; and
S is the area of the grid unit.
2.2.9. Landing and Border Processing
Oil spills will run aground on the shore as they move, but due to the limited carrying capacity of the coast, part of the oil spill will return to the seawater with waves and tides. The formula for calculating the amount of oil returned to the seawater over a period of time is as follows [
26]:
where
is the amount of oil returned to the water;
is the amount of oil reaching the shore; and
is the half-life of the shoreline.
When a particle moves to the shore boundary and its position exceeds the shore boundary at the next moment, the particle position is relocated to the intersection of the motion trajectory and the shore, and the particle is considered to be stranded on the shore. Considering the carrying capacity of the coast, a shore boundary grid is set to accommodate 3 shore-reaching particles at most. When a particle crosses the free boundary, it is considered to leave the calculation domain, and its movement is no longer calculated.
2.3. Buoys and Dye Trials
To verify the reliability of the oil spill drift and diffusion model, this study conducted buoy and dye release experiments in the coastal waters from Haikou to Danzhou, simulating the drift and dispersion processes of oil spills under realistic conditions. The offshore experiments were designed according to the “worst-case conditions—typical scenarios” principle, with the trial period selected during the South China Sea’s predominant strong wind season. Historical meteorological data indicate a 63% frequency of northeast winds during this interval, averaging 6.5 m/s and generating maximum wind-driven currents of 0.28 m/s, thereby rigorously testing the model’s predictive capability under extreme conditions. The experimental sites were located within a three-nautical-mile radius of a local oil terminal, an area that had experienced seven minor oil spills over the past five years (averaging 1.4 incidents annually), reflecting its high-risk status.
The trials took place from 18 to 22 December 2024, and included two buoy drift experiments and one dye dispersion test. The first buoy trial commenced at 15:00 on December 18, deploying 15 small GPS buoys at 110.0111° E, 19.9653° N, with data being sampled every 5 min. The second buoy trial began at 15:30 on December 20, releasing 12 buoys at the same location. The dye trial started at 13:30 on December 21 at 109.9677° E, 19.9570° N, monitoring the diffusion range and center position of sodium fluorescein every 10 min for 70 min. Equipment specifications are detailed in
Table 1.
Prior to deployment, wind field conditions were measured using anemometers and cross-validated with local weather forecasts. During the trials, the first buoy experiment experienced northeast winds averaging 5 m/s, while the second buoy trial and dye test were influenced by easterly winds at 6.4 m/s and 5.2 m/s, respectively. Buoys and dye were deployed downwind from predetermined positions (as shown in
Figure 3 and
Figure 4). The process of the three trials is as follows:
- (1)
The first buoy trial: A total of 15 small GPS buoys were deployed. Under the influence of the northeast wind, the buoys moved quickly toward the shore. Two hours after deployment, all the buoys ran aground on the beach southwest of the deployment point. The grounding location of the buoys was located by the computer GPS positioning system, and the buoys were quickly recovered at the corresponding location. However, due to the large number of obstacles in the nearshore area and the damage of some buoys by local fishermen, 12 buoys were successfully recovered in the end.
- (2)
Second buoy trial: A total of 12 small GPS buoys were deployed. Under the influence of the easterly wind, the buoys drifted along the nearshore sea surface of the coastline, maintaining good synchronization as a whole, and presenting a long strip distribution along the coast. A total of 30 h after deployment, all the buoys ran aground on the beach far west of the deployment point. After locating the grounding site through the GPS system, all the buoys were recovered.
- (3)
Dye trial: Because sodium fluorescein has low toxicity and high biodegradability and is not likely to cause long-term environmental pollution, sodium fluorescein was used as a dye in this trial. During the trial, 200 g of sodium fluorescein was mixed with 5 L of seawater to form a uniform solution, and the solution was evenly poured into the sea surface in the downwind direction. The distribution range of sodium fluorescein on the sea surface was photographed and monitored using a drone deployed on the shore, and image data were recorded every 10 min to analyze the drift and diffusion process of sodium fluorescein. After 70 min of release, the color gradually became less obvious due to the decrease in the concentration of sodium fluorescein, and the trial ended.
2.4. Scenario Condition Setting
The sea area from Haikou to Danzhou is located in the southern part of the Qiongzhou Strait. In recent years, the Hainan COSMO terminal in Chengmai Bay has gradually been improved and put into full use. With the increase in sea oil and gas exploitation in Hainan Province, the terminal provides oil and gas supply, maintenance, and cargo loading and unloading services for more ships, resulting in increasingly busy routes in Chengmai Bay, increasing the risk of ship collision and oil spill. According to AIS data, there are many oil tankers with a load of less than 6000 DWT entering and leaving Chengmai Bay. This study selected the dense route point of Chengmai Bay (110.0116° E, 19.9622° N) as the initial point of oil spill, simulated the instantaneous leakage of 100t of medium crude oil, set 2000 oil particles, and, combined with the average seawater temperature in summer and winter, analyzed the impact of oil spill (
Table 2) on five environmentally sensitive areas (ESAs) [
27], which rank from 1 to 5 under different tides and wind conditions as follows: the Silverlip Pearl Oyster Marine Reserve, Huachang Bay Mangrove Marine Reserve, Shatu Bay Beach, Hainan Old Town West Coast Beach, and Dongzhai Qinglan Port Mangrove Marine Reserve (
Figure 5).
Based on ECMWF ERA5 reanalysis data, the wind field characteristics of the sea area in the past 10 years are statistically analyzed (
Figure 6): the dominant wind direction in summer is southerly wind (S), with an average wind speed of 3.55 m/s; the dominant wind direction in winter is northeasterly wind (NE), with an average wind speed of 4.90 m/s. According to the relative position of the oil spill point and the ESA, the most unfavorable wind direction is determined to be easterly wind (E), with a wind speed of 7 m/s. Referring to the
Technical Guidelines for Environmental Risk Assessment of Oil Spills on Water [
28], the simulation time is set to 72 h, and the specific working conditions are shown in
Table 3.