553
1 INTRODUCTION
One of the principal challenges in global fisheries is
advancing sustainable fishing practices that minimize
environmental degradation while enhancing societal
advantages One of the most significant environmental
concerns is the high level of fuel consumption in
capture fisheries, which contributes to greenhouse gas
emissions. Many countries subsidize fuel, further
complicating efforts to curb emissions. Given the
economic reliance on fuel in the fishing industry,
finding fuel-efficient methods for seafood production
has been a long-standing priority. Over the years,
various technical innovations have been introduced,
such as modernizing fishing fleets, investing in energy-
efficient propellers and gears, replacing engines, and
improving vessel hull designs [13].
The design of a ship’s hull plays a crucial role in
hydrodynamics, stability, safety, and operational
efficiency. However, additional variables, including
hull biofouling, can modify the topographical
characteristics of the hull surface, leading to increased
hydrodynamic drag and elevated fuel expenditure
[4,5].
To address these challenges, regulatory bodies like
the International Maritime Organization (IMO) have
implemented strict energy efficiency measures. The
2014 amendment to MARPOL ANNEX IV introduced
the Energy Efficiency Design Index (EEDI) for new
CFD Analysis of Cobalt-Based Ceramic Coatings
for Energy Optimization in the Fishing Naval Industry
D.S. Sanz
1
, S. García
1
, J. García
1
, D. Boullosa-Falces
2
& A. Trueba
1
1
University of Cantabria, Santander, Spain
2
University of the Basque Country, Portugalete, Spain
ABSTRACT: The rising cost of fossil fuels has created a significant economic challenge for the fishing fleet, whose
performance heavily relies on marine diesel consumption. The increase in operational costs due to fuel price
surges negatively impacts the profitability of ships, particularly in the fishing industry, where profit margins are
often tight. Given this issue, it is crucial to explore solutions that reduce fuel consumption without compromising
the operational efficiency of ships. In this context, cobalt-based ceramic coatings, designed and tested in
accordance with the ASTM-D3623 procedure, emerge as an innovative and promising alternative. These coatings
reduce biofouling adhesion, a buildup of marine organisms on the ship’s hull that increases frictional resistance
to movement, consequently leading to higher fuel consumption. By decreasing hydrodynamic resistance, ships
require less energy for propulsion, thereby optimizing fuel consumption. Additionally, these coatings provide
anticorrosive protection, extending the service life of ships and reducing maintenance costs. The cobalt-based
coating has been tested under controlled laboratory conditions and subjected to hydrodynamic shear forces
representative of ship navigation at a speed of 10 knots. This article evaluates via CFD the impact of these coatings
on the drag resistance of a trawler ship, demonstrating that the increase in hull roughness due to biofouling
adhesion on the cobalt-based ceramic coating after one month of navigation results in a 0.02% increase in drag.
In contrast, the Intersleek 1001 coating leads to a 6.35% increase in drag under the same conditions.
http://www.transnav.eu
the International Journal
on Marine Navigation
and Safety of Sea Transportation
Volume 19
Number 2
June 2025
DOI: 10.12716/1001.19.02.27
554
ships and the Ship Energy Efficiency Management Plan
(SEEMP) for existing vessels. Moreover, IMO has set
ambitious targets, aiming for a 40% reduction in CO2
emissions by 2030 and a 70% reduction by 2050
compared to 2008 levels. These regulations push the
shipping industry to optimize vessel efficiency and
minimize environmental impact [6].
Computational Fluid Dynamics (CFD) has become
an essential tool for improving ship hull design,
allowing researchers to simulate hydrodynamic
conditions and predict the impact of hull roughness
and biofouling on frictional resistance. Several studies
have demonstrated CFD’s effectiveness in this area. [7]
employed roughness functions in CFD simulations,
though they did not incorporate Reynolds-averaged
Navier-Stokes (RANS) calculations, which are crucial
for accurately modelling turbulent flows. Izaguirre-
Alza et al. used Star-CCM+ software to validate surface
roughness effects but relied on built-in roughness
functions [8].
Recent CFD advancements have aimed at
improving predictive accuracy. [9] introduced a CFD
model incorporating fouling control coatings,
validating results against experimental data. Later,
they applied a RANS model with a modified roughness
function to predict efficiency losses due to hull
deterioration. Other researchers, such as [10,11],
analysed biofouling effects on propeller efficiency,
while [12] developed a methodology using
OpenFOAM to estimate frictional resistance.
As the shipping industry seeks energy-efficient
solutions to comply with environmental regulations,
advanced CFD models will play a key role in
optimizing hull design, reducing fuel consumption,
and minimizing greenhouse gas emissions [1].
Biofouling refers to the unwanted adhesion and
accumulation of biotic deposits on surfaces submerged
in seawater [1315]. These deposits primarily consist of
organic biofilms formed by microorganisms embedded
in a self-produced polymeric matrix, which can also
trap inorganic particles such as salts or corrosion by
products [16]. Over time, microbial biofouling
facilitates the settlement of larger macro-organisms,
leading to macro-fouling. This phenomenon presents a
significant challenge to various sectors of the blue
economy, particularly maritime transportation, where
biofouling can increase hydrodynamic drag, reducing
ship performance and elevating fuel consumption by
up to 30% to maintain operational efficiency. The
subsequent rise in fuel demand has both economic and
environmental consequences, as fuel costs account for
nearly 50% of marine transportation expenses and
increased combustion results in higher greenhouse gas
emissions [17].
To mitigate the adverse effects of biofouling in
fishing vessels, marine antifouling coatings have been
developed to inhibit biofouling growth either by
releasing active biocidal agents or by providing a
smooth, non-stick surface that prevents the settlement
of fouling organisms. These coatings significantly
reduce frictional resistance, thereby improving vessel
efficiency, lowering fuel consumption, and decreasing
greenhouse gas emissions. Given the high dependency
of fishing fleets on fuel, enhancing hydrodynamic
efficiency is crucial for achieving more sustainable and
cost-effective operations. However, different fouling
control coatings vary in surface roughness and
durability, influencing their effectiveness in
minimizing frictional resistance over time. Hence,
accurately assessing the impact of different coatings on
vessel performance is essential for optimizing their
application in the fishing industry. Studies suggest that
full-scale ship resistance and power requirements for
various antifouling coatings can be predicted using
laboratory drag measurements and boundary layer
similarity laws. Moreover, empirical correlations have
been established to estimate frictional resistance in
different fouling conditions, providing valuable
insights for coating selection and maintenance
strategies [18].
Biofouling and corrosion, though distinct processes,
often coexist and exacerbate material degradation in
marine environments, particularly in fishing vessels
that operate under diverse oceanographic conditions.
Biofouling involves the colonization of submerged
surfaces by microorganisms such as bacteria and algae,
which form biofilms that serve as substrates for larger
fouling organisms like barnacles and mussels [19]. The
development of biofilms is influenced by multiple
factors, including water temperature, nutrient
availability, flow dynamics, and surface composition.
Smooth, defect-free surfaces generally exhibit lower
biofouling susceptibility compared to rough or
irregular surfaces. Furthermore, nutrient-rich waters
encountered in fishing grounds facilitate microbial
proliferation, enhancing biofilm growth and
subsequent macro-fouling, which significantly affects
vessel performance by increasing drag and fuel
consumption [20].
Corrosion, on the other hand, is the deterioration of
materials through chemical or electrochemical
reactions with their environment. In marine settings,
saltwater exposure accelerates corrosion, leading to
significant structural damage over time. Corrosion
manifests in various forms, such as metal oxidation,
steel degradation, and material fatigue, and is driven
by factors including oxygen exposure, moisture, acidic
conditions, and saline environments. The
consequences of marine corrosion for fishing vessels
include reduced hull integrity, increased maintenance
costs, and potential safety hazards. For instance,
corrosion-induced weakening of hull structures can
elevate the risk of leaks, compromising vessel safety
and operational efficiency. Therefore, corrosion
prevention strategies are crucial for ensuring the
longevity and reliability of fishing fleets. Common
mitigation measures include the use of corrosion-
resistant materials, protective coatings, and
comprehensive corrosion monitoring systems [21]. The
interaction between biofouling and corrosion is
intricate, as biofouling can accelerate corrosion
through microbial metabolic activity. Certain
microorganisms produce acidic compounds that
degrade material surfaces, while biofilms create
microenvironments that promote localized corrosion.
Additionally, biofouling can trap moisture and
corrosion-inducing agents, further exacerbating
material degradation. Consequently, integrated
antifouling and anticorrosion strategies are essential
for comprehensive fishing vessel maintenance. The
development of advanced coatings that
simultaneously provide antifouling and anticorrosion
protection is a promising approach to mitigating these
555
challenges and improving the sustainability of fishing
operations [13,22,23].
Among these advanced coatings, cobalt-based
enamels have demonstrated exceptional mechanical
properties, including high resistance to erosion and
corrosion [18]. Research has shown that these coatings
exhibit high hardness and durability, making them
highly effective in protecting fishing vessels against
biofouling and material degradation. The surface of
cobalt-based enamels is highly hydrophilic, preventing
the adhesion of organic moleculesa critical initial
step in biofilm formation. Additionally, the vitreous
structure of ceramic glazes provides a smooth surface
with an average roughness height of approximately
0.189 μm, reducing microbial adhesion and facilitating
the removal of biofouling through hydrodynamic
forces, thereby improving vessel efficiency and
reducing fuel consumption.
One significant advantage of cobalt-based enamels
is their environmental compatibility. Unlike traditional
antifouling coatings that may contain toxic heavy
metals such as copper, cobalt-based enamels do not
release harmful substances into marine ecosystems.
This environmentally friendly characteristic aligns
with sustainability goals, making them a viable
alternative for long-term applications in fishing fleets.
Furthermore, these coatings exhibit remarkable
durability, maintaining their protective properties
even in highly corrosive marine environments, which
is essential for vessels operating in harsh fishing
grounds [24]. Biofouling and corrosion are major
concerns for fishing vessels, impacting fuel efficiency,
structural integrity, and environmental sustainability.
The use of advanced antifouling and anticorrosion
coatings, such as cobalt-based enamels, represents a
promising solution for mitigating these challenges.
Continued research and innovation in coating
technology will play a pivotal role in enhancing the
performance and longevity of fishing vessels while
reducing operational costs and environmental impacts,
contributing to the decarbonization of the fishing
industry.
A Computational Fluid Dynamics (CFD)-based
Reynolds Averaged Navier-Stokes (RANS) model was
employed to analyse the hydrodynamic performance
of a fishing vessel by modifying experimental
variables. OpenFOAM (Open Field Operation and
Manipulation), an open-source, C++-based software,
was utilized due to its high accuracy in simulating
rough hull conditions and its capability to integrate
custom algorithms.
This study examined the impact of hull biofouling
on surface roughness under both static and dynamic
navigation conditions. Laboratory tests were
conducted on a cobalt-based ceramic coating and a
reference Intersleek 1001 coating to assess their
performance. The experimentally obtained roughness
data were then incorporated into the CFD model to
evaluate the reduction in frictional resistance achieved
through cobalt-based ceramic coatings. The findings
demonstrate the potential of advanced coatings to
enhance fuel efficiency and decrease greenhouse gas
emissions in fishing vessels, contributing to the
decarbonization of the fishing industry.
2 MATERIALS AND METHODS
2.1 CFD simulations
In this study, OpenFOAM is utilized as the primary
Computational Fluid Dynamics (CFD) simulation tool.
As an open-source C++ library, it adheres to an object-
oriented programming paradigm and is distributed
under a GPL license. OpenFOAM provides a wide
range of solvers for simulating laminar and turbulent
flows, single and multiphase systems, and adaptive
meshing. Its modular architecture allows for extensive
customization, making it particularly suitable for
analyzing multi-phase turbulent flows involving
floating bodies.
2.1.1 Solver numerical model
The hydrodynamic simulations in this study were
performed using the interFoam solver in OpenFOAM,
designed for incompressible, two-phase flow. The
solver employs the Navier-Stokes Equation (1) and
continuity equations Equation (2) to model laminar
flow of Newtonian fluids.
2
· p µ g
t

+ = − + +


(1)
(2)
Here, υ is the velocity, p is the pressure, ρ is density, μ
is the dynamic viscosity, g is the acceleration due to
gravity and 2 is the Laplace operator.
Specifically, interFoam solves the Reynolds-
Averaged Navier-Stokes (RANS) equations for two
immiscible phases, air and water, using finite volume
discretization and the Volume of Fluid (VOF) method
to capture the free surface. The VOF method treats air
and water as a single effective fluid, with the volume
fraction of water (α) in each computational cell
indicating the phase distribution: α = 1 for water, α = 0
for air, and 0 < α < 1 for mixed cells. The air-water
interface is approximated by the iso-surface α = 0.5. The
local density (ρ) and viscosity (μ) of the fluid, with
μwater = 1.00E-3 Pa·s and μair = 1.48E-5 Pa·s, are
determined for each computational cell using
Equations (3) and (4).
( )
1
water air
= +
(3)
( )
1
water air
= +
(4)
2.1.2 Fishing Vessels model, settings, and simulation
mesh
The hull model selected for this study corresponds
to a factory trawler with a length between
perpendicular of 48.17 meters, a beam of 12.00 meters,
and a depth of 10.16 meters. It has a draft of 4.49
meters, a wetted surface area of 780.82 m², and a
displacement volume of 1479.38 (Table 1). Las
formas del modelo de buque se muestran en la Figure
1. The body plan of the ship model is presented in
Figure 1.
556
Table 1. Full-scale trawler ship model specifications
Specification
Trawler ship
Length between perpendicular
Lpp (m)
48.17
Breadth
B (m)
12.00
Depth
D (m)
10.16
Draft
T (m)
4.49
Wetted surface area
S (m2)
780.82
Displacement volume
V (m3)
1479.38
Figure 1. Body plan of the trawler ship model
In this study, the nutkRoughWallFunction model
was implemented in OpenFOAM to account for
surface roughness effects in viscous resistance
predictions. Wall functions, derived from empirical
equations, were applied to define near-wall boundary
conditions for momentum and turbulence transport
equations. The wall function model in OpenFOAM
was based on the formulations proposed by [25] and
[26]. To ensure the precise application of wall functions
in mesh design, the y+ parameter, representing the
non-dimensional boundary layer thickness, had to
remain within a valid range. This constraint-imposed
limitations on the size of the mesh cells adjacent to the
hull surface, requiring their height to exceed the
roughness value for proper recognition in the
simulation.
The roughness model incorporated the sand-grain
roughness height on the hull surface to improve the
accuracy of viscous resistance predictions.
Additionally, the roughness parameter influenced
turbulence properties. The nutkRoughWallFunction
boundary condition provided a turbulent kinematic
viscosity definition to account for roughness effects,
relying on turbulence kinetic energy.
OpenFOAM supported three turbulence models: k-
ε, k-ω, and SST k-ω. In this study, turbulence was
modeled using the Reynolds-Averaged Stress (RAS)
SST k-ω two-equation model, with second-order
upwind discretization applied to ensure numerical
stability and accuracy.
To achieve accurate numerical simulation results,
the mesh previously utilized in the investigations of
[18] was adopted. The computational domain
underwent five successive refinements in the vicinity
of the hull to ensure adequate resolution, allowing for
the precise capture of pressure and viscous forces. The
snappyHexMesh utility was then employed to
integrate the hull geometry into the simulation
domain. This process included one additional
refinement near the hull, two refinements directly on
the hull surface, and the generation of boundary layers
to maintain the required non-dimensional wall
distance (y⁺). The final mesh configuration, consisting
of 6.95 million cells, resulted in a y⁺ value of
approximately 100. All simulations were conducted
assuming a static hull, with no mesh motion or
morphing applied.
2.1.3 Frictional resistance calculation using CFD
prediction
The CFD model was employed to numerically
predict viscous and pressure resistance, thereby
determining the total drag along the X-axis for a
navigation speed of 10 knots. To transform the
experimentally obtained average roughness heights
into equivalent sand-grain roughness heights, the
conversion equation proposed by [27] (Equations (5))
was applied. The conversion equation used was:
5.863·
sa
kk=
(5)
Here, ks is the equivalent sand grain roughness height,
and ka is the average roughness height.
By incorporating Equation (5) into the CFD
computational model, it was possible to avoid directly
modeling the rough surface caused by biofouling,
thereby reducing the computational effort required by
the designer. This approach proved particularly
advantageous when analyzing multiple roughness
levels, as was the case in the present study. Moreover,
accurately discretizing surface roughness would have
necessitated a significantly refined mesh, leading to a
considerable increase in computational cost and
simulation time. This methodology was feasible
because the correlation between the roughness
function and the roughness Reynolds number was well
established. Consequently, this correlation was
incorporated into the OpenFOAM solver through the
wall function and turbulence model, enabling an
efficient and accurate representation of roughness
effects.
CFD simulations were performed on a workstation
equipped with a 13th Gen Intel® CoreTM i9-13900KF
processor clocked at 3.00 GHz and 128 GB of installed
RAM. For each simulation, 6 cores were employed.
2.2 Experimental Phase
2.2.1 Preparation and composition of experimental panels
The preparation of experimental steel panels coated
with ceramic enamel followed a standardized
multistep procedure to ensure optimal coating
performance. The ceramic coating was applied to an
S235JR steel panel measuring 150 × 400 mm with a
thickness of 3 mm. Prior to the application of the
ceramic coating, the steel surfaces underwent abrasive
blast cleaning to achieve a surface roughness
corresponding to grades Sa 2.5 or Sa 3 in accordance
with ISO 8503. Surface cleanliness was ensured
following ISO 8501 standards, and contamination
levels were controlled so as not to exceed rating 2 as
defined in ISO 8502-3. Subsequently, the application of
a groundcoat layer was carried out
The enamel coating process consisted of sequential
application of a primer (100 g/m²), groundcoat (450
500 g/m²), and covercoat (700–800 g/m²). Thermal
treatment was carried out in a muffle furnace at 840°C
for the primer and groundcoat layers, and at 820°C for
the covercoat. Firing duration was adjusted according
to substrate thickness to achieve proper sintering
without compromising structural integrity. The
enamel slurry was prepared using 100% TSSG00505/1
enamel frit combined with 40% water by weight. The
557
mixture was homogenized for 10 minutes and sieved
through a 60-mesh screen to ensure particle size
uniformity. The final slurry exhibited a specific gravity
in the range of 1.751.77 g/ml, suitable for achieving
consistent coating thickness.
The qualitative chemical composition of the enamel
included more than 30% SiO₂ as the primary glass-
forming oxide. Additionally, 5–15% of Na₂O and B₂O₃
acted as fluxing agents, while minor constituents (<5%)
such as Li₂O, K₂O, CaO, MnO, CoO, CuO, Al₂O₃, TiO₂,
ZrO₂, and F provided specific functional and structural
properties to the final enamel layer.
2.2.2 Static and Dynamic Testing
The experimental test was conducted in accordance
with ASTM D4939-89 at the Biofouling Laboratory of
Cantabria University. This facilitys strategic location
allowed for the direct intake of seawater from the Bay
of Santander (43°28N, 3°48W) using a Grundfos
SP 14-4R stainless steel submersible pump, which
transported the water to the laboratory.
Throughout the experimental phase, a reference
panel coated with the commercial antifouling paint
Intersleek 1001 was evaluated in parallel with the
cobalt-based ceramic coating. The test panels were
fully immersed in seawater under static conditions
(static test) inside an experimental reactor to facilitate
biofouling development. These panels were mounted
on a drum using plastic supports.
After a one-month static exposure, the dynamic test
was initiated. This phase involved rotating the drum,
where the test panels were mounted, at a speed of 10
knots. This setup was designed to replicate the
hydrodynamic shear stress experienced on the hull of
a fishing vessel during navigation (Figure 2).
Figure 2. Experimental reactor: (a) static test, (b) Dynamic
test.
2.2.3 Surface Roughness Characterization of Ceramic
Coating
The surface roughness of the cobalt-based ceramic
coating was measured three times throughout the
experimental phase, following the guidelines
established in ASME/ANSI B46.1-2009. The first
measurement was taken prior to experimentation, the
second after one month of seawater immersion in the
reactor (static test), and the third following a one-
month dynamic test. The initial roughness
characterization of both the ceramic coating and the
Intersleek 1001 marine paint was carried out using a
contact profilometer (Taylor-Hobson, Surtronic 3+).
After the static and dynamic exposures, when the
surfaces were covered with biofouling, roughness
measurements were performed using the TQC surface
roughness tester (DC9000 Series), which is accurate to
±5 microns or <2%.
3 RESULTS AND DISCUSION
3.1 Experimental study
The results obtained during the experimental phase are
presented in Table 2. Before exposure in sea water, the
cobalt-based ceramic coating exhibited significantly
lower roughness values (Ka = 0.25 μm, Ks = 1.44 μm)
compared to Intersleek 1001 (Ka = 5.12 μm, Ks = 30.02
μm), representing a 1948.0% higher roughness for
Intersleek 1001.
After 30 days of static seawater exposure, both
coatings experienced an increase in roughness due to
biofouling accumulation. The cobalt-based coating
reached Ka = 10.21 μm and Ks = 59.86 μm, whereas
Intersleek 1001 showed lower roughness values (Ka =
7.83 μm, Ks = 45.91 μm), indicating 23.31% less
biofouling adhesion compared to the ceramic coating.
The roughness data obtained for the cobalt-based
ceramic coating exhibited highly similar values
compared to those reported by [4,5] in his studies on a
coating with similar characteristics, showing a
roughness variation of only 1.47%.
Under dynamic conditions, after 30 days of
exposure simulating navigation at 10 knots, the
roughness of the cobalt-based ceramic coating
significantly decreased (Ka = 0.31 μm, Ks = 1.82 μm),
approaching its initial values. In contrast, the Intersleek
1001 coating maintained a higher roughness after the
dynamic phase (Ka = 6.47 μm, Ks = 37.93 μm),
suggesting a lower self-cleaning efficiency compared
to the ceramic coating, un 1987.10% mas de rugosidad,
with a 1987.10% higher roughness.
Table 2. Ceramic coatings roughness under different
experimental conditions.
Condition
Coating
Ka
μm
Ks
μm
No exposure
Blue (Co)
0.25± 0.09
1.44±0.55
Intersleek 1001
5.12± 1.24
30.02±7.27
30-days exposure
Blue (Co)
10.21±1.79
59.86±7.92
(Static Conditions)
Intersleek 1001
7,83±1.18
45.91±7.05
30-days exposure
Blue (Co)
0.31±0.10
1.82±0.32
(Dynamic Conditions)
Intersleek 1001
6.47±0.93
37.93±6.86
558
3.2 CFD simulation at different fouling levels
Each simulation required 50637.4 seconds of physical
time to complete, producing a computational mesh of
7986487 cells, ensuring high-resolution domain
discretization.
The CFD simulation results for a trawler hull model
coated with the evaluated coatings under different
exposure conditions are summarized in Table 3. Before
seawater exposure, the hull coated with the cobalt-
based ceramic exhibited a viscous resistance (Rv) of
1.47 kN and a total resistance (Rtot) of 5.45 kN. In
contrast, the Intersleek 1001-coated hull showed
slightly higher resistance values, with Rv = 1.50 kN and
Rtot = 5.75 kN, representing a 5.58% increase in total
resistance. Figure 3, 4 and 5 presents the results of the
CFD simulation for the six different cases outlined in
Table 3.
Figure 3. Velocity distribution on the hull surface under no-
exposure condition: (a) Cobalt-based ceramic coating, (b)
Intersleek 1001 coating.
Figure 4. Velocity distribution on the hull surface under 30-
days exposure in static conditions: (a) Cobalt-based ceramic
coating, (b) Intersleek 1001 coating.
Figure 4. Velocity distribution on the hull surface under 30-
days exposure in dynamic conditions: (a) Cobalt-based
ceramic coating, (b) Intersleek 1001 coating.
After 30 days of static seawater exposure,
biofouling accumulation led to an increase in hull
resistance for both coatings. The trawler coated with
the cobalt-based ceramic reached Rv = 1.73 kN and Rtot
= 6.01 kN, reflecting a 10.40% rise in total resistance.
The hull coated with Intersleek 1001 also experienced
an increase, with Rv = 1.64 kN and Rtot = 5.89 kN,
corresponding to an 8.09% increase in total resistance.
Under dynamic conditions, after 30 days of
simulated navigation at 10 knots, the cobalt-based
ceramic coating effectively restored its initial resistance
values (Rv = 1.47 kN, Rtot = 5.45 kN), indicating
successful biofouling removal and a negligible
resistance variation of 0.02%. Conversely, the hull
coated with Intersleek 1001 maintained higher
resistance values (Rv = 1.51 kN, Rtot = 5.79 kN), leading
to a 6.35% increase in total resistance, suggesting lower
self-cleaning efficiency compared to the ceramic
coating.
Table 3. Simulation results for ceramic and commercial
coatings.
Condition
Coating
Viscous
resistance
Rv
Total
resistance
Rtot
Difference
%
No exposure
Blue (Co)
1.47 E+04
5.45 E+04
0,00
Intersleek 1001
1.50 E+04
5.75 E+04
5.58
30-days
exposure
Blue (Co)
1.73 E+04
6.01 E+04
10.40
(Static
Conditions)
Intersleek 1001
1.64 E+04
5.89 E+04
8.09
30-days
exposure
Blue (Co)
1.47 E+04
5.45 E+04
0.02
(Dynamic
Conditions)
Intersleek 1001
1.51 E+04
5.79 E+04
6.35
4 CONCLUSIONS
This study investigated the hydrodynamic
performance and biofouling behavior of a cobalt-based
ceramic coating compared to Intersleek 1001
antifouling paint through both experimental and
computational approaches. The results provide
valuable insights into the effectiveness of ceramic
coatings in reducing biofouling accumulation and
minimizing drag resistance in trawler hulls.
Experimental roughness measurements demonstrated
that, prior to seawater exposure, the cobalt-based
ceramic coating exhibited significantly lower
roughness values than Intersleek 1001. However, after
30 days of static seawater exposure, the ceramic coating
showed a greater increase in roughness due to
biofouling adhesion, with roughness values 23.31%
higher than those observed for Intersleek 1001. Despite
this, the roughness values obtained for the ceramic
coating closely aligned with previous studies,
confirming the reproducibility of the experimental
findings.
Under dynamic conditions, simulating navigation
at 10 knots, the ceramic coating displayed remarkable
self-cleaning properties. After 30 days of dynamic
testing, its roughness values returned to near-initial
conditions, whereas the Intersleek 1001-coated surface
retained a significantly higher roughness (1987.10%
greater than that of the ceramic coating). These
findings suggest that the cobalt-based ceramic coating
offers superior biofouling removal capabilities when
subjected to hydrodynamic shear stress.
CFD simulations further validated the experimental
findings by quantifying the impact of surface
roughness on the total drag resistance of a trawler hull.
Prior to exposure, the ceramic-coated hull exhibited
lower total resistance than the Intersleek 1001-coated
hull. After 30 days of static exposure, biofouling-
induced roughness increased the total resistance of
both coatings; however, dynamic conditions effectively
restored the ceramic coating’s initial resistance values,
559
whereas the Intersleek 1001-coated hull retained a
6.35% increase in resistance.
Overall, the results highlight the potential of cobalt-
based ceramic coatings as a viable alternative to
conventional antifouling paints, offering both lower
initial roughness and enhanced self-cleaning
performance under operational conditions.
ACKNOWLEDGEMENTS
The authors acknowledge Vibrantz Technologies for their
invaluable collaboration in the manufacturing of the
experimental panels. Their support has been instrumental in
making this research possible.
FUNDING
This work was supported by the Government of Cantabria
and the European Union Next GenerationEU/PRTR [project
Environmentally friendly bioactive coatings for energy
improvement and emissions reduction in the fishing
shipbuilding industry - BIO-ENER.].
REFERENCES
[1] H. Lindstad, R.M. Bright, A.H. Strømman, Economic
savings linked to future Arctic shipping trade are at odds
with climate change mitigation, Transp Policy (Oxf) 45
(2016) 2430.
https://doi.org/10.1016/j.tranpol.2015.09.002.
[2] S. García, A. Trueba, D. Boullosa-Falces, H. Islam, C.
Guedes Soares, Predicting ship frictional resistance due to
biofouling using Reynolds-averaged Navier-Stokes
simulations, Applied Ocean Research 101 (2020).
https://doi.org/10.1016/j.apor.2020.102203.
[3] L. Trueba-Castaneda, D.S. Sanz, S. Garcia, A. Trueba,
Analysis of biofouling economic impact on the Cantabria
fishing fleet, in: OCEANS 2021: San Diego Porto, IEEE,
2021: pp. 16.
https://doi.org/10.23919/OCEANS44145.2021.9706079.
[4] N. Hadžić, I. Gatin, T. Uroić, V. Ložar, Biofouling dynamic
and its impact on ship powering and dry-docking, Ocean
Engineering 245 (2022) 110522.
https://doi.org/10.1016/j.oceaneng.2022.110522.
[5] D.S. Sanz, S. García, L. Trueba-Castañeda, D. Boullosa-
Falces, A. Trueba, Antifouling and Anticorrosive
Protection of Renewable Energy Marine Structures with
TiO2-Based Enamel, TransNav, the International Journal
on Marine Navigation and Safety of Sea Transportation
18 (2024) 419424. https://doi.org/10.12716/1001.18.02.21.
[6] C.-C. Chang, M.-L. Huang, C.-H. Li, Analysis of emission
reduction strategies for the use of alternative fuels and
natural carbon sinks in international bulk shipping,
Energy Conversion and Management: X 24 (2024) 100702.
https://doi.org/10.1016/j.ecmx.2024.100702.
[7] M. Leer-Andersen, L. Larsson, An
experimental/numerical approach for evaluating skin
friction on full-scale ships with surface roughness, J Mar
Sci Technol 8 (2003) 2636. https://doi.org/10.1007/s10773-
003-0150-y.
[8] D.S. Sanz, S. García, A. Trueba, L. Trueba-Castañeda, H.
Islam, C. Guedes Soares, D. Boullosa-Falces, Numeric
analysis of the biofouling impact on the ship resistance
with ceramic coating on the hull, in: Trends in Maritime
Technology and Engineering Volume 1, CRC Press,
London, 2022: pp. 443449.
https://doi.org/10.1201/9781003320272-49.
[9] Y.K. Demirel, O. Turan, A. Incecik, Predicting the effect of
biofouling on ship resistance using CFD, Applied Ocean
Research 62 (2017) 100118.
https://doi.org/10.1016/j.apor.2016.12.003.
[10] S. Song, Y.K. Demirel, M. Atlar, An investigation into the
effect of biofouling on the ship hydrodynamic
characteristics using CFD, Ocean Engineering 175 (2019)
122137. https://doi.org/10.1016/j.oceaneng.2019.01.056.
[11] D. Owen, Y.K. Demirel, E. Oguz, T. Tezdogan, A. Incecik,
Investigating the effect of biofouling on propeller
characteristics using CFD, Ocean Engineering 159 (2018)
505516. https://doi.org/10.1016/j.oceaneng.2018.01.087.
[12] N. Speranza, B. Kidd, M.P. Schultz, I.M. Viola, Modelling
of hull roughness, Ocean Engineering 174 (2019) 3142.
https://doi.org/10.1016/j.oceaneng.2019.01.033.
[13] D.S. Sanz, S. García, L. Trueba, A. Trueba, Bioactive
Ceramic Coating Solution for Offshore Floating Wind
Farms, TransNav, the International Journal on Marine
Navigation and Safety of Sea Transportation 15 (2021).
https://doi.org/10.12716/1001.15.02.19.
[14] S. García, D. Boullosa-Falces, D.S. Sanz, A. Trueba, M.A.
Gomez-Solaetxe, Artificial-intelligence-model to
optimize biocide dosing in seawater-cooled industrial
process applications considering environmental,
technical, energetic, and economic aspects, Biofouling 40
(2024) 366376.
https://doi.org/10.1080/08927014.2024.2363241.
[15] E. Robertson, V. Choudhury, S. Bhushan, D.K. Walters,
Validation of OpenFOAM numerical methods and
turbulence models for incompressible bluff body flows,
Comput Fluids 123 (2015) 122145.
https://doi.org/10.1016/j.compfluid.2015.09.010.
[16] A. Trueba, L.M. Vega, S. García, F.M. Otero, E.
Madariaga, Mitigation of marine biofouling on tubes of
open rack vaporizers using electromagnetic fields, Water
Science and Technology 73 (2016) 12211229.
https://doi.org/10.2166/wst.2015.597.
[17] D. Boullosa-Falces, D.S. Sanz, S. Garcia, L. Trueba-
Castañeda, A. Trueba, Predicting tubular heat exchanger
efficiency reduction caused by marine biofilm adhesion
using CFD simulations, Biofouling (2022) 111.
https://doi.org/10.1080/08927014.2022.2110493.
[18] D.S. Sanz, S. García, A. Trueba, H. Islam, C.G. Soares,
Prediction of biological development effects on drag
forces of ceramic hull coating using Reynolds-averaged
NavierStokes-based solver, Biofouling (2023) 114.
https://doi.org/10.1080/08927014.2023.2209020.
[19] A. Trueba, S. García, F.M. Otero, L.M. Vega, E.
Madariaga, Influence of flow velocity on biofilm growth
in a tubular heat exchanger-condenser cooled by
seawater, Biofouling 31 (2015).
https://doi.org/10.1080/08927014.2015.1070404.
[20] D.S. Sanz, S. Garcia, A. Trueba, L.M. Vega, L. Trueba-
Castaneda, D. Boullosa-Falces, Application of ceramic
coatings to minimize the frictional drag penalty on ships,
in: OCEANS 2021: San Diego Porto, IEEE, 2021: pp. 15.
https://doi.org/10.23919/OCEANS44145.2021.9706045.
[21] D. Boullosa-Falces, S. García, D. Sanz, A. Trueba, M.A.
Gomez-Solaetxe, CUSUM chart method for continuous
monitoring of antifouling treatment of tubular heat
exchangers in open-loop cooling seawater systems,
Biofouling 36 (2020) 7385.
https://doi.org/10.1080/08927014.2020.1715954.
[22] D. Boullosa-Falces, M.A. Gomez-Solaetxe, Z. Sanchez
Varela, S. García, A. Trueba, D. Sanz, Study of the impact
of EMF on the reduction of biofouling in heat exchangers,
in: Trends in Maritime Technology and Engineering
Volume 1, CRC Press, London, 2022: pp. 511515.
https://doi.org/10.1201/9781003320272-57.
[23] S. Garcia, A. Trueba, L. Vega, E. Madariaga,
Improvement of Electromagnetic Fields treatment for
biofouling growth control in tubular heat exchanger-
condenser cooled by seawater, in: OCEANS 2017 -
Aberdeen, IEEE, 2017: pp. 15.
https://doi.org/10.1109/OCEANSE.2017.8084576.
560
[24] P. Yi, H. Jia, X. Yang, Y. Fan, S. Xu, J. Li, M. Lv, Y. Chang,
Anti-biofouling properties of TiO2 coating with coupled
effect of photocatalysis and microstructure, Colloids Surf
A Physicochem Eng Asp 656 (2023) 130357.
https://doi.org/10.1016/j.colsurfa.2022.130357.
[25] T. Cebeci, P. Bradshaw, Momentum Transfer in
Boundary Layers, Hemisphere Publishing Corporation,
New York, 1977.
[26] G. Kalitzin, G. Medic, G. Iaccarino, P. Durbin, Near-wall
behavior of RANS turbulence models and implications
for wall functions, J Comput Phys 204 (2005) 265291.
https://doi.org/10.1016/j.jcp.2004.10.018.
[27] T. Adams, C. Grant, H. Watson, A Simple Algorithm to
Relate Measured Surface Roughness to Equivalent Sand-
grain Roughness, International Journal of Mechanical
Engineering and Mechatronics (2012).
https://doi.org/10.11159/ijmem.2012.008.