33
1 INTRODUCTION
1.1 Motivation
Oceans are resourceful. Ocean operations are vastly
increasing the last decades, and so is the interest in
unmanned surface vehicles (USV) and autonomous
ships. The ocean’s extreme weather and far distances
can result in high-risk-high-cost work conditions.
The world’s economy is mainly defined by three
areas: energy, transportation and communication
(Rifkin, 2012). Ocean industries push the boundaries
of these three areas to the limits. Unmanning of
maritime assets by excessive automation and remote
control could reduce or eliminate the risk imposed on
crew; however, infrastructure cost will increase. A
huge safety potential is accompanied with more
benefits; by removing crew from the assets, crew
related costs are, in theory, removed, costs such as
cooling, heating and ventilation. Accommodation
spaces are, in theory, no longer required, less power
consumption is projected and the chain of promises
goes on.
The drivers for unmanning maritime assets are
developing into motivations for building and
operating autonomous vessels. One example is, Yara
Birkeland, a 120 TEU open-top zero emissions
autonomous containership, planned launch is
expected before 2020, the ship is under construction
with Kongsberg technology. Another example is the
car ferry Falco that was built using Rolls-Royce
technology, launched late 2018.
As operators are moved from the far end of the
operation to shore control centers, their experiences
are changed, their feelings and senses while on duty
from one hand, their toolboxes and control authority
on the distant ship from the other hand. The current
remote-control technologies and their limitations are
subject to discussion. In this paper, the challenges of
Is Full-autonomy the Way to Go Towards Maximizing
the Ocean Potentials?
R. Zghyer, R. Ostnes & K.H. Halse
Norwegian University of Science and Technology, Trondheim, Norway
ABSTRACT: Growth prospects for ocean economy are promising because ocean industries are addressing
challenges such as food security, energy security and climate change. However, safety and efficiency are the
general challenges of ocean operations. Increased automation is believed to solve these problems. This paper
discusses the impact of automation on safety and efficiency. A literature review of ‘Human factors’ mainly from
the aviation and maritime industries is presented to untangle the human-machine relationship characteristics
when increased automation is introduced to operators. A literature review of Hydrodynamics, Guidance,
Navigation and Control (GNC) technologies is presented to introduce the state-of-art and associated
limitations. It is concluded that, if the industry’s drive is safety and efficiency, then full-autonomy is, at present,
not the way to go. Remote control, instead, could facilitate a feasible future, while focused research and
development are in need.
http://www.transnav.eu
the International Journal
on Marine Navigation
and Safety of Sea Transportation
Volume 13
Number 1
March 2019
DOI: 10.12716/1001.13.01.02
34
unmanning maritime assets and the transfer towards
full-autonomy will be discussed.
1.2 Introduction
This literature review is part of a PhD study with the
objective of “evaluation of technology using
simulators”. Main research area is safety and
efficiency of semi-autonomous vessels and the
research scope includes Hydrodynamics; simulation
and testing; and semi-autonomous maneuvering in
close proximity to structures. The simulator facilities
at NTNU include a variety of simulators used for
teaching and research. The use of simulators enables
operators-in-the-loop testing, connecting technology
to humans. The author is studying the man-machine
semi-autonomous maneuvering problem from both
the technology side and the human side. The
technology side is broken down to four scientific
fields: Hydrodynamics; Guidance; Navigation; and
Control. Those four fields reflect the state-of-art in ship
motion prediction and enhancement of automation
level in ship maneuvering. Whereas the field of
Human Factors (relevant to remote operators) is the
field representing the human side of the problem.
These five fields are reviewed briefly in this paper.
The terms may have multiple definitions, therefore, in
this review, the main fields are defined as follows:
Hydrodynamics field in this review refers to
methods that describe the motions and responses
of a ship moving in water using maneuvering and
seakeeping theories such as unified models (Skejic
and Faltinsen, 2008).
Guidance, navigation and control (GNC) is a well-
established technical term used in engineering and
control (cybernetics) fields in topics related to
traveling vehicles; cars, ships, or planes.
Guidance module is the brain of the robotic
controller that is responsible for trajectory
planning, collision avoidance and conforming
to protocol (such as COLREG) (Fossen, 2011).
Navigation module is responsible for
estimating own state, that is, identify own
position and motion information using sensors
and GNSS signals, as well as estimating external
situation, including environment perception
(wind, waves, water depth, etc.) and obstacle
state estimation, that is, identify obstacle
position and motion information (Farell, 2008).
Control is the translation of guidance (desired
trajectory) into actuator instructions that result
in an actual trajectory as close as possible to the
desired one and provides stability to the vehicle
(Pérez, 2005).
Human factors refers to reflections from human
operators as more automation is introduced to
their operations. Sections below include reviews of
each of the fields separately.
2 LITERATURE REVIEW
After the fields of interest were defined, 59 relevant
articles were reviewed in those fields of interest.
Challenges and conflicts are presented. The literature
is found in two ways: Education and search.
Education literature is based on relevant courses and
their relevant references. While Search literature is
based on digital databases search of the following
keywords: Hydrodynamics; seakeeping; maneuvering;
ship simulation; semi-autonomous vessels; unmanned
surface vehicles; guidance; navigation; control; and human
factors. Search results were filtered based on relevance
to the already defined subjects of interest.
2.1 Topic 1: Hydrodynamics
Dynamics is broken down by the studies of
kinematics and kinetics, the former deals with
geometrical aspects of motion and the latter deals
with forces causing the motion. This review is
concerned with ship dynamics, therefore this section
starts with the maneuvering and the seakeeping
theories as foundation for ship dynamics models. The
former is the study of ship moving in constant speed
in calm waters with the assumption that ship motion
is frequency independent, that is, no wave excitation
takes place. The latter is the study of ship motion at
zero or constant speed in waves using frequency
dependent hydrodynamic coefficients.
An overview of methods for describing
maneuvering and seakeeping are grouped into
experimental methods, unified methods, two-time
scale methods and direct calculations by
Computational Fluid Dynamics (CFD) tools
(Quadvlieg et al., 2014). The research is focused on
real-time simulations and on including dynamics-in-
the-loop for marine control systems, therefore the
interest lies in fast mathematical methods such as the
unified methods and the two-time scale methods.
CFD tools are high computationally demanding and
not suitable for real-time simulations. Both examples
presented below, the unified model and the two-time
scale method, are suitable for real-time simulations.
The unified model is a vectorial model that
describes both the maneuvering and seakeeping ship
motions and dates back to 1991 (Fossen, 1991) and is
considered by the international community as a
“standard model” for marine control systems design.
The “standard model” is an upgrade of an earlier
model (the “classical model”) that represents the ship
motion in a component form instead of vector form
and is mostly used in hydrodynamic modeling where
isolated effects are studied.
The 6 degrees-of-freedom (6-DOF) model is
represented as (Fossen, 2011):
 
0

wind wave
Mv C v v D v v g η
g
ττ τ
(1)
where
,,,,,
T
xyz

and
,, , ,,
T
v uvwpqr
are vectors of position / Euler angles and velocities
respectively.
τ vectors are vectors of environment
and control forces and moments. The model matrices
 
, andMC v Dv are inertia, Coriolis and
damping matrices respectively. While

g η is a
vector representing gravitational and buoyancy forces
and
0
g
is a representation of ballast restoring forces
and moments. The model is formulated in the time
domain using the Cummins equation that considers the
impulse response function over the past history of the
excitation force, known as fluid memory effects
(Cummins, 1962).
35
The two-time scale method was proposed by Skejic
and Faltinsen in 2008. It is also a vectorial unified
model that describes both the maneuvering and
seakeeping ship motions. The time domain of the
simulation is divided into two time scales, a slowly
and a rapidly varying one associated with the
maneuvering and the seakeeping respectively. This
method estimates the mean second-order wave loads
(that result in lateral drift caused by incident waves
and wind) “as accurate as possible and at the same
time to be able to simulate real-time maneuvers with
acceptable CPU time.” (Skejic and Faltinsen, 2008, p.
374). The model is represented in a 4-DOF (surge,
sway, roll and yaw) form as follows:
(2)
The main advantage of the two-time scale model is
that it captures the second-order lateral drift
phenomenon. It has better performance in incident
waves, where the mean second-order wave loads
heavily influence the maneuvering behavior. As it
considers theories covering the whole range of
important wavelengths.
For both methods, the potential theory is the main
tool for calculating the hydrodynamic coefficients and
thus forces. This theory assumes water flow across the
rigid body as constant, irrotational, and
incompressible. Chapter 5 of Fossen (2011) covers
hydrodynamic concepts and numerical approaches.
The most common numerical approaches for
calculating the hydrodynamic coefficients are;
Strip Theory; a 2-D theory that considers the flow
variation in the longitudinal-section is much
smaller than that of the cross-section plane of the
ship.
Panel Methods; 3-D integration method that
divides the surface of the ship and the
surrounding water into discrete panels, assigned a
distribution of sources and sinks that fulfil the
Laplace equation.
A comparison of the unified model and the two-
time scale method is of interest for this research,
because the hydrodynamic differences affecting ship
control require further research (Liu et al., 2016).
Several examples of unified numerical models have
been developed in the last three decades and here is a
summary of the latest progress. The method proposed
by Skejic and Faltinsen in 2008 was verified and
validated for calm water. This method is further
developed in a study on ship-to-ship hydrodynamic
interaction effects between two ships going ahead in
regular waves, it highlights critical maneuvering
situations and it still requires experimental validation
(Skejic and Berg, 2010). In 2013, the two-time scale
model was applied to irregular seas and validated for
a container ship (Skejic and Faltinsen, 2013).
Hermundstad and Hoff (2009) implemented a time
domain unified model on submarines and compared
with experimental results. It was argued that the used
unified model did not describe the diving maneuvers
correctly because the depth dependency of the
coefficients was not incorporated. A practical method
for ship motion simulation using the two-time scale
method is presented by Yasukawa and Nakayama
(2009) that derives 6-DOF equations of motion for the
high frequency problem and 4-DOF equations of
motions for the low frequency problem. Wave
induced motions for turning maneuver are predicted
for a container ship of geometry S-175 and the
predictions resulted in rough agreement with free
model tests. Yasukawa, Amri Adnan and Nishi
(2010)
compared, numerical estimates of hydrodynamic
forces and wave-induced motions taking into account
lateral drift, with experiments showing that drift
effects are not negligible and that the method is able
to capture them. Seo and Kim (2011) extended the
WISH (computer program for nonlinear Wave
Induced load and Ship motion analysis) by coupling
the maneuvering and the seakeeping models, and
verified it by comparing with published experiment
data in calm weather and regular waves. The
simulations showed fair agreement of overall
tendency in maneuvering trajectories.
Beside lateral drift, the broaching phenomenon is
another challenge for hydrodynamic models, it
concerns loss of stability while sailing in following
seas where the kinetic energy of the ship along the
forward axis transfers to roll motion and leads to
strong heel, loss of heading, even capsize (Wu,
Spyrou and McCue, 2010). Generally, maneuvering in
waves is a challenge for both experimental and
numerical modelling. For simulating ship motion in
waves, forces and hydrodynamic coefficients need to
be calculated dependent on wave frequency, ship
heading and angle of attack angle between wave
direction and ship course (Kim et al., 2014).
2.2 Topic 2: Guidance
The guidance system receives information about the
world, both internal information concerning the ship
maneuvering and engine status, and external
concerning the surroundings, environmental loading
and nearby target ships and other objects and
translates this information into instructions to
controllers. Guidance is responsible for path planning,
including collision avoidance. Fossen (2011) defines
motion objectives categories. The guidance system
together with the control system should fulfill the
motion objectives according to one of the following
categories:
1 Setpoint regulation: heading angle is constant with
no consideration of time.
2 Path following: heading angle is variable,
following a path, no consideration of time.
3 Trajectory tracking: heading angle is variable,
following a trajectory in both space and time.
4 Maneuvering: considers the overall feasibility of the
path, often with more importance to space than
time. To incorporate COLREG, the guidance
system shall consider both space and time
parameters because velocities of maneuvers are
critical.
36
The guidance system tasks are grouped into two:
global and local path planning. The global path
planning approach is the deliberate part of the
guidance system. It is an optimized plan of the path
from starting point of the trip to the end point, it
includes known information about traffic, weather
forecast, ship properties, land/islands, shallow waters
and buoys. This is a multi-objective optimization
problem and usually done offline and requires large
computational requirements, in which, optimization
methods and heuristic search algorithms are the two
main methods. While local path (re)planning
approach is the reflexive part of the guidance system,
it takes charge of planning local deviations from the
global plan, in case the navigation system detected an
approaching object. A characteristic requirement of
local path re-planning is the low computational
requirements, where real-time methods such as line-
of-sight (LOS) and potential fields are common.
Polvara et al. (2018) presented a review of global
and local planning methods including a section for
advanced computing-based methods. The author
stated that almost all of the methods reviewed did not
consider uncertainties due to environment loads and
vehicle dynamics. A recent review of trajectory
planning and tracking review (for autonomous
driving systems) concluded that even most advanced
guidance and control algorithms, with today’s
available sensor technology, work well under
regulated environments assuming knowledge of
surroundings and weather conditions. It also states
that the inclusion of vehicle dynamics and
environmental loads increases the effectiveness of
such controllers (Dixit et al., 2018). Lately, Wiig et al.
proposed an integral line-of-sight law in the presence
of constant ocean currents (2018).
LaValle in his tutorials points out that “the basic
problem of computing a collision-free path for a robot
among known obstacles is well understood and
reasonably solved; however, deficiencies in the
problem formulation itself and the demand of
engineering challenges in the design of autonomous
systems raise important questions and topics for
future research” (LaValle, 2011, p. 108)
Polvara et al in their recent review stated the
following: “It has been concluded that almost all the
existing methods do not address sea or weather
conditions, or do not involve the dynamics of the
vessel while defining the path. Therefore, this
research area is still far from being considered fully
explored.” (Polvara et al., 2018, p. 241).
2.3 Topic 3: Navigation
The navigation system collects data from various
sources such as sensors, cameras and satellites, and
transfers the data into information of two kinds, state
estimation and environment perception. State
estimation is information about the ship’s motion,
mainly location and velocities. Environment
perception is weather information, wind, waves,
currents, and information about the surrounding as
well, including state of target ships and objects. The
scope of this system vastly increases as level of
automation increases; the number of datasets, their
resolution, frequency, quality and size are vastly
increasing in remotely controlled vessels comparing
to conventional ones. Moreover, since making sense
of the collected data is considered part of the
navigation system, its scope should then include
advanced computing methods in order to deliver a fit-
for-purpose output. Methods such as machine
learning, sensor fusion, computer vision, prediction,
and anomaly detection are now used within the
navigation system for making sense of the collected
data.
On board sensors are susceptible to disturbances
that come from the environment, ship motion and
other noise sources. The disturbances cause
uncertainties in the perception model. This leads to
control errors that accumulate over time, and result in
undesired control behavior. Therefore, data from
multiple sources are correlated against each other to
calculate position and velocity estimates as accurate
as possible. Data sources involved in a navigation
system are:
1 Inertial measurement unit (IMU) is an onboard
three-dimensional navigation system that
comprises of three mutually-orthogonal
accelerometers and three gyroscopes to give the
position, velocity and altitude of own ship. IMU is
often used with (and aided by) satellite positioning
to provide drift-free positioning.
2 Automatic Identification System (AIS) is a very-
high frequency communication system used by
ships to transmit their identity, position, velocity,
destination and other information and in return
they receive information of nearby ships. Even
though AIS is mandatory for commercial vessels,
not all boats have it onboard!
3 GNSS is a global positioning solution system. It
transmits radio signals from satellites orbiting the
planet to the ship. There are a number of GNSS
solution providers including GPS, GLONASS,
Beidou and Galileo.
4 Radar, an acronym for radio detection and
ranging, uses radio waves to detect ships and
obstacles within a long range but its capability of
detecting small moving targets is limited. Radar
wavelength passes through fog and rain and it
provides nearly all-weather data imagery.
5 Lidar, an acronym for light detection and ranging,
is a high resolution and accuracy object detection
sensor for near-range.
6 Sonar, an acronym for sound navigation ranging,
detects submerged objects such as reefs, sunken
ships and submarines. The sonar transmits
ultrasonic pulses, receives the reflected echoes and
displays a picture of the detected objects.
7 Other types of sensors and tools are used for
navigation purposes such as cameras, infrared
sensors, compass systems, navigation lights and
ship whistles.
Most common method for fusing the navigation
data as of today is the Kalman filter. The Kalman
filter, invented by Kalman in 1960, is a real-time
Bayesian estimation algorithm that uses all available
measurements over time, and uses knowledge of
deterministic and statistical properties of the system
parameters in order to provide optimal minimum-
error state estimations (Groves, 2013).
Examples of recent perception technologies in
navigation systems are:
37
1 Non-linear observers: Advanced alternatives to the
well-established Kalman filter, with proven
stability properties and lower computational
demands (Fossen and Strand, 1999; Aschemann,
Wirtensohn and Reuter, 2016; Bryne, 2017).
2 Extended Kalman filter (EKF) for position and
velocity estimation using GPS and compass
measurements (Caccia et al., 2008; Bibuli et al.,
2009; Tran et al., 2014).
3 Unscented Kalman filter (UKF) for state estimation
without previous knowledge of noise
characteristics (Peng, Han and Huang, 2009;
Vasconcelos, Silvestre and Oliveira, 2011).
4 Inverted Kalman filer (IKF) bounds model
uncertainties that come from environment
variability (Motwani et al., 2013).
5 The eXogenous Kalman filter (XKF) for providing
covariance estimates for the estimated states
generated by non-linear observes (Johansen and
Fossen, 2017).
6 Wave information perception using camera (Liu
and Wang, 2013). Stereo vision system that
generate probabilistic hazard maps and provide
estimates for speed and heading of target objects
(Huntsberger et al., 2011).
2.4 Topic 4: Control
The control system is responsible to translate the
information collected from the guidance system and
communicate it with the actuators as commands.
Actuators such as propellers, thrusters, and rudder
receive commands from the control system and
execute actions producing forces and moments that
affect the state of the ship, approaching the desired
state. The control system is responsible to make sure
that the generated actuator commands are practical
for the underactuated ship given the actuator
limitations and ship dynamics.
Control literature is rich with control design
approaches that extend from the classical
proportional-integral-derivative (PID) controllers to
the more advanced artificial-intelligence (AI) based
controllers. Practical ship control often applies a
combination of different control methods. PID control
approaches are the most favored, they are, however,
suitable for single-input-single-output cases such as
heading control (Minorsky, 1922). This approach
could suffer severe actuator damage caused by high
waves. Simultaneous control of velocity and heading
solves this problem. Multivariable control was
realized by multi-loop PID control (Lefeber, Pettersen
and Nijmeijer, 2003) and fuzzy adaptive control
techniques (Le et al., 2003).
Multivariable control has been widely approached
by optimal control techniques such as H-infinity and
Linear quadratic optimal techniques. Nonetheless,
Linear Quadratic Regulator (LQR) controller suffers
from the assumption that all states are measurable
and known, which is not the case. Linear Quadratic
Gaussian (LQG) controller together with a Kalman
filter estimates in real-time the unknown states,
however, suffers from instability. Instability outside
predefined domain and discontinuities are major
drawbacks of adaptive linear control methods (Liu et
al., 2016). Non-linear methods, such as Fuzzy logic
control, Neural networks and Lyapunov-based
methods argue that they can potentially overcome
stability related issues while maintain smooth time-
parametrized trajectories (Aguiar and Hespanha,
2003).
2.5 Topic 5: Human Factors
In this section the definition of levels of automation
(or autonomy; since both terms are used
interchangeably) is presented and followed by
explanations of the human factors faced by operators
introduced to increased automation in their
operations.
2.5.1 Levels of automation (LOA)
Levels of automation were developed in the 1978.
They were used to describe systems and aid the
communication in the design phase of automated
systems (Sheridan and Verplank, 1978). Multiple
versions of LOAs have been issued since then. In the
ship industry, LOA proposals exist from multiple
sources such as Bureau Veritas, Lloyd’s Register, the
Norwegian Forum for Autonomous Ships (NFAS),
Rolls-Royce, and others. Table 1 shows the LOAs as
proposed by NFAS. General agreement exists in the
different definitions as they range from human-
operated ship (lowest level) to fully autonomous ship
(highest level).
Explicitly, all the different variations of LOA
classifications, agree that, on the highest level of
automation, the machine decides and acts, and
requires no communication with the human.
2.5.2 Increased automation
Automation is intended to increase safety and
efficiency, however, in complex tasks (dynamic
environments involving many variables) it changes
the nature of the human-role in the task, it affects
areas such as workload and cognitive demands.
Moreover, the resultant impact of (increased)
automation turns out to be more complex than
anticipated. The changes are qualitative in context
rather than quantitative and uniform (Woods et al.,
1996). Main human factors involved in the operator-
technology interface are summarized as follows,
including responsibility, surprises of automation,
management by exception and communication:
1 Responsibility: Decisions that the human operator
is used to take and implement will be routinely
delegated to machines. However, can
responsibility be delegated as well? Responsibility
perception and calibration of trust between
humans and machines are important to safe
autonomous operations (Muir, 1987).
Jordan was one of the first to stress out that “we
can never assign them (i.e., the machines) any
responsibility for getting the task done;
responsibility can be assigned to man only”
(Jordan, 1963, p. 164). As suggested by Billings,
human operators bear ultimate responsibility for
operational goals, they must be in command, well
involved and well informed about ongoing
autonomous activities (Billings, 1991).
38
Table 1. LOAs as proposed by NFAS (Rødseth and Nordahl,
2017).
_______________________________________________
Level LOA Description
name
_______________________________________________
1 Decision This corresponds to today’s and
support tomorrow’s advanced ship types with
relatively advanced anti-collision radars
(ARPA), electronic chart systems and
common automation systems like
autopilot or track pilots. The crew is still
in direct command of ship operations
and continuously supervises all
operations. This level normally
corresponds to "no autonomy".
2 Automatic The ship has more advanced
automation systems that can complete
certain demanding operations without
human interaction, e.g. dynamic
positioning or automatic berthing. The
operation follows a pre-programmed
sequence and will request human
intervention if any unexpected events
occur or when the operation completes.
The shore control centre (SCC) or the
bridge crew is always available to
intervene and initiate remote or direct
control when needed.
3 Constrained The ship can operate fully automatic in
autonomous most situations and has a predefined
selection of options for solving
commonly encountered problems, e.g.
collision avoidance. It has defined limits
to the options it can use to solve
problems, e.g. maximum deviation from
planned track or arrival time. It will call
on human operators to intervene if the
problems cannot be solved within these
constraints. The SCC or bridge
personnel continuously supervises the
operations and will take immediate
control when requested to by the
system. Otherwise, the system will be
expected to operate safely by itself.
4 Fully The ship handles all situations by itself.
autonomous This implies that one will not have an
SCC or any bridge personnel at all. This
may be a realistic alternative for
operations over short distances and in
very controlled environments.
However, and in a shorter time
perspective, this is an unlikely scenario
as it implies very high complexity in
ship systems and correspondingly high
risks for malfunctions and loss of
system.
_______________________________________________
2 “Automation surprises”: It could be difficult for
the operators to follow up with the autonomous
vehicle and understand the grounds for its
decisions. When the actions of the “machine” are
not similar to what the human operator would do
if placed in the same situation then the human
would lose track and fail to predict next steps. A
simulator experiment to evaluate pilots’ mode
awareness was carried out that confirmed that
“automation surprises” are experienced even by
operators with extensive amount of line experience
on similar highly autonomous aircrafts. It was
shown that in non-normal situations, more
problems related to “automation surprises”
occurred (Sarter and Woods, 1994). A previous
study by Wiener, who conducted a survey of B-757
pilots, resulted that 55% of respondents were still
being surprised by the automation after more than
one year of line experience on the aircraft (Wiener,
1989). Norman referred to the phenomenon of
human operator losing track of machine’s behavior
as ‘breakdown in mode awareness’ which has been
linked strongly to the following factors:
automation surprises, increased error possibilities,
new cognitive demands, and failure to intervene
appropriately. Increased automation would also
cause surprise to the ship designers and owners
who experience unexpected consequences because
their automated system fails to behave as was
intended (Norman, 1988).
3 Management by exception: A remote operator,
whether monitoring or supervising, is in a double
bind dilemma with the machine. A dilemma
between trust and takeover. Dekker and Woods
explained this phenomenon in their work titled
“To intervene or not to intervene: the dilemma of
management by exception” (Dekker and Woods,
1999). Supervisory control places the operator in a
decision-making situation. A trade-off between
intervening too early, before enough evidence is
collected about the situation, and intervening too
late, after it escalades into an irreversible crisis.
The operators, for every moment in time, must
assess the criticality of the situation and decide
whether to intervene or not. Late decisions are
catastrophic. Early decisions are not justified.
Decision aids and prediction tools are required,
but how much should they be trusted? (Sheridan,
2000).
Human-machine interaction is changing in nature.
Increased automation reduces workload in
normal-times and increase them dramatically in
non-normal times. In non-normal times the
‘automation surprises’ factor is higher, the ‘mode
awareness’ factor is lower, the attentional
demands and the cognitive demands are highly
increased. Given the dilemma, this setting is
critical in non-normal times as it leads to less
situational awareness (SA) and less intervention
capabilities. Thus, safety is a big concern if things
went wrong in non-normal times. Sarter et al
define the term Mode awareness as “the ability of
a human operator to track and anticipate the
behavior of automated system” (Sarter, Woods
and Billings, 1997, p. 6). Situational awareness,
according to Endsley, is “the perception of the
elements in the environment within a volume of
space and time, the comprehension of their
meaning, and the projection of their status in the
near future” (Endsley, 1995, p. 36).
4 Communication: For example, the grounding of
the Royal Majesty is referred to as a loss of
situational awareness problem; a communication
problem because of increased automation. Among
other factors, the GPS has failed, positioning
information were incorrect, autopilot used the
faulty information, the ship drifted and that was
not apparent to the crew. They believed that the
sailing was flawless but in fact, it lead to a
grounding (Lützhöft and Dekker, 2002).
Researchers emphasized on the value of
communication and collaboration with the
machine for safer autonomous navigation.
Effective communication and coordination
39
between humans and machines is believed to be
key for successful operations (Sarter, Woods and
Billings, 1997).
3 RESULTS
This review covers topics concerning the future of
autonomous vessels from three perspectives. First
from the side of the technology advancements that
make such a future possible. Second from the human
operator side and the challenges faced while teams
are operating highly autonomous systems, remotely.
Third from the levels-of-automation side, multiple
versions of LOA definitions for the maritime industry
that classify human-machine relationship as
automation increases. Trying to answer the article’s
question.
One may argue that the supporting technologies
are already available, as there are booming examples
of domain-specific advancement, but this review
identifies a shortage in the studies that show how
well these building blocks work out together, and
under uncertainties. Analysis and breakdown of this
identified shortage follows:
There is interaction and signal flow among the
GNC and hydrodynamics fields, as shown in Figure
1. One publication proposing a novel path planning
method would have built-in assumptions regarding
(and pre-selections of) ship dynamic models,
navigation methods, and control design approaches.
For example, Liu, Bucknall and Zhang (2017)
proposed a guidance “fast marching” method for a
USV, and presented their results of full-scale
experiments. They used a preselection of navigation
methods (Kalman filter), control methods (PID
autopilot), and vehicle dynamics model (3-D model)
as in (Motwani et al., 2013).
Figure 1. GNC module interaction and signal flow (Fossen,
2011, p. 233)
Given the interrelation, applications of semi-
autonomous vessels, both real (full-scale) and virtual
(simulators), require a package of GNC and
hydrodynamics technologies interacting together. It is
widely agreed that the performance of methods is
largely altered by uncertainties coming from
environmental loads and ship dynamics (LaValle,
2011; Liu et al., 2016; Polvara et al., 2018).
“Automation could increase sources of error”.
Precautionary perspective is necessary in research
and development. Porathe et al. (2018) includes a
fictive story that predicts a possible future scenario in
one of the Norwegian fjords and provides a forecast
of the risk picture in the maritime industry.
Human operators face challenges with highly
automated systems. In the future, as autonomous
ships become reality, advancing through the LOA
scale, until eventually, full-autonomous vessels are
realized in a safe and efficient manner, remote control
will be essential. Safe and efficient remote operations
are as important as, or even more important than, no-
human-interaction type of control (according to LOA
definitions of full autonomy). There is a literature
shortage in this multidisciplinary field of “ship
remote control”. It should cover remote control-
centered topics of ship design, GNC systems design,
human-machine interaction, navigation functions,
interface, and remote control center design.
4 DISCUSSIONS
4.1 Disagreements
Viewpoints such as “A ship must follow and adhere
to the international regulations for preventing
collisions at sea (COLREGS)” are common in GNC
technology research. However, these viewpoints
oversimplify the problem. They inherently assume
that traffic in the sea is well regulated and all players
follow the rules. In reality, operators and crew do
violate procedures, for different reasons, as shown by
a research that collected 1262 questionnaires from
tankers and bulk carriers crew (Oltedal, 2011).
Some collision avoidance methods enable manual
input of waypoints by a supervisor operator for
replanning the path and avoiding approaching
obstacle. As Campbell et al describe them: “This is not
the most efficient method for avoidance and is subject
to operator error” (Campbell, Naeem and Irwin,
2012). This view is common. It promote two points.
First, that human operators are subject to more errors
than machines. Second, researchers are oriented to
develop technologies with high automation level and
low human interaction, to avoid human errors. This
view conflicts with the status quo of technology,
because also machines are subject to error, and it
conflicts with human factors research, that having less
human interaction with highly automated system
introduces the dilemma of management by exception
and it can be avoided by having human input and
authority over the machines even for highly
autonomous systems.
In a recent survey on communication technologies
(Zolich et al., 2018) a relation of LOAs with
communication requirements was presented. It says,
basically, that the higher the LOA is, the lower the
amount of data the ship would require to
communicate with land. This view conflict with the
human operator’s requirements for safe and efficient
monitoring, supervision and control of the
autonomous remote asset.
The definition of full-autonomy, in all the
variations of LOA scales, emphasizes on “no human
interaction; machine ignores human; no human
input”. These definitions favor automation over safety
and efficiency of the asset because full-autonomous
ships need to be remotely controlled, on demand,
upon the decision of the supervisor in charge. In such
a dynamic multi-objective shipping task, the option of
40
remote control is necessary; the reason for this desire
of remote control could be any of the following
examples:
Business and market fluctuation
Environment regulations and emission related
rules
Cyber-attacks, piracy and hijacking
Environment loads and extreme weather
Incidents at ports such as fires or chain-reaction
accidents
Customer relations; cargo health; maintenance
issues and etc.
4.2 Main challenges
Main challenges from the different perspectives are
summed up in this section as follows.
Motion coupling: control advancements consider
a simplified ship model, similar to that of a 3-D
unicycle model. The effect of motion coupling to
stability requires further analysis.
Ship motion in waves: Describing ship motion in
harsh weather is a challenge; there is no standard way
of doing it. Hydrodynamic research considers that
ship motion in calm water is assumed to converge to
an underlying true trajectory. The maneuvering
committee of the 27
th
ITTC address this issue as a
challenge for both experimental work and numerical
modelling (Quadvlieg et al., 2014, sec. 6.4).
Co-simulation of digital models: The
development of GNC algorithms has boomed lately. It
is challenging to know how they will work together
under the influence of stochastic environmental
loading and uncertainties. In addition, how will the
human (remote) operator experience those
advancements?
Remote operator input: How well does these
technologies workout together? Research towards
enhancing the performance of man-machine systems
in dynamic control tasks is crucial in design and
operation of future maritime operations. Effects of
LOA towards situation awareness and mental
workload are researched (Kaber and Endsley, 2004).
However, it is challenging to judge automation based
on the LOA scale because the whole scale is course;
massive variations could be possible within one LOA
level. Variations in terms of interface, controllers,
inputs, outputs and engagement level are expected
for each level.
4.3 Full Autonomy
The main challenges of the previous section maps
man-machine challenges that are valid for the
maritime industry as of today. Worldwide research
and development projects will certainly tackle them
and innovations will pave the way, gradually, to
realizations of higher levels of ship autonomy. The
progress will be gradual, evolutionary rather than
revolutionary, because of the political, legal and
financial inertia involved in such industry.
Systematic bias: Assume that “we are dealing
with a transition towards fully autonomous systems”
with the main objectives “safety and efficiency”. The
way the developers perceive the future is key in
determining the safety and efficiency of that future.
The definitions of LOAs form a huge anchoring bias
that weakens the focus on the objectives and
strengthens the following views:
1 The ultimate goal is full autonomy
2 Full autonomy is that systems run by themselves
with no human interaction
3 Human input is a negative contribution to system
objectives
And those views are expanding systematically
within and across industries and can be seen popular
in technical scientific disciplines and among the youth
in societies of most industrial countries. Broek et al
(2017) mentioned the need of a man-machine
“collaboration framework” even for fully autonomous
systems.
Towards full-autonomy: As it brings value to
other industries, the values of advanced technology
must be harvested in the shipping industry as well.
We strive for fewer accidents, less social and
environmental impact by the utilization of tools such
as data analytics, decision support aids, and advanced
autopilots. Surprisingly, the GNC literature shows
that technology is being developed towards a future
with no human interaction. However, I think that the
values of full-autonomy cannot be harvested unless
the technology becomes developed towards a future
with full human interaction.
5 CONCLUSIONS
If the industry’s drive is safety and efficiency, then
full-autonomy is, at present, not the way to go.
Remote control, instead, could facilitate a feasible
future, while focused research and development are
in need. From the technology side, the literature
shows that uncertainties coming from environmental
loads and ship dynamics largely affect the
performance of GNC technologies in a semi-
autonomous vessel. Thus, accurate modeling and
prediction of semi-autonomous maneuvering is
fragile under uncertainties. From the human side, the
literature shows that as automation is increased and
interaction is decreased the operators face the
management by exception dilemma. Operators
undertaking safe and efficient ship remote control,
even for highly autonomous ships, require high
interaction and high authority over the system.
Automation is promising because of the possible
reduction of cost and risk involved in maritime
operations, nevertheless; it could bring in new sources
of error, while human operators face serious
challenges dealing with highly automated systems.
There is a rush of technology-related research but
there is a lack of holistic research focusing on “ship
remote control”. Research that tests the GNC
technologies under uncertainties with human-
operator in-the-loop is needed. Digital advancements
enable virtual experiment environments with human
interaction such as simulators. Those safe
environments could be the only tools available, for
now, to enable us research whether it is full-
autonomy the right way to go for exploiting the ocean
potentials.
41
REFERENCES
Aguiar, A. P. and Hespanha, J. P. (2003) ‘Position Tracking
of Underactuated Vehicles’, Acc 2003. doi:
10.1109/ACC.2003.1243366.
Aschemann, H., Wirtensohn, S. and Reuter, J. (2016)
‘Nonlinear Observer-Based Ship Control and
Disturbance Compensation’, IFAC-PapersOnLine.
Elsevier B.V., 49(23), pp. 297–302. doi:
10.1016/j.ifacol.2016.10.358.
Bibuli, M. et al. (2009) ‘Path-following algorithms and
experiments for an unmanned surface vehicle’, Journal of
Field Robotics. doi: 10.1002/rob.20303.
Billings, C. E. (1991) Human-Centered Aircraft Automation: A
Concept and Guidelines, NASA Technical Memorandum
103885. doi: 10.1207/s15327108ijap0104_1.
Broek, V. der et al. (2017) ‘Approaching full autonomy in the
maritime domain: paradigm choices and Human Factors
challenges’, in MTEC 2017, pp. 1–11.
Bryne, T. H. (2017) Nonlinear Observer Design for Aided
Inertial Navigation of Ships. Available at:
http://hdl.handle.net/11250/2440519.
Caccia, M. et al. (2008) ‘Basic navigation, guidance and
control of an Unmanned Surface Vehicle’, Autonomous
Robots, 25(4), pp. 349–365. doi: 10.1007/s10514-008-9100-
0.
Campbell, S., Naeem, W. and Irwin, G. W. (2012) ‘A review
on improving the autonomy of unmanned surface
vehicles through intelligent collision avoidance
manoeuvres’, Annual Reviews in Control. Elsevier Ltd,
36(2), pp. 267–283. doi: 10.1016/j.arcontrol.2012.09.008.
Cummins, W. E. (1962) ‘The impulse response function and
ship motions’, Symposium on Ship Theory. Hamburg.
Dekker, S. W. . and Woods, D. . (1999) ‘To intervene or not
to intervene the dilemma of management by exception’,
Cognition, Technology & Work. doi:
10.1007/s101110050035.
Dixit, S. et al. (2018) ‘Trajectory planning and tracking for
autonomous overtaking: State-of-the-art and future
prospects’, Annual Reviews in Control. Elsevier Ltd, 45,
pp. 76–86. doi: 10.1016/j.arcontrol.2018.02.001.
Endsley, M. R. (1995) ‘Toward a Theory of Situation
Awareness in Dynamic Systems’, Human Factors: The
Journal of the Human Factors and Ergonomics Society. doi:
10.1518/001872095779049543.
Farell, J. A. (2008) Aided Navigation: GPS with High Rate
Sensors, The McGraw-Hill Companies. doi:
10.1017/CBO9781107415324.004.
Fossen, T. I. (1991) ‘Adaptive control of nonlinear
underwater robotic systems’, Modeling, Identification and
Control, 12(2), pp. 95–105.
Fossen, T. I. (2011) Handbook of Marine Craft Hydrodynamics
and Motion Control, John Wiley & Sons Ltd.
Fossen, T. I. and Strand, J. P. (1999) ‘Passive nonlinear
observer design for ships using Lyapunov methods’,
Automatica, pp. 3–16.
Groves, P. (2013) Principles of GNSS, Inertial, and Multisensor
Integrated Navigation Systems, Journal of Chemical
Information and Modeling. doi: 10.1109/MAES.2014.14110.
Hermundstad, E. M. and Hoff, J. R. (2009) ‘Time domain
simulation of the maneuvering of a vessel in a seaway’,
in Proceedings of the International Conference on Offshore
Mechanics and Arctic Engineering - OMAE. doi:
10.1115/OMAE2009-79946.
Huntsberger, T. et al. (2011) ‘Stereo vision-based navigation
for autonomous surface vessels’, Journal of Field Robotics.
doi: 10.1002/rob.20380.
Johansen, T. A. and Fossen, T. I. (2017) ‘The eXogenous
Kalman Filter (XKF)’, International Journal of Control. doi:
10.1080/00207179.2016.1172390.
Jordan, N. (1963) ‘Allocation of functions between man and
machines in automated systems’, Journal of Applied
Psychology. doi: 10.1037/h0043729.
Kaber, D. B. and Endsley, M. R. (2004) ‘The effects of level of
automation and adaptive automation on human
performance, situation awareness and workload in a
dynamic control task’, Theoretical Issues in Ergonomics
Science. doi: 10.1080/1463922021000054335.
Kim, Y. et al. (2014) Seakeeping Committee Final Report and
Recommendations to the 27th ITTC. Copenhagen.
Available at:
https://ittc.info/downloads/proceedings/27th-
conference-copenhagen-2014/volume-1/ on 19/12/2018.
LaValle, S. (2011) ‘Motion Planning: Part II’, IEEE Robotics &
Automation Magazine, 18(2), pp. 108–118. doi:
10.1109/MRA.2011.941635.
Le, M. D. et al. (2003) ‘A new and effective fuzzy PID
autopilot for ships’, in Proceedings of IEEE International
Symposium on Computational Intelligence in Robotics and
Automation, CIRA. doi: 10.1109/CIRA.2003.1222204.
Lefeber, E., Pettersen, K. Y. and Nijmeijer, H. (2003)
‘Tracking control of an underactuated ship’, IEEE
Transactions on Control Systems Technology. doi:
10.1109/TCST.2002.806465.
Liu, K. K. and Wang, J. H. (2013) ‘A Method of Detecting
Wave Grade Based on Visual Image Taken by USV’,
Applied Mechanics and Materials. Trans Tech Publications,
291–294, pp. 2437–2441. doi: 10.4028/
www.scientific.net/AMM.291-294.2437.
Liu, Y., Bucknall, R. and Zhang, X. (2017) ‘The fast marching
method based intelligent navigation of an unmanned
surface vehicle’, Ocean Engineering. Elsevier Ltd,
142(April), pp. 363–376. doi: 10.1016/
j.oceaneng.2017.07.021.
Liu, Z. et al. (2016) ‘Unmanned surface vehicles: An
overview of developments and challenges’, Annual
Reviews in Control. Elsevier Ltd, 41, pp. 71–93. doi:
10.1016/j.arcontrol.2016.04.018.
Lützhöft, M. H. and Dekker, S. W. A. (2002) ‘On your watch:
Automation on the bridge’, Journal of Navigation. doi:
10.1017/S0373463301001588.
Minorsky, N. (1922) ‘Directional Stability of Automatically
Steered Bodies’, Journal of the American Society for Naval
Engineers. doi: 10.1111/j.1559-3584.1922.tb04958.x.
Motwani, A. et al. (2013) ‘Interval Kalman filtering in
navigation system design for an uninhabited surface
vehicle’, Journal of Navigation. doi: 10.1017/
S0373463313000283.
Muir, B. M. (1987) ‘Trust between humans and machines,
and the design of decision aids’, International Journal of
Man-Machine Studies. doi: 10.1016/S0020-7373(87)80013-5.
Norman, D. A. (1988) The Psychology of Everyday Things, New
York, Basic Books. New York, Basic Books. doi:
10.2307/1423268.
Oltedal, H. A. (2011) Safety culture and safety management
within the Norwegian-controlled shipping industry.
University of Stavanger.
Peng, Y., Han, J. and Huang, Q. (2009) ‘Adaptive UKF based
tracking control for Unmanned Trimaran Vehicles’,
International Journal of Innovative Computing, Information
and Control, 5, pp. 3505–3515.
Pérez, T. (2005) Ship Motion Control: course keeping and roll
stabilisation using rudder and fins. Springer. doi: 10.1007/1-
84628-157-1.
Polvara, R. et al. (2018) ‘Obstacle Avoidance Approaches for
Autonomous Navigation of Unmanned Surface
Vehicles’, Journal of Navigation, 71(1), pp. 241–256. doi:
10.1017/S0373463317000753.
Porathe, T. et al. (2018) ‘At least as safe as manned shipping?
Autonomous shipping, safety and “human error”’, in
Safety and Reliability – Safe Societies in a Changing World.
Proceedings of ESREL 2018, June 17-21, 2018, Trondheim,
Norway.
Quadvlieg, F. et al. (2014) Manoeuvring Committee: Final
Report and Recommendations to the 27th ITTC.
Copenhagen. Available at:
https://ittc.info/downloads/proceedings/27th-
conference-copenhagen-2014/volume-1/ on 19/12/2018.
Rifkin, J. (2012) The Third Industrial Revolution- How Lateral
Power Is Transform-ing Energy, The Economy, and the
42
World, World Future Review (World Future Society). doi:
978-0230341975.
Rødseth, Ø. J. and Nordahl, H. (2017) Definitions for
autonomous ships by NFAS. Available at:
http://nfas.autonomous-ship.org/resources/autonom-
defs.pdf.
Sarter, N. B. and Woods, D. D. (1994) ‘Pilot Interaction With
Cockpit Automation II: An Experimental Study of Pilots’
Model and Awareness of the Flight Management
System’, The International Journal of Aviation Psychology.
doi: 10.1207/s15327108ijap0401_1.
Sarter, N., Woods, D. D. and Billings, C. E. (1997)
‘Automation surprises’, Handbook of human factors and
ergonomics. doi: 10.1207/s15327108ijap0204_5.
Seo, M. G. and Kim, Y. (2011) ‘Numerical analysis on ship
maneuvering coupled with ship motion in waves’, Ocean
Engineering. doi: 10.1016/j.oceaneng.2011.09.023.
Sheridan, T. B. (2000) ‘HCI in supervisory control: Twelve
dilemmas’, in Human Error and System Design and
Management. doi: 10.1038/scientificamerican0798-64.
Sheridan, T. B. and Verplank, W. L. (1978) Human and
Computer Control of Undersea Teleoperators.
Skejic, R. and Berg, T. E. (2010) ‘Combined seakeeping and
maneuvering analysis of a two ship lightering
operation’, Proceedings of the 29th Annual International
Conference on Offshore Mechanics and Arctic Engineering
(OMAE 2010). doi: 10.1115/OMAE2010-20616.
Skejic, R. and Faltinsen, O. M. (2008) ‘A unified seakeeping
and maneuvering analysis of ships in regular waves’,
Journal of Marine Science and Technology, 13(4), pp. 371–
394. doi: 10.1007/s00773-008-0025-2.
Skejic, R. and Faltinsen, O. M. (2013) ‘Maneuvering behavior
of ships in irregular waves’, in Proceedings of the ASME
32nd International Conference on Ocean, Offshore and Arctic
Engineering OMAE 2013. Nantes, France.
Tran, N.-H. et al. (2014) Tracking Control of an Unmanned
Surface Vehicle, AETA 2013: Recent Advances in Electrical
Engineering and Related Sciences. Edited by I. Zelinka, V.
H. Duy, and J. Cha. Berlin, Heidelberg: Springer Berlin
Heidelberg.
Vasconcelos, J. F., Silvestre, C. and Oliveira, P. (2011)
‘INS/GPS Aided by Frequency Contents of Vector
Observations With Application to Autonomous Surface
Crafts’, IEEE Journal of Oceanic Engineering, 36(2), pp.
347–363. doi: 10.1109/JOE.2011.2126170.
Wiener, E. L. (1989) Human Factors of Advanced Technology
(‘Glass Cockpit’) Transport Aircraft, NASA-CR-177528. doi:
10.13246/j.cnki.iae.2015.02.004.
Wiig, M. S. et al. (2018) ‘An Integral Line-of-Sight Guidance
Law with a Speed-dependent Lookahead Distance’, in
European Control Conference 2018 (ECC 2018). Limassol,
Cyprus, pp. 1269–1276. doi: 10.23919/ECC.2018.8550243.
Woods, D. D. et al. (1996) ‘Bridging The Gap Between User-
Centered Intentions and Actual Design Practice’, in
Proceedings of the Human Factors and Ergonomics Society
40th Annual Meeting-1996, pp. 967–971.
Wu, W., Spyrou, K. J. and McCue, L. S. (2010) ‘Improved
prediction of the threshold of surf-riding of a ship in
steep following seas’, Ocean Engineering. doi:
10.1016/j.oceaneng.2010.04.006.
Yasukawa, H., Amri Adnan, F. and Nishi, K. (2010) ‘Wave-
Induced Motions on a Laterally Drifting Ship’, Ship
Technology Research. doi: 10.1179/str.2010.57.2.001.
Yasukawa, H. and Nakayama, Y. (2009) ‘6-DOF motion
simulations of a turning ship in regular waves’, in
International Conference on Marine Simulation and Ship
Manoeuvrability.
Zolich, A. et al. (2018) ‘Survey on Communication and
Networks for Autonomous Marine Systems’, Journal of
Intelligent and Robotic Systems: Theory and Applications,
pp. 1–25. doi: 10.1007/s10846-018-0833-5