53
1 INTRODUCTION
Global seaborne trade is doing well, but is facing a
number of political (inward-looking policies and
protectionism), economical (merger and alliances of
large shipping lines), and environmental challenges
(carbon footprint and sulphur cap) making its long-
range outlook quite uncertain
1
. This calls the maritime
administrations and industries, among others, for
maximizing efficiency of the shipping routes.
Shipping routes are specific tracks that vessels follow
when traveling between ports and form a global
maritime exchange network. Their spatial
characteristics are not static though, as over time
1
https://unctad.org/en/PublicationsLibrary/rmt2018_
en.pdf
shipping responds to a number of external
parameters. The actual tracks may for instance result
from either security reasons (e.g. avoiding risk of
exposure to piracy), or economic reasons leading to
an optimization process for avoiding rough marine
weather (Arguedas et al., 2018). To what extent tracks
are already optimized in actual business is not
straightforward to assess, but at the same time it
would represent an extremely valuable piece of
information.
We contribute to this topic via an inter-comparison
exercise between reported vessel tracks as collected
through the Automatic Identification System (AIS)
and tracks planned through the ship route
optimization model VISIR. This paper is part of a
broader strategy for a comprehensive verification and
evaluation of the VISIR ship routing model. The
verification (i.e., checking if its equations are solved
Preliminary Inter-comparison of AIS Data and Optimal
Ship Tracks
G. Mannarini & L. Carelli
CMCC, Lecce, Italy
D. Zissis, G. Spiliopoulos & K. Chatzikokolakis
M
arine Traffic, London, United Kingdom
ABSTRACT: Optimal ship tracks computed via the VISIR model are compared to tracks recorded by the
Automatic Identification System (AIS). The evaluation regards 43 tracks in the Southern Atlantic Ocean, sailed
during 2016-2017 by different bulk carriers. In this exercise, VISIR is fed by wave analysis fields from the
Copernicus Marine Environment Monitoring Service (CMEMS). In order to reproduce vessel speed loss in
waves, a new methodology is developed, where kinematic information from AIS is fusioned with wave
information from CMEMS. Resulting VISIR tracks are analyzed along with AIS tracks in terms of their
topological features and duration. The tracks exhibit quite diverse topological shapes, including orthodromic,
loxodromic, and other paths with complex and dynamic diversions. The distribution of AIS to VISIR track
durations is analyzed in terms of several parameters, such as the AIS to VISIR track length and their Fréchet
distance. Model features of VISIR affecting the results are discussed and future developments suggested by the
results are outlined.
http://www.transnav.eu
the International Journal
on Marine Navigation
and Safety of Sea Transportation
Volume 13
Number 1
March 2019
DOI: 10.12716/1001.13.01.04
54
correctly) included comparison to both static
(Mannarini et al., 2016) and time-dependent
(Mannarini & Carelli, 2019) analytical solutions as
well as comparison to the outcomes of another path
planning model (Mannarini et al., 2018). The first step
in evaluation (i.e., assessing whether the model fairly
represents reality) is presented in this work.
The core of the paper is organized into a
description of the evaluation methodology (Section 2)
and a presentation of the results (Section 3), which
precede conclusive remarks and outline (Section 4).
2 METHODOLOGY
In this Section we describe the methodology
employed for this evaluation experiment. The
description comprises the extraction of AIS tracks
from reported raw data (Section 2.1) and the
computation of optimal tracks through VISIR (Section
2.2). The key development enabling the comparison of
AIS and VISIR tracks is the computation of the vessel
speed loss in waves out of AIS kinematic information,
VISIR vessel performance model, and sea state
analysis fields (Section 2.3).
Abbreviations and symbols employed throughout
this text are defined in Table 1.
Table 1. Some abbreviations and symbols employed in this
manuscript.
_______________________________________________
Symbol Meaning Units
_______________________________________________
SOG Speed Over Ground kts
STW Speed Through Water kts
COG Course Over Ground deg
HDG Heading deg
EOT Engine Order Telegraph %
UKC Under Keel Clearance m
HWHM Half Width Half Maximum [-]
L
wl Length at the waterline m
B
wl Beam at the waterline m
T
avg Average draught m
P
max (fitted) maximum engine rating kW
V
max (fitted) maximum speed in calm water kts
H
s Significant wave height m
α Wave direction relative to vessel heading deg
_______________________________________________
2.1 AIS tracks
Nowadays, a multitude of tracking devices and
systems produce massive amounts of maritime data
on a daily basis. The most commonly used of such
tracking systems is the Automatic Identification
System (AIS), a collaborative, self-reporting system
that allows vessels to broadcast their identification
information, characteristics and destination, along
with other information originating from on-board
devices and sensors, such as location, speed and
heading (M.1371, 2014). AIS messages are broadcast
periodically and they are received by other vessels
equipped with AIS transceivers, as well as by on-
ground stations and satellites.
A growing body of literature describes methods of
exploiting AIS data for safety and optimisation of
seafaring, traffic analysis, anomaly detection, route
extraction and prediction, collision detection, path
planning, weather routing and many more (Tu et al.,
2018). As the amount of available AIS data grows to
massive scales though, researchers are realising that
computational techniques must contend with
difficulties faced when acquiring, storing, and
processing the data. Applying traditional techniques
to data processing can lead to processing times of
several days, if applied to global data sets of
considerable size. Additionally, algorithms are
challenged by difficulties related to the datasets
themselves; including highly skewed, not uniform,
and uncertain data. For example, the update interval
for AIS is not constant, but dependent on a ship’s
behaviour; as such it is common for a vessel to
broadcast its data every three minutes if moving no
faster than 3 knots, while every two seconds if
travelling above 14 knots and changing course. In
such occasions, the collected positions are spatially
and temporally closer. In addition to this, depending
on AIS coastal coverage, there are geographical areas
where huge amounts of data are collected (e.g. busy
ports) while others with much less data. Furthermore,
human errors during data entry generate
discrepancies for example in naming of ports and
areas, generating further uncertainty and ambiguity.
In our previous work (Spiliopoulos et al., 2017), we
proposed an efficient big data approach for building a
global network of sea routes from AIS data. As a first
step of this process we reassign departure and arrival
information for each vessel trajectory as although this
data is existent in AIS it is error prone and often an
issue of confusion. In the present work we consider
vessel trajectories travelling between the ports of
Buenos Aires (Argentina) and Port Elizabeth (South
Africa) from July 2016 until the end of December
2017. These trajectories amounted to more than
160,000 AIS messages in total. Each of these messages
consists of the coordinates of the vessel transmitting
the message, the corresponding kinematic
characteristics of the vessel (such as SOG, COG, and
HDG), and a timestamp. For identification purposes
an anonymised vessel’s identifier and voyage data are
bound to each message.
Within this dataset, we identified 61 different
trajectories in total, each of them being performed by
a different vessel. Moreover 51 of these trajectories
correspond to dry bulk carriers and the remaining 10
trajectories to wet bulk carriers. For each voyage, the
vessel’s dimensions, engine power and maximum
historical speed information were included in our
analysis. The summarised characteristics of the
anonymised vessels in this dataset are shown in Table
2.
Table 2. Vessel characteristics. and stand for average
and standard deviation of the vessel sample.
_______________________________________________
Dry bulk Wet bulk
_______________________________________________
Number of vessels 51 10
Length [m] 201 182
19 2
Width [m] 32 31
2 2
Engine Power [kW] 8600 9000
1500 2200
Max Speed [knots] 12 12
2 3
_______________________________________________
55
2.2 VISIR tracks
VISIR (an acronym for "discoVerIng Safe and effIcient
Routes") is the ship routing model resulting from the
prototype first published in Mannarini et al. (2013).
The model eventually evolved to compute least-time
tracks in presence of time-dependent fields from
wave models (Mannarini et al., 2016), and has been
recently extended to deal also with ocean currents in
Mannarini et al. (2019).
VISIR is based on a graph-search algorithm, with
graph edges accounting for vessel COG, and edge
weights depending on the sailing time between graph
nodes. Graph edges crossing the landmass are
pruned, enabling computation of tracks even in
coastal waters or in vicinity of islands. Furthermore,
vessel intact stability can be accounted for through
checks on parametric roll, pure loss of stability, or
surfriding/broaching-to (IMO, 2007). Either
intentional speed reduction (EOT<1) or course change
can be exploited by VISIR for fulfilling the stability
constraints.
2.2.1 Path planner setup
For this actual evaluation experiment, the VISIR
model configuration is described through the
parameters provided in Table 3. The graph resolution
parameters are chosen to compromise between spatial
and angular accuracy on the one hand and
computational effort on the other one. UKC is not
checked in this exercise, as the employed GEBCO
bathymetry
2
would not allow, for the actual vessel
draught, a UKC>0 at the Western end of the AIS
tracks (located at the estuary of Rio de la Plata, in
South America).
Safety constraints and intentional speed reduction
are also disabled as, according to Mannarini et al.
(2019), for this actual route, they do not to
significantly impact the results. Conversely, disabling
them allows reducing the computer RAM allocation
of the computations and, thus, increasing the
maximum number of time steps considered for VISIR
paths. In fact, the vessel type considered in this
experiment is (both dry and wet) bulk carrier, which
top speeds (Table 2) are generally lower than
container ships considered instead in Mannarini et al.
(2019). A lower speed implies a longer sailing time,
which is represented in VISIR through a larger
number of time steps and, thus, requires higher RAM
allocation.
Table 3. VISIR configuration for the computations of this
experiment.
_______________________________________________
Feature Value
_______________________________________________
Grid resolution 1/7 deg ( = 8.6 nmi in meridional
direction)
Angular resolution 8.1 deg
Ocean currents neglected
Safety constraints shoreline only, UKC>0 and intact
stability checks disabled
Intentional speed disabled
reduction
_______________________________________________
2
https://www.gebco.net/data_and_products/gridded_b
athymetry_data/)
For the same reason, ocean currents are neglected
in this first version of the evaluation experiment. The
role of this approximation with respect to the self-
consistency of the whole methodology is discussed in
Section 2.3.1.
Finally, CMEMS
3
three-hourly wave fields are
averaged into daily fields before being employed by
VISIR. Again, as discussed in (Mannarini et al., 2019),
the reason for this approximation is the reduction of
RAM allocation.
2.3 Vessel response function
As mentioned above, the critical modeling piece for
the evaluation of VISIR vs. AIS tracks is the vessel
response function. It defines the involuntary speed
loss in waves due to the added resistance R
aw
(Bertram and Couser, 2014). Because of speed loss,
path diversions may allow sailing at an higher speed
than along the least-distance track. Thus, by taking a
diversion, destination may be reached earlier. Since
the objective of the track optimization is to minimize
such a sailing time, diversions that optimally
compromise between reduced speed loss and
increased track length are chosen by the algorithm.
The vessel response function is defined within
VISIR as the STW sustained at specific values of
significant wave height H
s (Mannarini et al., 2016). Its
functional form is obtained from a power balance at
the ship propeller and makes use of a parametrization
of R
aw based on a statistical reanalysis of numerical
simulations via the Gerritsma and Beukelman's
method (Alexandersson, 2009). The parametrization
depends on just three main geometrical parameters:
vessel length L
wl, beam Bwl, and draught Tavg. When
applied to the VISIR power balance, it results into a
Gaussian-shaped vessel response function, with a
peak value given by vessel top speed in calm water
V
max and a HWHM proportional to maximum engine
break power P
max.
The dependence on wave direction (relative to
vessel heading) is presently neglected within the
VISIR vessel model. The impact of this approximation
is discussed in the subsection below.
2.3.1 Developments for the current experiment
For the current evaluation experiment, the five
parameters (L
wl, Bwl, Tavg, Pmax, Vmax) needed for
computing the vessel response function are identified
in the following way:
1 Hull geometry parameters (L
wl, Bwl, Tavg)
correspond to the vessel details identified through
the IMO-number provided along with the AIS
record;
2 Propulsion and performance parameters (P
max,
V
max) are fitted to the data of speed loss in waves,
generated as in the following.
A speed loss diagram (scatter plots of SOG vs.
corresponding H
s experienced by the vessel) is first
produced.
3
http://marine.copernicus.eu/
56
H
s is not part of the AIS record. However, it can be
estimated from sea state model analysis. Analyses are
the best available reconstructions of an environmental
state, making use of both observations and
geophysical model outputs. We employ CMEMS
analysis fields which assimilate observations of
significant wave height from Jason 2 & 3, Saral and
Cryosat-2 altimeters into the MFWAM sea state
model
4
. For each AIS track leg, we extract the spatially
and temporally nearest CMEMS gridpoint value. This
value represents our best estimation of H
s
encountered by the vessel at that specific AIS spatial
location and time.
Resulting data of speed loss in waves are
displayed in Figure 1, where they are plotted vs.
either H
s or relative wave direction α. Despite the fact
that, even for a specific vessel, data are strongly
scattered, a general trend for speed loss with
increasing H
s is recognized. Also, maximum loss is
achieved for head waves. This is consistent with
calculations and towing tank data of wave added
resistance provided by Tsujimoto et al. (2013).
SOG
a)
b)
Figure 1. Speed loss in waves out of AIS kinematic data and
CMEMS sea state fields. Panel a) displays the dependence
on significant wave height, while b) the dependence on
wave-vessel relative direction (α = 0 deg means head seas, α
> 0 refer to waves from the starboard). Marker grey tones
refer to individual voyages/vessels.
Datapoint scattering in Figure 1 can be attributed
to:
4
http://cmems-
resources.cls.fr/documents/PUM/CMEMS-GLO-PUM-001-
027.pdf
1 neglect of ocean currents, which combines with
STW for producing SOG;
2 limited skill of the CMEMS analysis fields in the
reconstructing the sea state.
The difference between SOG and STW, being
related to ocean current magnitude and direction
(Mannarini & Carelli, 2019), can be as high as several
knots. However, identifying STW with SOG is
consistent with the fact that, for this experiment, we
are neglecting ocean currents also for the computation
of VISIR optimal tracks. A more accurate treatment is
planned for future developments of this evaluation
methodology.
009
002
051
016
a)
b)
c)
d)
SOG
Figure 2. SOG dependence on significant wave height. AIS
data are represented as gray dots, while the VISIR response
function for fitted P
max and Vmax parameters is shown as a
dashed line. Each panel refers to a different voyage/vessel,
identified by the 3-digit code in the top-right.
Both the response function resulting from AIS
kinematical data and CMEMS wave fields and the
fitted VISIR speed loss curves are displayed in Figure
2 for selected vessels. The data have been previously
57
filtered by pruning data in the vicinity of the track
endpoints (harbours). There, due to vessel
acceleration, SOG can strongly vary even in
correspondence of a constant H
s value.
VISIR speed loss curve generally fits well to the
data, but in the case when they include more than a
branch at larger H
s. If this is the case, the VISIR curve
is fitted in between the branches (Figure 2b).
The presence of multiple speed loss branches may
be due to vessel performance in head waves, which
are not accounted by the present ship model. In some
cases, even a speed increases with H
s is observed
(Figure 2d), which may be due either to EOT changes
or to ocean currents.
In Fujii et al. (2017) too, a speed loss curve out of
AIS and model wave data is displayed. The authors
considered tracks of container ships and pure car
carriers in the North Pacific. However, their data do
not support a clear relation between SOG and H
s.
Finally, we note that the information in step b) of
the present procedure just represents the P
max and
V
max fit parameters and does not necessarily agree
with the actual vessel parameters for maximum
engine brake power and speed.
3 RESULTS
In this Section, results relative to systematic
application of the methodology of Section 2 to
transatlantic passages between Buenos Aires
(Argentina) and Port Elizabeth (South Africa) and
vice-versa during the years 2016 and 2017 are
presented and discussed. The results for a few
individual tracks (Section 3.1) precede the analysis of
the ensemble of the tracks from the topological
viewpoint (Section 3.2) and by investigation of the
dependence of track duration on several variables
(Section 3.3).
3.1 Individual tracks
A one-to-one comparison of VISIR to AIS track
topology is displayed for a selection of all the voyages
in Figure 3.
The tracks are portrayed on top of the CMEMS
wave fields. The latter are generally taken at the
timestamps of the AIS waypoints and displayed as
adjacent vertical stripes. VISIR instead employs daily
averages of the CMEMS waves. These are shown in
Figure 3.b-c in the portion of the map containing the
VISIR trajectory.
Finally, each panel displays also the geodetic track
computed by VISIR. This track is, in the open ocean,
an arc of great circle joining the track endpoints. In
that case, it is identical to the orthodromic path. The
main features of the various panels of Figure 3 are
described in the following:
1 For this ship voyage, both VISIR and AIS tracks
take an initial Northbound diversion and then sail
to destination along a trajectory close to the
loxodromic (i.e., constant bearing) path. The
diversion is slightly anticipated by VISIR, which in
the second part of the voyage computes a
diversion approaching the geodetic track.
2 A significant track disagreement is noticed, with
AIS close to orthodromic navigation and VISIR
taking a wide Northbound diversion. VISIR
diversion is instrumental in avoiding the rough
seas between 25-10
o
W and 5-15
o
E. This
corresponds to the second part of the voyage, at 6
and 12 days since departure respectively. AIS track
may result from the fact that the rough sea in the
second part of the voyage could not be exactly
forecast at the time of departure.
3 AIS track diverts North while VISIR track diverts
even South of the geodetic. The different wave
field stripes in the upper and lower part of the
map make clear that the diversion computed by
VISIR is instrumental in avoiding rough seas at the
latitude of the geodetic and above.
4 Both AIS and VISIR tracks sail a path between the
rhumb-line and the least-distance track. VISIR
track is closer to the latter in vicinity of both
departure and arrival locations, resulting in a
shorter track length than AIS data.
In both Figure 3.b-c an appreciable difference
between three-hourly and daily-averaged wave fields
can be noticed. The observed divergent AIS and VISIR
trajectories might be ascribed to this fact.
3.2 Tracks ensemble topology
The original 61 AIS tracks were reduced to 43 due to
the fact that some tracks presented an anomalous
length or that the response function, upon spatial
pruning, contained too few datapoints. Each of the 43
voyages in the AIS record is sailed by a different
vessel, for which a speed loss curve was computed as
described in Section 2.3.1.
It is employed for computing an optimal track via
VISIR between the actual AIS endpoints, for their
specific departure date and time.
In Figure 4 all Eastbound tracks from both AIS and
VISIR between July 31 2016 and December 13, 2017
are displayed. The tracks form a bundle with an
appreciable meridional extent. In the middle of the
passage, the bundle extent is about 12
o
for AIS and
about 20
o
for VISIR, which bundle extends even South
of the geodetic track. For both AIS and VISIR, there is
a general trend to larger diversions for tracks sailed
during (Southern hemisphere) winter months. A
significant meridional dispersion of the AIS tracks is
also noticed in the data published by Fujii et al.
(2017).
58
H
s
[m]
c)
051
VISIR
AIS
b)
002
VISIR
AIS
d)
VISIR
AIS
a)
009
VISIR
AIS
016
Figure 3. Tracks of voyages referenced also in Figure 2 displayed on top of CMEMS three-hourly wave fields. The white
arrows denote wave direction. In each panel, AIS, VISIR, and geodetic tracks are displayed as a green solid line, a red
dashed line, and a blue dotted line respectively. Panels b) and c) are split vertically, with the daily-averaged wave fields
displayed in correspondence of the VISIR tracks and the three-hourly fields in correspondence of the AIS tracks.
3.3 Tracks ensemble – key metrics
The track ensemble is also analyzed in terms of track
duration (sailing time), which is the optimization
objective for VISIR. Track duration could be the
guiding principle for shipmaster decisions as well,
since costs relative to bunker and onboard personnel
are proportional to duration.
First of all, durations of VISIR optimal tracks (T
V)
are compared to durations on the geodetic track (T
G).
They are computed by VISIR accounting for speed
loss in waves. Figure 5a confirms that VISIR’s
optimization works as expected, saving up to about
two days with respect to orthodromic navigation.
T
V are also compared to AIS durations (TA) in
Figure 5b, finding T
V < TA always. Again, VISIR
savings exceed two days in some cases.
In order to get a deeper insight on the reasons for
the better performance of VISIR, the relative duration
saving -
va of VISIR to AIS trajectory (va = TV/TA -1)
is displayed in Figure 6. in dependence of four
different variables:
59
VISIR
AIS
a)
b)
009
016
051
002
009
016
051
002
Figure 4. Route departing from Buenos Aires (Argentina) arriving in Port Elizabeth (South Africa). a) displays AIS tracks,
while b) shows VISIR simulations departing on the same dates. Dashed bold tracks with numeric labels refer to the selected
voyages referenced also in Figure 2.
Figure 5. Analysis of VISIR optimal track durations TV. Panels a) and b) compare to the duration of VISIR geodetic TG and
AIS tracks T
A respectively. In every panel, the codes of the voyages are put in evidence, which speed losses in waves are
displayed in Figure 2.
Vessel maximum speed (fit parameter Vmax). There
possibly is some correlation between
va and Vmax.
This is one of the fit parameters in Figure 2 and is
related but not identical to the actual vessel top
speed. This trend should be confirmed by a larger
track statistics;
The length ratio of AIS to VISIR tracks (L
A/LV). It
is close to unity with a high precision (within 5%) in
most cases. When departing from unity, a trend of -
va increasing with LA/LV seems to be supported by
the data;
1 The month of departure of the track. No clear
seasonal trend is recognized, but a larger -
va
variability in (Southern Hemisphere) summer
months. Before making hypothesis on its origin,
the statistics should be enlarged;
60
Figure 6. Analysis of relative VISIR to AIS track duration savings -va =1-TV/TA with respect to several parameters. Panel a)
displays the dependence on the V
max fit parameter; b) dependence on the AIS to VISIR track length ratio LA/LV; c)
dependence on month of departure; d) dependence on Fréchet distance F
d between VISIR and AIS tracks. In every panel, the
codes of the voyages are put in evidence, which speed losses in waves are displayed Figure 2.
2 The Fréchet distance
5
Fd, representing VISIR and
AIS track similarity. It seems that the maximum -
va increases with Fd, but the data are quite
scattered.
4 CONCLUSIONS
We have completed a preliminary evaluation of
VISIR ship routing model with respect to vessel
tracks reported through AIS data. The evaluation has
included the development of a new methodology for
determining speed loss in waves from fusion of
empirical data (AIS) and meteo-oceanographic model
output (wave analysis fields). A case study of 43
voyages in the Southern Atlantic Ocean has been
considered. The results indicate that:
1 VISIR track topology generally agree well with
AIS one, though in some cases completely
different tracks, with a large Fréchet distance from
AIS, are computed;
2 There is a significant annual variability of the
spatial structure of the tracks, both in the AIS
records and in the VISIR simulations. A number of
5
http://www.kr.tuwien.ac.at/staff/eiter/et-
archive/cdtr9464.pdf
AIS tracks are close to the loxodromic track and
the VISIR tracks at times extend even South of the
orthodromic track;
3 VISIR sailing times are always shorter (4-14%)
than AIS ones, while their total length remains
quite close to the AIS one in most cases;
4 VISIR time savings are larger for AIS tracks longer
than VISIR ones. Furthermore, the duration
savings could also be related to top vessel speed
and seems to vary most during (Southern
Hemisphere) summer months.
These conclusions are supportive of the fact that
some kind of optimization (either automated or
manual) took place in most of the actually sailed
tracks.
Discrepancies between VISIR and AIS both in
terms of duration and spatial structure can be
ascribed to at least two factors:
1 VISIR model features and approximations. For
this first exercise:
three-hourly wave fields have been averaged
into daily fields, for the sake of reducing the
computer RAM allocation, cf. Mannarini &
Carelli (2019). As noted in Section 3.1, this may
be one of the reasons for the larger
discrepancies between AIS and VISIR tracks;
61
ocean currents have not been considered,
neither for the optimal track computation, nor
for reconstructing STW out of SOG;
a directional wave response is not yet modeled.
Its impact depends on how the a-directional
VISIR response function was fitted to the AIS-
CMEMS data in Figure 2. Whenever VISIR
curve is adjusted to the steepest branch of the
data, its speed loss is overestimated to always
be the one relative to head waves. On the other
hand, adjusting to the most constant branch
would underestimate the speed loss in head
wave. Overestimation of speed loss would
induce an overestimation of the spatial
diversion along the optimal tracks;
2 Unavailability, for the ship officer in charge of
track planning, of long enough wave forecasts.
Since forecasts are usually limited to a few days
lead time (e.g. 5 days for CMEMS
6
, 7.5 days for
NCEP
7
), a sub-optimal track may result from this
fundamental knowledge gap at the time of vessel
departure. This is especially relevant when rough
seas are encountered several days after departure
(cf Figure 3.b,c). This may induce a conservative
approach by the shipmaster, for instance avoiding
diversions towards circumpolar latitudes (cf
Figure 4). VISIR tracks instead, employing
reconstructions of the sea-state (analysis fields),
are not affected by this fundamental limitation.
We consider these preliminary results quite
encouraging. A larger track statistics and
consideration of routes in other parts of the global
ocean should enable an even deeper insight into
optimization choices by actual vessels. Also, this
approach provides further indication of what VISIR
model features need to be developed more urgently.
ACKNOWLEDGEMENTS
This work has received funding from the European Union
Horizon 2020 research and innovation programme under
grant agreements No. 633211 (AtlantOS) and No. 732310
(BigDataOcean) and the Italy-Croatia Interreg V-A
programme under project ID 10043587 (GUTTA).
REFERENCES
Alexandersson, M. A study of methods to predict added
resistance in waves. Master’s thesis, KTH Centre for
Naval Architecture, 2009.
Arguedas et al. 2018. Maritime Traffic Networks: From
historical positioning data to unsupervised maritime
traffic monitoring. IEEE-ITS, 19(3): pp.722732.
Bertram V. and Couser. P. 2017. Computational methods
for seakeeping and added resistance in waves. In B.
Volker (ed.), 13th International Conference on Computer
and IT Applications in the Maritime Industries, Redworth,
12-14 May 2014, pages 8–16. Harburg: Technische
Universität Hamburg.
6
GLOBAL_ANALYSIS_FORECAST_WAV_001_027
7
https://polar.ncep.noaa.gov/waves/product_table.s
html?-hs-multi_1-atlantic-latest-
Fujii M., Hashimoto H., Taniguchi Y. 2017: Analysis of
Satellite AIS Data to Derive Weather Judging Criteria
for Voyage Route Selection. TransNav, the International
Journal on Marine Navigation and Safety of Sea
Transportation, Vol. 11, No. 2, doi:10.12716/1001.11.02.09,
pp. 271-277
IMO. MSC.1/Circ.1228 Revised guidance to the Master for
avoiding dangerous situations in adverse weather and
sea conditions. Technical report, International Maritime
Organization, London, UK, 2007.
M.1371: Technical characteristics for an automatic
identification system using time-division multiple
access in the VHF maritime mobile band.” [Online].
Available: https://www.itu.int/rec/R-REC- M.1371/en
Mannarini G., Coppini G., Oddo P. and Pinardi N. 2013. A
prototype of ship routing decision support system for
an operational oceanographic service. TransNav, the
International Journal on Marine Navigation and Safety of
Sea Transportation, 7(1): 53–59.
Mannarini G., Pinardi N., Coppini G., Oddo P. and Iafrati
A. 2016. VISIR-I: small vessels – least-time nautical
routes using wave forecasts. Geoscientific Model
Development, 9(4): 1597–1625. https://www.geosci-model
dev.net/9/1597/2016/gmd-9-1597-2016.pdf
Mannarini G. & Carelli L. 2019. VISIR-I.b: waves and ocean
currents for energy efficient navigation. Geoscientific
Model Development Discussions, (under review)
Spiliopoulos, G., Chatzikokolakis, K., Zissis, D., Biliri, E.,
Papaspyros, D., Tsapelas, G. and Mouzakitis, S. 2017.
Knowledge extraction from maritime spatiotemporal
data: An evaluation of clustering algorithms on Big
Data. In Big Data (Big Data), 2017 IEEE International
Conference on (pp. 1682-1687). IEEE.
Tu E., Zhang G., Rachmawati L., Rajabally E., and Huang
G. B. 2018, Exploiting AIS Data for Intelligent Maritime
Navigation: A Comprehensive Survey From Data to
Methodology. IEEE Trans. Intell. Transp. Syst., vol. 19,
no. 5, pp. 1559–1582.
Tsujimoto, M., Kuroda, M. and Sogihara, N. 2013.
Development of a calculation method for fuel
consumption of ships in actual seas with performance
evaluation. In ASME 2013 32nd International
Conference on Ocean, Offshore and Arctic Engineering
(pp. V009T12A047-V009T12A047). American Society of
Mechanical Engineers