BALDI Licentiatethesis Improvingshipenergyefficiencythroughasystemsperspective PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 135

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/263467202

Improving ship energy efficiency through a systems perspective

Thesis · December 2013

CITATIONS READS

10 753

1 author:

Francesco Baldi
ENEA
44 PUBLICATIONS   449 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Vessels Fuel Consumption Forecast and Trim Optimisation: a Data Analytics Perspective View project

Hybrid propulsion system simulation and control View project

All content following this page was uploaded by Francesco Baldi on 28 June 2014.

The user has requested enhancement of the downloaded file.


THESIS FOR THE DEGREE OF LICENTIATE OF ENGINEERING

Improving ship energy efficiency through a systems


perspective

FRANCESCO BALDI

Department of Shipping and Marine Technology

CHALMERS UNIVERSITY OF TECHNOLOGY


Gothenburg, Sweden 2013
Improving ship energy efficiency through a systems perspective

FRANCESCO BALDI
Francesco.baldi@chalmers.se
+46 (0)31 77 22 615

© FRANCESCO BALDI, 2013

ISSN: 1652-9189
Technical report no 13:147

Department of Shipping and Marine Technology


Chalmers University of Technology
SE-412 96 Gothenburg
Sweden
Telephone + 46 (0)31-772 1000

Cover:
Artist impression of ship energy systems

Printed by Chalmers Reproservice


Gothenburg, Sweden 2013
Abstract

The last years have been particularly challenging for the shipping industry.
Fuel prices have increased to levels only seen during the oil crisis in the
70’s, and environmental regulations have grown much stricter than in the
past. Climate change, at a global level, is going to become a major threat
to society.
Increasing energy efficiency is one of the only possibilities of reducing fuel
costs and environmental impact of the shipping sector without influencing
the output. However, despite the recent developments in several aspects of
ship technology, little effort has been made in looking at the whole ship as
an energy system.
This licentiate thesis aims at filling a gap in the existing scientific knowledge
on the way energy in its different forms is generated, converted, and used
on board of a vessel. This is done by applying energy and exergy analysis
to ship energy system analysis. The results of this analysis allow improv-
ing the understanding of energy flows on board and identifying the main
inefficiencies and waste flows.
As a further development of this work, these results are used as a basis
for generation and evaluation of alternatives for improving ship energy ef-
ficiency. This is applied to the three main categories of: ship operations,
retrofitting, and design. Engine-propeller interaction, waste heat recovery
systems and the early stages of ship design are identified as relevant as-
pects and their evaluation indicates that there is a relevant potential of
improvement.
ii
Acknowledgements

The first version of my licentiate thesis was more than 100 pages long. After
discussion with my supervisors, it went down to roughly 60 pages. I will
therefore take what is my only chance to be lengthy for thanking those who
have had, in different ways, a role in the achievement of this key stone of
my education, and of my life.
First of all, thanks to my supervisors, Karin and Cecilia. It must not have
been easy, I acknowledge that, as my way of moving forward is not always
structured and methodical. This is why I need to thank you twice: once for
helping me, and once for standing me.
Big thanks also go to Bengt-Olof, Pär Brandholm, Mikael Karlsson, and in
general all the people at Laurin Maritime who contributed to the develop-
ment of my work providing me with data and support. I enjoyed very much
your expertise, your professionalism, and your cordiality in my regards.
Thanks also to Jon Agust, Kristinn and Stefn at Marorka. Working with
you has been motivating and I learnt a lot during my visits to Iceland. All
people I collaborated to, both in Laurin and Marorka, have been examples
of what a company should have for being successful in its business.
Thank you Gesa, thank you very much. You deserve your own little para-
graph, for all your support, ranging from work to personal life. If there is
one person I can say ”I would not be here without you”, that is definitely
you.
A lot of thanks to all my colleagues in the environmental group. Thanks to
Selma, for making a patient smile every time I came to your door just for
having a short chat. Thanks to Mathias, for your lessons on the Swedes, and
for coming to the cinema with me for watching trash movies. Thanks to,
Hannes for introducing me to Laurin and for the numerous confrontations
on energy efficiency, and research in general. Thanks to Philip for making
me feel you counted on me also outside of the working hours. Thanks to
Erik, for the all the afterworks we had together that reminded me there is
life out there. Thanks to Steven, Luis, Florian, Hedy, Nicole, Henrik, and
all those who make me feel, every time I wake up in the morning, that I am
happy to go to work because I am surrounded by wonderful people.
Thanks to my friends, for their support in the happy and the sad moments.
Thanks to (in random order, I hate alphabetic rankings!) Erry, Marco, Ste-
fano, Silvia, Ann, Alberto, Oana, Raquel, Jack, Pablo, Gabo, Bernadette,
Roberta, Ignacio, Saimir, Stella, Chiara, Dado, Karl, Josefin, Angela, Lu-
ana, Suny, Romolo, Bartolo, Loris, Marzi, Ale, Giaele, Filo, Fierro, ... I
am not thanking you individually for the only reason that would require a
whole thesis alone. But I cannot help giving a special thank to Maurizio,
who introduced me to sailing, the new little world where I just can feel good
when I need to escape from everything else.
Last, but definitely not least, thanks to Antonella and Sandro, papá e
mamma. I thanked all the people who had a role in making me the re-
searcher I am, but if I have to thank somebody for being the person I am,
that is you.
Appended papers

This thesis represents the combination of the research presented in the three
following appended papers:

Paper I : Baldi, F. , Gabrielii, C. & Anderssson, K. (2013) Energy and


exergy analysis of a ship: the case study of a chemical tanker, ready
for submission to Energy.
The study focuses on the analysis of ship energy system using en-
ergy and exergy analysis. These methods are applied to a specific
case-study. The results identify the main energy flows and the most
important inefficiencies on board of the selected vessel.
Paper II : Baldi, F. , Larsen, U. , Gabrielii, C. & Andersson, K. (2013)
A validated zero-dimensional four-stroke medium speed Diesel engine
model for waste heat recovery marine applications, submitted to Ap-
plied Energy on the 2013-06-10.
The study focuses on the description of a model for predicting medium-
speed marine Diesel engine performance and on its application to the
case of engine-propeller interaction modelling for a case-study ship.
Results underscore how the use of presented models can help identify-
ing more efficient operations.
Paper III : Baldi, F. , Bengtsson, S. & Andersson, K. (2013) The influ-
ence of propulsion system design on the carbon footprint of different
marine fuels, Low Carbon Shipping Conference, London, 9-10 Septem-
ber 2013.
The study focuses on proposing a method for comparing the carbon
footprint of different fuels and propulsion systems. A number of possi-
ble arrangements for the case-study ship were analysed and compared
in their energy performance and carbon footprint. The results show
how fuel and propulsion system choice in the design phase must be
tackled contemporarily and how a carbon footprint analysis can give a
much more detailed understanding of the environmental performance
of different propulsion systems and fuels.

For all the appended papers, the author of this thesis contributed to the
ideas presented and had a major role in planning the paper, data collection,
performing the analysis, and writing the manuscript.
iv
Contents

List of Illustrations vii


Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii
Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii

Symbols and abbreviations ix

1 Introduction 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Aim and research questions . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Delimitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.4 Thesis outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2 The ship as complex energy system 9


2.1 An introduction to complexity . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 The ship . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

3 The study of ship energy systems with a systems approach 17


3.1 The case study of a chemical tanker . . . . . . . . . . . . . . . . . . . . 18
3.2 The analysis of ship energy systems . . . . . . . . . . . . . . . . . . . . 22
3.3 The improvement of ship energy performance . . . . . . . . . . . . . . . 27

4 Ship energy system analysis 37


4.1 Energy analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.2 Exergy analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

5 Ship energy performance improvement 43


5.1 Ship operations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.2 Retrofitting and waste heat recovery . . . . . . . . . . . . . . . . . . . . 44
5.3 Ship design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

6 Discussion 49
6.1 Energy and exergy analysis . . . . . . . . . . . . . . . . . . . . . . . . . 49
6.2 Energy performance improvement . . . . . . . . . . . . . . . . . . . . . . 50
6.3 Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

v
CONTENTS

7 Future work and recommendations 57


7.1 Proposals to the scientific community . . . . . . . . . . . . . . . . . . . 57
7.2 Recommendations to the industry . . . . . . . . . . . . . . . . . . . . . 58

8 Conclusions 59

References 61

vi
List of Illustrations

Figures

1.1 World trade evolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2


1.2 Residual fuel price evolution and forecast . . . . . . . . . . . . . . . . . 3
1.3 Forecasted shipping contribution to global CO2 emissions . . . . . . . . 4

2.1 Schematic representation of ship energy system . . . . . . . . . . . . . . 15


2.2 Schmeatic representation of ship energy flows . . . . . . . . . . . . . . . 16

3.1 Graphical representation of thesis methodology . . . . . . . . . . . . . . 18


3.2 Conceptual representation of ship energy systems and flows . . . . . . . 20
3.3 Conceptual representation of the coherency of mechanical and thermal
energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.4 Conceptual representation of the property of exergy . . . . . . . . . . . 25
3.5 Conceptual representation of white and black box modeling . . . . . . . 29
3.6 Ship speed distribution over one year of operations for the case study ship 32
3.7 Propulsion system and individual engines load distribution over one year
of operations for the case study ship . . . . . . . . . . . . . . . . . . . . 32
3.8 Schematic representation of the two operational modes compared in the
study . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.9 Schematic representation of the propulsion arrangement with WHR sys-
tem installed: Case 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

4.1 Sankey diagram for the case study ship . . . . . . . . . . . . . . . . . . 39


4.2 Grassmann diagram for the case study ship . . . . . . . . . . . . . . . . 40

5.1 Comparison of case study ship specific fuel consumption, fixed-speed


versus variable-speed setup . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.2 Propeller power versus propeller speed, for different values of ship speed
and propeller pitch . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.3 Fractional coverage of the auxiliary need for different recovery arrange-
ments versus cycle exergy efficiency . . . . . . . . . . . . . . . . . . . . . 45
5.4 Comparison of case study ship specific consumption, standard setup ver-
sus WHR retrofitting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

vii
LIST OF ILLUSTRATIONS

5.5 Cumulative fossil energy consumption over one year of operation for the
alternative design cases. . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

Tables

2.1 Performance parameters of Diesel engines . . . . . . . . . . . . . . . . . 12

3.1 Available measurements from on board energy management system . . . 21


3.2 Comparison of mechanistic and empirical models . . . . . . . . . . . . . 29
3.3 Alternative arrangements evaluated in ship design . . . . . . . . . . . . 36

4.1 Energy flows and efficiencies for system components . . . . . . . . . . . 41


4.2 Exergy flows and efficiencies for system components . . . . . . . . . . . 42
4.3 Energy and exergy analysis of waste heat flows . . . . . . . . . . . . . . 42

viii
ex Exergy

Abbreviations
AE Auxiliary engine
CAC Charge air cooler
Symbols and CO2 Carbon dioxide

abbreviations CPP
ECA
Controllable pitch propeller
Emission controlled area
EEDI Energy Efficiency Design Index
EGE Exhaust gas economiser

Roman Symbols FPP Fixed pitch propeller

EX Exergy [kW ] GHG Greenhouse gas

H Enthalpy [kJ] HFO Heavy fuel oil

I Irreversibility [kJ] HHV Higher heating value

m Mass [kg] HT High Temperature

M CR Maximum continuous rate [kW ] HVAC Heat, ventilation, and air condition-
ing
P Power [kW ]
IMO International Maritime Organisation
Q Heat [kJ]
JW Jacket wataer
S Entropy [ kJ
K
]
g LHV Lower heating value
SF OC Specific fuel oil consumption [ kW h
]
kg LNG Liquified Natural Gas
SSF C Specific ship fuel consumption [ nm ]
LO Lubricating oil
T Temperature [K]
LT Low Temperature
W Work [kJ]
MDO Marine diesel oil
Greek Symbols ME Main engine
δ Contribution to total exergy destruc- NOX Nitrogen oxides
tion
PM Particulate matter
η Efficiency
S/G Shaft generator
λ Irreversibility ratio
SEEMP Ship Energy Efficiency Management
Plan
Subscripts
SOX Sulphur oxides
0 Reference ambient conditions
WHR Waste heat recovery
en Energy

ix
SYMBOLS AND ABBREVIATIONS

x
1

Introduction

The shipping industry is today facing very strong challenges. In a period of low freight
rates, fuel prices have increased to levels only seen during the oil crisis in the 70’s.
Stricter environmental regulations are putting additional stress on the sector. Mean-
while, the latest IPCC report highlighted the increased confidence in the existence of
an anthropic contribution to global warming. Shipping, though only contributing by an
estimated 3% to global CO2 emissions, is expected to increase its share in the future.
In such a context, it is not surprising that the interest in energy efficiency has
exponentially grown during last years. The critical role of shipping in global economy
implies that increasing the efficiency of the sector is one of the only ways to reduce
its consumption without decreasing its output. There is the need of addressing energy
efficiency in shipping from a number of different angles. This thesis approaches this
challenge from a technical perspective1 .
This chapter provides an introduction to the subject of the thesis. In Section 1.1 a
short description of the shipping sector and its current challenges is presented, leading
to the identification of the problem. The aim of the thesis and the research questions
are then explicitly defined in Section 1.2, while Section 1.3 draws the delimitations.
Finally, the thesis outline presented in Section 1.4 helps the reader in the orientation
through the different chapters and sections of this work.

1.1 Background
Shipping has always been intrinsically related to the history of mankind, from the
Pheonicians to the Romans, the Venetians, the Hanseatic League and the journeys of
Colombo, Diaz, Caboto, Zheng He. Today, shipping is one of the largest drives of
world’s globalised economy, as it contributes to more than 80% of global world trade
by volume, and 70% by value (UNCTAD, 2012). Figure 1.1 shows the evolution of
global trade in the last decades; even after the step back caused by the economic crisis
in 2008, global trade has already taken back on its previous pace, and most analysts
1
For a more interdisciplinary perspective, Towards understanding energy efficiency in shipping, by
Johnson et al. (2013), is a very informative reading.

1
1. INTRODUCTION

Figure 1.1: World trade evolution, 1950-2012; detail of the 2007-2012 period1

agree that this growing trend is very likely to continue in the future, fostered by the
growth in non-OECD countries (UN, 2013).
As a result of this trend, merchant shipping has been growing steadily over the
past years, hand in hand with world trade. In the period between 1999 and 2004
merchant shipping increased its economic turnover by a striking average of 22% per
year2 . This astonishing growth, together with the rising global economy, is explained by
phenomena like containerisation, increased economy of scale, and advances in marine
engineering. Under these conditions, the cost of freight is not a major concern anymore
when deciding where to purchase goods and materials (Stopford, 2009).

1.1.1 The evolution of bunker prices


The low cost of transport by sea has also been historically connected to very low prices
for marine fuels (normally referred as ”bunkers”). During latest years, however, the
increase in bunker prices has made fuel cost the largest element for virtually every
shipping company (DNV, 2012c). If as late as in the early 70s the fuel bill accounted
for around 13% of total ship costs, for the period between 2006 and 2008, fuel costs
were estimated to account for between 43% and 67% of total operating costs depending
on vessel type (Kalli et al., 2009). Figure 1.2 presents bunker prices evolution from
1984 until 2012, showing how they have seen a sharp increase since the 80’s and,
1
Author’s elaboration, from WTO (2012).
2
Source: Douglas-Westwood Ltd, available on (Stopford, 2009)

2
1.1 Background

'History'
800 'Forecast'
Bunker price (HFO) [USD/ton]

600

400

200

1984 1994 2004 2014 2024 2034


Time

Figure 1.2: HFO price historical evolution (1984-2012) and forecast (2012-2035)1

especially, in the last ten years. As it can be observed, even if prices fell around 2009
in coincidence with the economic crisis, they have already promptly recovered. In
the moment this thesis is going to print, the price for residual fuels ranged from 586
USD/ton in Rotterdam to 718 USD/ton in Sydney3 .
This is not the first period in history when oil prices (and, consequently, bunker
prices) have experienced this kind of increase. During the oil crisis of the 70s fuel costs
had risen to over 50% of ship operating costs (Buxton, 1985), creating the deepest re-
cession for the maritime sector since the Great Depression (Böme, 1983). Nevertheless,
in spite of the large number of studies connected to the reduction of fuel consumption
that was produced (Gauthey & DeTolla, 1974; DeTolla & Fleming, 1981; Brady, 1981;
Sack, 1981), most of the technical and managerial improvements discussed in those
years just faded as bunker prices dropped (Johnson et al., 2013). However, even if
there is disagreement among experts on the forecasts, reference scenarios hypothesised
by the major international agencies assume increasing prices in medium to far time
horizons (EIA, 2013; DNV, 2012b). This is a crucial matter for the subject of this
thesis, since fuel prices have a direct, strong impact on the uptake of new technologies
for increasing energy efficiency, as well as on the implementation of existing ones (DNV,
2012c).

3
Source:bunkerworld.com. Prices refer to IFO380, a type of residual fuel. Last updated 2013-10-20
1
Author’s elaboration from Brett (2008), Mazraati (2011), EPA (2008), Vivid & Economics (2010),
Sabinsky, SCC (2005), SSPA (2011)

3
1. INTRODUCTION

1.1.2 The influence of environmental concerns on fuel costs


Yet, this is only the economic part of the picture. In 2013 there could be no discussion
connected to the transport sector without a mention to greenhouse gases (GHG) and,
in general, emissions to air. Transportation by sea requires energy for propulsion, which
with today’s technological standard is provided by the combustion of fossil fuels. The
oxidation of carbon content in the fuel, in turn, releases carbon dioxide (CO2 ), which
stays in the atmosphere for centuries and contributes to global warming.
Shipping contribution to global CO2 emissions is relatively low and hard to evaluate.
Estimates relative to 2007 give an upper and lower boundary of respectively around 600
M T CO2
year (IEA, 2012) and around 1250 MyearT CO2
(Buhaug et al., 2009), which correspond
to a share of global CO2 emissions of between 2.7% and 3.6%. Taking into account
the contribution to the overall emissions of GHG, shipping is estimated to account for
1.2% to 2.5% of the total.
According to Rogelj et al. (2011), global GHG emissions need to be reduced to
GT CO2,eq GT CO2,eq
approximately 43 year by 2020 and 17 year by 2050 in order to have a 90%
likelihood of keeping the temperature from increasing more than 2◦ C compared to pre-
industrial levels. On the other hand, shipping emissions are not expected to decrease
at all. Even when the implementation of all cost efficient measures in a high carbon-tax
scenario is accounted for, projections do not forecast any reduction in total emissions
(Buhaug et al., 2009; Eide et al., 2011; Faber et al., 2009). As shown in Figure 1.3,
shipping might become the major contributor to global GHG emissions if present trends
are not diverted.
Two main policy instruments have been issued by the International Maritime Or-

Figure 1.3: Forecasted shipping contribution to global CO2 emissions. Lines represent
different scenarios for global CO2 emissions, while the gray area represents the possibility
space for shipping-related CO2 emissions as forecasted by Buhaug et al. (2009). From
Gilbert & Bows (2012)

4
1.2 Aim and research questions

ganisation (IMO)1 in the effort of reducing shipping impact on global warming: the
Energy Efficiency Design Index (EEDI), which sets minimum limits on the emissions
of CO2 per unit of transport work from newly built vessels, and the Ship Energy Ef-
ficiency Management Plan (SEEMP), which aims at improving awareness for energy
efficiency on existing vessels2 . Their effectiveness, however, has been put under question
(Johnson et al., 2013; Bazari & Longva, 2011; Devanney, 2011).
A part from GHG3 , sulphur oxides (SOX ) and nitrogen oxides (SOX ) are often
discussed as they have an influence on fuel costs. New, stringent limits on the emissions
of these two pollutants are expected to be enforced in the coming years. As reported
by the European Environmental Agency (EEA), shipping contribution to the national
SOX and N OX deposition is estimated to be between 10% and 30% of the total for
most of the European countries having a significant portion of their borders facing the
sea (EEA, 2013). Meeting the requirements imposed by new regulations on the matter
(especially in Emission Control Areas (ECAs), where limits are even more stringent)
will require either the installation of costly equipment on board, or the switch to cleaner
and more expensive fuels4 . In both cases, fuel-related costs are expected to increase in
the near future because of the more stringent requirements on emissions to air (DNV,
2012c).

1.1.3 Energy efficiency and the need for a systems approach


This thesis deals with the challenge of increasing ships energy efficiency, i.e. reducing
fuel consumption without decreasing the output. But if research and development
have reached very high standards in the technology of engines and propellers, the same
cannot be said of the design of ship energy systems. Until the recent past, low fuel prices
have generated very little demand for more energy efficient ships from the industry; as
a consequence, technical knowledge in this field has stagnated. There is the need of
approaching the subject with systems perspective, which regards the ship as a complex
system rather than focus on individual components (Lassesson & Andersson, 2009;
DNV, 2012a). This thesis aims at addressing this gap in the scientific knowledge in the
field.

1.2 Aim and research questions


The aim of this licentiate thesis is to contribute to a better understanding of the ship
as an energy system and to use this acquired knowledge in order to analyse possible
1
The processes inside the International Maritime Organisation (IMO), the body of the UN respon-
sible for international shipping, can be very complex. An analysis of one such example, related to
sulphur dioxides emissions from shipping, can be found in Svensson (2011).
2
EEDI and SEEMP have been adopted at MEPC 62 in July 2012 as an amendment to the Maritime
Pollution (MARPOL) convention.
3
There is still an open discussion, especially from a juridical point of view, on whether CO2 should
be considered a pollutant or not. For further details on the subject please refer to (Linné, 2012)
4
There is a large debate on what will be the marine fuels in the future. A thorough environmental
comparison of different marine fuels is presented in Bengtsson (2011).

5
1. INTRODUCTION

improvements to its energy efficiency.


To improve the understanding of ship energy systems translates in the objective
of evaluating components performances and energy flows sizes. The importance and
efficiency of producers, converters and consumers from one side and the size of inputs,
outputs and internal flows from the other should be evaluated. This can be translated
in the following research questions:
How can energy and exergy analysis be used to identify:

• identify the main energy producers and consumers on board? (Paper I)

• the main waste heat flows? (Paper I)

Once the main possibilities for improvement are identified, alternatives to the cur-
rent practice should be generated and evaluated. In general, three kinds of intervention
are considered: ship operations, retrofitting and design. In this thesis, this challenge is
addressed using a systems approach, focusing on the system as a whole rather than on
its individual components. This objective can be translated in the following research
questions:
How can the use of a systems perspective and of mathematical modelling assist in:

• the improvement of the energy efficiency of the propulsion system through a better
engine-propeller interaction? (Paper II)

• the evaluation of the feasibility of the installation of waste heat recovery systems
on ships? (Paper I, Paper II)

• the estimation and comparison of the energy consumption and the carbon foot-
print of alternative arrangements in the initial stage of ship design? (Paper III)

Results from the application of energy and exergy analysis are further developed in
Paper I in order to evaluate WHR feasibility. In Paper II, models for the main part
of ship energy systems are built in order to evaluate alternative operational modes
for propulsion, also including a WHR system. Finally, in Paper III alternative fuels
and propulsion system designs are evaluated in their annual energy consumption and
carbon footprint.

1.3 Delimitations
International shipping is the largest contributor both to the benefits (trade volume)
and to some of the drawbacks (GHG emissions) of shipping in general. This thesis will
hence focus on the types of vessels mostly used in international shipping. This kind
of ships are generally operated in stable conditions over long distances, which makes a
steady-state approach the most appropriate for tackling this subject. Dynamic effects
are, in most cases, of little interest and they are therefore not taken into account.
Furthermore, this thesis aims at providing tools and information on how to improve
the energy efficiency of ship energy systems, by means of generating and evaluating

6
1.4 Thesis outline

alternatives to current setups. The results of the thesis should, as a consequence, be


seen as a part of a larger picture, also including economic, environmental, and human
aspects. From an energy perspective, the analysis sets its boundaries at the ship as a
system, and again the results should be seen as part of a larger figure, where energy
needs for fuel production and transportation, and for ship production are also taken
into account.

1.4 Thesis outline


This thesis presents a synthesis of the research conducted over two and a half years on
the subject of improving ship energy efficiency through a systems perspective. Although
additional details can be found in the three appended papers, the thesis is supposed to
be understandable as a stand-alone work.
This thesis is subdivided in eight chapters; after the introduction to the back-
ground, subject and scope of the thesis, Chapter 2 describes the ship as an engineering
and energy system and motivates the use of a systems approach by introducing to
its complexity. Chapter 3 presents the main methodological framework of the work,
first describing the case study that have been used in the work, and then introducing
the reader to the concepts of systems analysis, energy analysis and system modeling.
Chapters 4 and 5 represent the core of the thesis, presenting the results of energy and
exergy analysis of ship systems and providing relevant examples of how these can be
used in order to generate and evaluate alternatives for the improvement of ship energy
efficiency. Results and methodology are then discussed in Chapter 6, while Chapters
7 and 8 respectively present suggestions for further research and recommendations to
the industry, and draws conclusions from the work and.

7
1. INTRODUCTION

8
2

The ship as complex energy


system

Concepts such as ”energy system” or ”systems approach” have already been mentioned
many times in this work. In particular, the aim of the thesis lies in the utilisation of a
systems approach in order to identify possible improvements of ship energy efficiency.
This is justified by the fact that in complex systems a major contributory factor [to
erroneous predictions of systems behavior] has been the unwitting adoption of piecemeal
thinking, which sees only parts and neglects to deal with the whole ˝Flood & Carson
(1993, p. 14). Inefficient design is often connected to erroneous predictions of sys-
tem behavior, which are normally originated by counter-intuitive behavior. However,
referring again to Flood & Carson (1993, p. 14),
this [counter-intuitive behavior] is not an intrinsic property of phenomena;
rather, it is largely caused by our neglect of, or lack or respect being paid
to, the nature and complexity that we are trying to represent. That is one
reason why we need systems thinking, methodologies, and models. We argue
that without this formal thinking we see only parts, the extremes, the simple
explanations or solutions.
This thesis addresses the subject of ship energy efficiency using a systems perspec-
tive. This kind of approach is most suitable, if not required, when a complex system
is to be understood and improved without risking sub-optimisation. Section 2.1 in-
troduces the subject of complexity and identifies what features of the ship make it
identifiable as a complex system. This leads to a description of the ship as an engineer-
ing system, and in particular of the features related to its energy efficiency, presented
in Section 2.2.

2.1 An introduction to complexity


Scientific literature can propose several examples which refer to the ship as a com-
plex system from the point of view of control theory, social sciences, environmental

9
2. THE SHIP AS COMPLEX ENERGY SYSTEM

sciences, and other more (Dupuis & Neilson, 1997). The interest of this work, how-
ever, lies specifically in looking at the ship as an energy system, and components and
sub-systems are considered under the light of their influence on the energy balance of
the ship. Components which are extremely relevant from other perspectives, as for
example navigation equipment, can be considered as a mere fixed, small contribution
to the overall energy balance of the ship.
The complexity of a system can be broken down to a number of attributes (Yates,
1978; Flood & Carson, 1993; Checkland, 1999):

• Large number of parts and significant interactions

• Non-linearity

• Emergence

Ship energy systems have a number of significantly interacting parts that is large
enough not allow intuitive prediction of all input-output relations, but small enough to
enable a clear identification of all components and of their interactions.
A system is non linear when at least one element in the system relates to and
varies in a non-linear way with another ˝ (Flood & Carson, 1993, p. 28). Linearity
allows very simple intuitive estimations. A ship is made of several complex components,
few of which can be described with a linear behavior. These components also give rise
to asymmetrical relationships.
Emergence refers to the property of the system to be more than the sum of its
parts. The ability of a ship to fulfil its mission depends on the coexistence of a number
of interacting components. A ship is more than the sum of an engine, and propeller, a
hull, and many other subsystems; it is the unique combination of all these components
that makes the system to be able to deliver its final function.
The ship, and in particular its energy systems, can therefore be said to be a complex
system which can take a benefit from the application of a systems approach. Before
proceeding further with the description of the work the led to the results of this thesis,
the most relevant features of a ship are introduced, with a special focus on those that
are most closely connected to the thematic of energy efficiency.

2.2 The ship


A ship is a floating, autonomously propelled platform which is designed for performing
a specific mission. As this generally implies moving the ship through water, propulsion
is often one of the major sources of energy consumption. The propulsion system fulfils
this function on board of the ship. On the other hand, a ship has additional func-
tional requirements, such as providing accommodation to the crew, supplying cooling
and lubrication, etc. These functions require additional power, in the form of electric,
mechanical, and thermal energy. A short overview of the most relevant auxiliary con-
sumers and of the most common technologies for the generation of auxiliary power and

10
2.2 The ship

heat is then presented. A visual representation of ship energy systems is provided in


Figures 2.1 and 2.21 .

2.2.1 The propulsion system


A detailed description of ship resistance would be out of the scope of this work; however,
for giving an estimation of how the system is influenced by external parameters, it is
often assumed that the power required for ship propulsion can be approximated as
shown in equation 2.1 (Woud & Stapersma, 2008, p. 52).

P = y · c0 (v)v 3 (2.1)

The factor y accounts for the influence of non-design conditions, such as hull fouling,
displacement, sea state, and water depth. The coefficient c0 , representing the charac-
teristic behavior of a specific ship, is normally increasing with speed, meaning that the
final dependance of ship propulsion power requirement with speed is not exactly a third
power curve (Woud & Stapersma, 2008, p. 52). The function of the propulsion system
is to provide the ship with the ability to move. Even if the propulsion arrangement
can vary substantially from vessel to vessel, the most common configuration can be
described by one or more prime movers, coupled to one or more propellers. This is the
typical arrangement for most of today’s commercial vessels.

Propellers
The propeller is the most widespread solution among for converting of the rotating
mechanical power from the engine shaft into a thrust force. Thrust bearings connect
the shaft to the ship, thus allowing to convert thrust force into ship motion.
Fixed pitch propellers (FPP) are characterized by having blades whose angle relative
to the axis of the shaft (pitch) is fixed. FPPs are generally directly connected to
low-speed two-stroke engines, therefore building a very solid, reliable, and efficient
propulsion train. However, this system suffers low flexibility and scarce manoeuvrability
(Molland, 2008). FPPs are the most widespread solution for ship propulsion, and are
particularly common among container ships, tankers, and bulk carriers (Carlton, 2007).
Controllable pitch propellers (CPP) allow the variation of the pitch. This ability
provides the CPP with an extra degree of freedom in addition to its rotational speed. As
a consequence, CPPs are installed for increasing ship manoeuvrability, for improving the
ability of adapting load to drive characteristic, and for giving the possibility to generate
constant-frequency electric power with a generator coupled to the main engines (Woud
& Stapersma, 2008). As a consequence of the system being more complex, CPPs are
more expensive and delicate than FPPs. CPPs are most favoured in passenger ships,
ferries, general cargo ships, tugs, and fishing vessels (Carlton, 2007).
1
Please note that the arrangement represented in these figure is only intended to give a feeling of
what are the main components positioned on board of the case study vessel. Their location in the
actual arrangement might differ substantially.

11
2. THE SHIP AS COMPLEX ENERGY SYSTEM

Prime movers
Diesel1 engines are the most widespread solution for the generation of mechanical power
from chemical energy. Firs installed on a ship in 1903, Diesel engine finally substituted
steam turbines in the 60s, and today constitute 96% of installed power on board of
merchant vessels larger than 100 gross tons (Eyring et al., 2010). Marine Diesel engines,
in fact, can achieve efficiencies up to 50%, allow operations to very low load (down to
10% of maximum continuous rating, MCR (Laerke, 2012)), and are designed to burn
both residual fuels (heavy fuel oil, HFO, and intermediate fuel oil, IFO), and distillates
(marine diesel oil, MDO, and marine gas oil, MGO) (Woud & Stapersma, 2008, p.
132). Recent developments also made dual-fuel engines available on the market, which
can run both on liquid fuels and on natural gas (Aesoy et al., 2011). The most relevant
features of Diesel engines are summarised in Table 2.12 .

Diesel Engines
Low-speed Medium-speed High-speed
Process 2-stroke 4-stroke 4-stroke
Construction Crosshead Trunk piston Trunk piston
Output power range [kW] 8000 - 80000 500 - 35000 500 - 9000
Output speed range [rpm] 80 - 300 300 - 1000 1000 - 3500
Fuel type HFO/MDO HFO/MDO MDO
SFOC [g/kWh] 160 - 180 170 - 210 200 - 220
Specific mass [kg/kW] 60 - 17 20 - 5 6 - 2.3

Table 2.1: Performance parameters of Diesel engines, state of art 2001 (Woud & Sta-
persma, 2008, p. 136)

Gas turbines are today the only alternative to Diesel engines for ship propulsion.
Despite being less efficient (efficiency for gas turbines ranges between 30% and 40%),
and less flexible with load and fuel quality (Woud & Stapersma, 2008, p. 138) than
Diesel engines, their main advantage lies in their higher power density. This makes
them suitable for applications where high power and low weight are required, as in the
case of fast ferries or naval vessels.

2.2.2 Auxiliaries
Ship auxiliary systems are a vital part of the ship. They are generally connected with
energy demands that can, depending on ship type, represent a significant portion of
1
As Diesel is the surname of the inventor of this type of engine, Rudolf Diesel, I will refer to Diesel
engines with capital letter.
2
For a detailed description of the different principles the reader is invited to refer to the extensive
literature on the subject, such as (Stone, 1999; Heywood, 1988; Kuiken, 2008).

12
2.2 The ship

overall ship energy use. As a ship at sea cannot use an external source of energy for its
auxiliary needs, these must be provided by on board machinery. This can be divided
as a function of the purpose (consumers, producers) or of the type of energy processed
(thermal, electrical).

Auxiliary consumers
A number of components on board require electric or mechanical power. Pumps are of-
ten a major consumer in this category, and they can be found in fuel and lubrication sys-
tems. Compressors are installed on board for air conditioning (HVAC, especially when
additional accommodation is required for passengers), refrigeration and compressed air
systems. Fans, both in the engine room and in cargo spaces, cargo handling, in the
form of pumps or cranes, lighting, especially for passenger vessels, and all navigation
equipment can also be a relevant source of electric power consumption. Ballast water
pumps also constitute large auxiliary consumers, as most ships need to load water into
specifically allocated tanks in order to maintain stability, especially when sailing with
empty holds1 . Most ships are also equipped with bow thrusters, propulsors located in
the fore of the ship, that provide additional manoeuvrability in port.
Cooling demand is met by using sea water as a cooling flow and is also associated to
relevant power demand, especially for pumps. The main cooling demand is represented
by the main engines, which is subdivided into cooling of the cylinder walls (normally
referred as jacket cooling, JW), of the charge air flow (charge air cooler, CAC), and of
the lubricating oil (lubricating oil cooler, LO). In general, a high temperature (HT) and
a low temperature (LT) circuits are used not to provoke too high thermal stress in the
components. Additional systems are installed when refrigeration or air conditioning
are required. The first is often needed by fishing vessels and reefers, while the second
is connected to accommodation facilities and is a major consumer in cruise ships and
ferries. Cooling demand generally translates in additional auxiliary power requirement
for the operation of cooling pumps and refrigeration systems.
Heat consumers can also be relevant to the overall energy balance of the ship,
especially for some specific vessel types, such as tankers and cruise ships. A large
heat demand is often connected to accommodation, both for heating and fresh water
generation. As HFO has very high viscosity at ambient temperature, fuel heating is
also often a major figure in this category. Finally, some vessel types (e.g. tankers) can
have specific mission-related heat demands, such as cargo heating.

Auxiliary producers
Electric power is often generated using auxiliary Diesel engines (AEs, also referred to as
auxiliary generators), coupled to electric generators and, in turn, to a main switchboard.
This solution is the most common when a 2-stroke engine is used for propulsion. When
a larger amount of auxiliary power is needed, a generator (in this case normally referred
1
A ship is said to be ”sailing ballast” when it is navigating with empty cargo holds for picking a
new cargo.

13
2. THE SHIP AS COMPLEX ENERGY SYSTEM

to as ”shaft generator”, S/G) can be coupled to the main engine shaft. This increases
the efficiency of auxiliary power generation since the main engines are generally more
efficient than the auxiliary engines. This solution requires however either a constant
engine speed (and, thus, a CPP), or the installation of power electronics for frequency
conversion, since the frequency of the current generated by the S/G directly depends
on the speed of the main engine (Woud & Stapersma, 2008).
Auxiliary heat is generated in different ways depending on the required quality
(temperature) and quantity. Heat exchangers recovering energy from the exhaust gas
(also referred to as exhaust gas economisers, EGE) are often employed when a relatively
small amount of process heat is required (i.e. only for fuel heating and accommodation,
which is the most usual case on merchant ships). A separate boiler is necessary for
higher heat demand, such as in the case of tankers and cruise ships. Heat is normally
distributed to different consumers either using steam or thermal oil. The freshwater
generator is a special case, as it can use low-grade heat and is therefore often located
on the high temperature cooling water systems.

Waste heat recovery systems


WHR systems refer to technical devices designed to make use of thermal energy that
would otherwise be wasted to the environment, a solution which is widely used in
various industrial sectors. The possibility of recovering waste heat from the main
engines exhaust gas to meet auxiliary heat demand has already been mentioned in
the previous section, and conceptually falls in this category. However the acronym
WHR will be used, in the continuation of this thesis, to identify systems whose main
purpose is to generate mechanical and/or electric power from a flow of waste heat. This
distinction is used since the technology required for the conversion of waste heat into
mechanical/electric power are conceptually different from what needed for the heat-to-
heat conversion.
Some different technologies exist for the conversion from thermal to mechanical
energy (Shu et al., 2013). However, this work focuses on the utilisation of systems
based on Rankine cycles. This technology has been particularly successful because of
its simplicity, safety, and relatively high efficiency (Tchanche et al., 2011). A Rankine
cycle is based on the generation of high-pressure steam and its subsequent expansion
in a turbine, which generates mechanical power. Organic Rankine cycles (ORC) are
often used when only low-temperature waste heat is available; their working process is
analogous to that of a standard Rankine cycle, but they make use of different working
fluids which allow additional freedom in the choice of the evaporating temperature.
On a ship, the main engines are the principal source of waste heat on board of most
of vessels, in particular through the exhaust gas and the cooling water flows. Despite
the application of WHR systems is still quite rare in shipping, they are considered to
have a high likelihood to be retrofitted on existent ships in the future (DNV, 2012c).

14
Accomodation
Fuel heating
Auxilary boilers

Exhaust gas Tank cleaning

15
economisers
Auxiliary engines
Engine room
auxiliaries N2
Bow thrusters

Propeller

Gearbox Main engines

Figure 2.1: Schematic representation of ship energy system


2.2 The ship
Mechanical energy consumer / producer

Heat consumer / producer


Exhaust gas
economisers
Accomodation Electric energy consumer / producer

Fuel heating
Auxilary boilers
HVAC Nitrogen
compressors
Tank cleaning

16
Auxiliary engines
Engine room
auxiliaries Bow thrusters

Propeller
2. THE SHIP AS COMPLEX ENERGY SYSTEM

Gearbox Main engines


Shaft generator
Cargo pumps

Figure 2.2: Schmeatic representation of ship energy flows


3

The study of ship energy systems


with a systems approach

In Chapter 2 ship energy systems have been described, and they have been shortly anal-
ysed from a systems science perspective. Having assessed the complexity of ship energy
systems, it is necessary to employ a systems perspective rather than a component-wise
optimisation. Systems analysis is a methodology suggested when dealing with problem-
solving in complex systems (Flood & Carson, 1993, p. 100).
The community of systems scientists only partially agrees on the specific procedures
to be applied in systems analysis (Checkland, 1999, p. 139), and it is common to adapt
the methodology to the specific need of the situation. In the case of this thesis, the
task of problem solving has to coexist with the more scientific need of improving the
understanding of a system.
The analysis of the problem constitutes the first part of a systems analysis. The
focus can be summarised in the question What are the limitations of the present system?
(Flood & Carson, 1993). Valuable information for answering to this question can be
provided by the results of energy and exergy analysis. A description of these methods,
together with some required thermodynamic background to the concept of exergy and
with an investigation of the existing literature, are provided in section 3.2.
The results from the energy and exergy analysis provide the basis for the following
steps of systems analysis: the generation and evaluation of alternative solutions. This
process requires an additional premise; in fact, during specific and in-depth studies,
conceptual or mental models are often not sufficient to cope with the type of complexity
involved. [...] It is, therefore, necessary to seek more formal structured approaches to
modeling (Flood & Carson, 1993). The second part of this thesis, hence, focuses on
the application of models to the evaluation of alternative solutions for improving the
system. This process is subdivided in its application to ship operations, retrofitting,
and design. More details on systems modeling and its application to ships in existing
literature are provided in Section 3.3.
The main interest of this study is to argue for the application of a systems perspec-
tive to ship energy systems. However, the presentation of such a method without any

17
3. THE STUDY OF SHIP ENERGY SYSTEMS WITH A SYSTEMS
APPROACH

Figure 3.1: Graphical representation of thesis methodology

reference to real cases would fall short both in explanatory power and persuasion. As
observed in most literature in the field, in the present work the proposed methods are
applied to a case study. The technical details of the selected case study ship and more
information about the data sources used in this study are here presented.

3.1 The case study of a chemical tanker


The case study represents an existing vessel, for which the company owning the ves-
sel (named the ”partner company”) provided extensive operational measurements and
technical information.

3.1.1 Description of the ship


The case study ship is a chemical/product tanker of 45 000 tons of deadweight which
largely operates in international waters and, often, in ECA areas. The propulsion
system is composed of two equally sized medium speed 4-stroke engines for a total
installed power of 7 680 kW. Both engines are connected to a common gearbox, which
in turn is connected to the propeller shaft, which provides the required thrust for
propulsion.
The ship is equipped with a S/G (rated 3200 kW), connected to the main gearbox,
which can provide auxiliary power. When operating in this configuration (”generator
mode”) both the main engines and the propeller need to be run at constant speed,
while acting on the pitch of the CPP enables setting the speed. Alternatively, the
ship can run in ”combinator mode”; this operational mode allows for variable propeller
speed, and consequently requires the use of at least on the two auxiliary engines (rated

18
3.1 The case study of a chemical tanker

682 kW each) for power generation. The ship is, however, operated in generator mode
during most of its operations.
When the main engines are running, the auxiliary heat is supplied by two EGEs,
capable of generating 700 kg/h of steam at 10 bar each, while two large auxiliary boilers,
rated 14 000 kg/h of steam at 14 bar, are used for peak demands and when the main
engines are not in operation.
For both electric power and heat, most auxiliary consumers are the same that can
typically be found on most merchant ships. Special functions connected to the ship
mission are the following:

Inert gas production and compression: Nitrogen needs to be produced on board


and pumped into cargo tanks when inflammable liquids are transported. Nitrogen
compressors have a high power demand, but are only operated intermittently.

Cargo pumping: When unloading the vessel, cargo pumps are required (high pressure
in the shore-based tanks is normally sufficient for cargo loading). They can require
a large amount of power when operated simultaneously.

Tank cleaning: After one cargo has been unloaded, tank cleaning is generally neces-
sary in order to prepare the cargo tanks for the following shipment. This operation
is performed either directly in port or during ballast trips, and requires the use
of the auxiliary boilers.

Cargo heating: Some specific liquids are characterized by very high viscosity at am-
bient temperature, which makes them unsuitable for handling. For this reason,
cargo heating can be ensured by means of process steam. This operation is,
however, very seldom required.

The energy system of the case study ship is schematically represented in Figure 3.2.

3.1.2 Input data


Measured data and technical information for the case study ship were provided by the
partner company. The different sources of quantitative and qualitative information are
hereafter described.

Continuous monitoring

The case study ship is equipped with an energy monitoring system which logs on board
measurements on a dedicated server with a frequency of acquisition ranging between 1
and 15 seconds. Data are automatically processed by the systems in order to produce
15 minutes averages, to check data reliability, and to filter output values. The list of
the measurements available on the energy monitoring system is presented in Table 3.1.

19
3. THE STUDY OF SHIP ENERGY SYSTEMS WITH A SYSTEMS
APPROACH

Mechanical energy

Electric energy

Thermal energy

Thermal auxiliary
Producers consumers
Converters
Engine
room

Auxiliary Nitrogen
boilers Compressor

Boiler
auxiliaries

Engine
room
Cargo
pumps
Fuel input Auxiliary Nitrogen
engines Switchboard Compressor
HVAC
Boiler
auxiliaries
Other
S/G consumers

Electric auxilairy
consumers
Main Gearbox Propeller
engines

Propulsion
EGE

Lub oil
cooler LT
cooling
Jacket
cooler
HT
cooling Cooling
CAC systems

Figure 3.2: Conceptual representation of ship energy systems and flows

20
3.1 The case study of a chemical tanker

Measured variable Unit


Ambient air
Dew point temperature ◦C

Relative humidity %
Auxiliary engines
ton
Fuel consumption 15mins
Power output kW
Shaft generator power output kW
Propeller
Power kW
Speed rpm
Torque kNm
ton
Main engines fuel consumption 15mins
Fuel temperature ◦C

Seawater temperature ◦C

Table 3.1: Available measurements from on board energy management system

Technical documentation
When direct measurements of ship and components performance are not available, they
can be calculated starting from available knowledge of the system. In this sense quite
extensive technical documentation was made available by the partner company for the
different components installed on board.

Main engines project guide contains information directly provided by the engine
manufacturer and publicly available online (MaK). Data here provided comply
with ISO 3046/1 and 15550 standards. Information connected to engine perfor-
mance, inlet and outlet flows, and thermal losses to the environment are used in
the study.

Main engine shop test contains experimental data provided by the manufacturer
and measured under well-defined conditions. Information on engine performance
for different load, including efficiency and exhaust temperature, is available from
this type of technical document.

Ship sea trials are performed when the construction of the ship is completed in order
to verify that the actual vessel performance conforms to the initial requirements
set by the customer. Ship sea trials were available for the case study ship and
for all its sister ships1 . On board sensors are generally used in this phase, but a
1
In shipping jargon, ”sister ships” are vessels built according to the same design

21
3. THE STUDY OF SHIP ENERGY SYSTEMS WITH A SYSTEMS
APPROACH

reasonable level of accuracy is guaranteed.

Propeller curves are represented as a diagram provided by the propeller manufac-


turer and generated through numerical codes. They provide information on pro-
peller performance for different values of propeller pitch, speed and power and
for different ship speeds.

Combinator diagram maps the characteristics of the control system installed on


board for engine-propeller interaction. The combinator diagram is used when the
ship is run in combinator mode, and is needed for engine protection versus too
high torque at low speed, which would result in excessive thermal loading for the
engine.

Ship electric balance is provided by the shipyard and summarises the expected
power consumption of different auxiliary components depending on ship oper-
ational mode.

Ship heat balance is supplied by the shipyard and provides details on the different
parameters used in the calculations such as heat exchange areas and heat transfer
coefficients.

Other technical documents provide design information to be used for the estima-
tion of the efficiency of the auxiliary engines, the shaft generator, and of other ship
components.

Noon reports and other data sources


On the case study ship, as on most vessels, information and measurements related to
on board fuel consumption and machinery relevant parameters is manually collected
daily by the crew and logged in paper and electronic format. Although the accuracy
and reliability of these data is often questioned (Aldous et al., 2013), they constitute
a broad source of knowledge and are used in this thesis when none of the previously
mentioned sources could provide the required information.

3.2 The analysis of ship energy systems


In this thesis, as energy efficiency is the main focus, the analysis of ship energy systems
is performed as first step of the systems analysis methodology. The analysis of ship
energy systems is subdivided in two main parts, energy and exergy analysis. While the
former is quite known in various fields of science and in the industry, exergy is only
seldom used, and will hence require some additional background.
The analysis of energy systems is often also divided between two main approaches
A top-down approach mostly relies on the analysis of extensive measurements carried
out in existing facilities (see, for example, Basurko et al. (2013)). On the other hand, a
bottom-up approach uses mechanistic knowledge of the system in order to simulate its

22
3.2 The analysis of ship energy systems

behavior and draw conclusions based on simulation results (see, for example, Nguyen
et al. (2013)). The choice of the type of analysis strongly depends on the availability
of input data and other information about the system.

3.2.1 Energy analysis


Energy analysis is defined as the process of determining the energy required directly and
indirectly to allow a system to produce a specified good or service(IFIAS, 1974), and
refers to the application of the first law of thermodynamics, which states the principle
of energy conservation (Clausius, 1850).
Energy analysis provides a quantitative insight of the relevance of different energy
flows. It can help to identify, for example, the largest consumers in a system, on which
efforts should be focused in order to improve the overall efficiency1 .
Studies reporting the analysis of the whole ship energy system are not common in
literature. The work of Thomas et al. (2010) and Basurko et al. (2013), related to fishing
vessels, propose an estimation of ship energy consumption and its repartition among
different consumers, followed by the generation of a number of different alternative
solutions for decreasing fuel consumption and their evaluation.

3.2.2 Exergy analysis


Energy analysis becomes incomplete and can be misleading when thermal energy flows
are compared to electric and mechanical flows; the first law of thermodynamics, in fact,
does not include any consideration about energy quality (Dincer & Rosen, 2013). The
concept of exergy can be of particular use in this case, as the exergy content of a flow
depends both on the quantity and on the quality of its energy content.

Heat and disorder


The need of introducing a new concept lies in the discrepancies between different energy
types and, especially, in the conversion from one to another. Electric energy can be
easily transformed into mechanical power and vice verse (electric motors have efficien-
cies of more than 90%, and the same holds for electric generators). Both these forms
of energy can also be easily converted into thermal energy (respectively, for instance,
through a fan and a resistance), with efficiencies close to 100%. The same cannot be
said of the opposite: in practical applications the conversion from thermal to mechani-
cal energy reaches maximum values of roughly 60% in advanced combined cycles. There
is always a part of the thermal energy input which cannot be converted into work. This
asymmetry is expressed by the second law of thermodynamics, which states that it is
not possible to have a cycle whose only result is to convert a given amount of heat into
work.
1
Using high efficiency led lamps can be a very good idea for an office building, where large amounts
of energy is required for lighting. On a container-ship, on the other hand, the same measure would
lead to much less rewarding improvements as overall consumption for lighting is almost negligible in
comparison to other needs.

23
3. THE STUDY OF SHIP ENERGY SYSTEMS WITH A SYSTEMS
APPROACH

Coherent No coherent
movement
movement

(a) Mechanical energy (b) Thermal energy

Figure 3.3: Conceptual representation of the coherency of mechanical and thermal energy

Explaining the reasons of this asymmetry requires to look into matter at microscopic
level. With reference to Figure 3.3, it is easy to observe that mechanical energy is
characterised by a coherent motion of particles, for which it is possible to identify a
principal direction. On the other hand, thermal energy is characterised by a completely
random motion, for which it is not possible to identify any main pattern. A cost is
associated to the passage from chaos to order, and the second law of thermodynamics
qualitatively describes this cost (Atkins, 1984).
A quantification of this cost would allow to compare thermal flows among each
other and versus more coherent forms of energy. Sadi Carnot provided the tools for
this quantification: the Carnot efficiency is defined as the maximum efficiency that
could be achieved by an ideal engine in generating mechanical power when receiving
heat from a thermal source at a temperature Th and rejecting the waste heat to a
thermal sink at a temperature Tc , lower than Th :
Tc
ηex = 1 − (3.1)
Th

Exergy

The concept of exergy derives from a generalisation of the Carnot efficiency. As con-
ceptually represented in Figure 3.4, exergy represents the fraction of a given energy
flow that could be converted into work using an ideal, Carnot engine. In fact, for a
given amount of matter, its thermal exergy content is defined as showed in Equation
3.2. The remaining fraction of the initial energy flow represents the part that cannot

24
3.2 The analysis of ship energy systems

Energy

Exergy

Anergy

Figure 3.4: Conceptual representation of the property of exergy

be converted into work, even in ideal conditions, and is called anergy.

EX = m[(h − h0 ) + T0 (s − s0 )] (3.2)

where EX, h, T and s respectively represent exergy, specific enthalpy, temperature and
specific entropy, while the subscript 0 represents reference conditions. It is possible
to demonstrate that, under certain assumptions, Equation 3.2 can be derived from
Equation 3.1.
For energy in a coherent form, such as in the case of mechanical, chemical, and
electric, energy and exergy flows coincide1 . Exergy and energy flows are instead re-
markably different in the case of thermal energy. As it appears from Equation 3.1,
the higher the temperature of a flow, the higher the fraction of that flow that can be
converted to work.
Equation 3.2 shows how the exergy content of a flow of thermal energy also depends
on the ambient temperature, which in general represents the cold sink of the hypothet-
ical Carnot cycle. A flow of steam at 500◦ C could not be converted into work if its
environment was at the same temperature2 .
1
In this thesis, differently from common practice in exergy analysis, the lower heating value (LHV)
is used instead of the higher heating value (HHV) for the quantification of chemical exergy. The reason
for this choice lies in the fact that as conditions of exhaust gas condensation are never reached in ship
energy systems, the use of HHV does not add any additional accuracy to the analysis.
2
Despite this argument, there is still debate on whether a fixed reference temperature should be

25
3. THE STUDY OF SHIP ENERGY SYSTEMS WITH A SYSTEMS
APPROACH

Exergy flows and efficiencies


According to Dincer & Rosen (2013) exergy flows calculated according to equation 3.2
can be divided in three main categories:
˙ in ) : the flow of exergy entering the component.
Input (EX
˙ out ) : the flow of exergy leaving the component.
Output (EX
˙ : the amount of exergy lost in the component operation (also
Irreversibility (I)
known as exergy distruction). This part represents energy quality deterioration
and is defined as I˙ = T0 Ṡgen , where Ṡgen represents the rate of entropy generation
in the component. In practical terms, however, exergy destruction is normally
calculated as the difference between input and output exergy flows.

There are several different figures of merit that are commonly used in exergy anal-
ysis. In this thesis, four different quantities are used:
˙
EX
p
Exergy efficiency ηex is defined for this study as ηex = EX ˙ in , where the subscripts
p and in respectively refer to products and inputs. This definition of exergy
efficiency gives an estimation of how efficient the component is in the generation
of useful products. In the case of heat exchangers, as there is not a clear distinction
∆EX˙ c
between inputs and products, the alternative definition of ηex = ∆ EX˙ h is used,
where subscripts c and h respectively refer to the cold and the hot fluid. The
definition of the ∆EX are adapted depending whether the component is meant
for cooling or heating.

Irreversibility ratio λ is used according to the definition proposed by Kotas (1980),



i.e. λ = EX˙ in . The irreversibility ratio gives an estimation of how much energy
quality is lost in the component.

Irreversibility share δ is defined as the ratio between the exergy destroyed in the
component and the total rate of exergy destruction in the whole system, i.e.
˙
δ = II˙i .
tot

Task efficiency ηtask is used in this thesis in a modified form from what proposed
in Dincer & Rosen (2013), and is defined as the ratio between the irreversibility
in an ideal exchange at constant temperature difference (here arbitrarily fixed to
˙
10◦ C ) and the irreversibility in the actual process, i.e. ηtask = IIid˙ . The task
efficiency gives an estimation of how close the component behavior is to an ideal
process.

Exergy analysis allows overcoming many of the shortcomings related to energy


analysis. The most typical and clear example is that of a heat exchanger: from a first-
law analysis, the component can be assumed to have a 100% efficiency, if the small
used instead for the evaluation of exergy flows (Pons, 2009).

26
3.3 The improvement of ship energy performance

heat losses to the environment are neglected, regardless the temperatures of inlet and
outlet flows. If a second-law analysis is performed instead, the exergy losses connected
to inefficient thermal exchanges (those, for instance, in which a low-temperature fluid is
heated using a very high-temperature fluid, as generally happens in households boilers)
can be identified. Exergy analysis provides, in general, an insight when both thermal
and mechanical/electrical energy flows are present in the systems (Dincer & Rosen,
2013).
Exergy analysis, however, is not widely used in shipping. Only few examples of
the application of this method to ships can be found in literature, and only focus on
individual components (Matuszak, 2008; Leo et al., 2010) or subsystems (Lijun et al.,
1996; Choi & Kim, 2012). In the only example of the application of exergy analysis
to the whole ship, Dimopoulos et al. (2012) presented the utilisation of exergy analysis
for the optimisation of a marine WHR system.

3.2.3 Procedure
In the application of the analysis of energy systems to the case study presented in this
thesis, a top-down approach is used, as extensive measurements are available. Input
data from the continuous monitoring system for propeller power demand, MEs and
AEs fuel consumption, S/G and AEs power output are used. The whole analysis of
energy and exergy flows represents an aggregation over one year of operation of the
case study ship.
Mechanistic knowledge of the system is used to break down the analysis of the energy
flows among different components. Data from the different sources presented in Section
3.1.2 are used to identify thermal flows in the exhaust gas, cooling systems, and in the
heat consumers. Data from the electric balance is elaborated to subdivide the overall
auxiliary power consumption among different users. Records from noon reports are used
to calculate the fuel input to auxiliary boilers. Heat exchange areas and coefficients
from the heat balance are used to estimate auxiliary heat consumption. Available
measurements of external seawater temperature are used as reference temperature for
the calculation of exergy flows.

3.3 The improvement of ship energy performance


As introduced in the beginning of this chapter, the results provided by the application
of energy and exergy analysis to ship energy systems should be used for the subse-
quent process of proposing and evaluating possible improvements to the system under
study. This thesis focuses on the application of such process to the three main types
of intervention in systems improvement: operational, retrofitting, and design.
In Section 2.1 the ship has been identified as a complex energy system. As suggested
by Flood & Carson (1993, p. 151) during specific and in-depth studies, conceptual or
mental models are often not sufficient to cope with the type of complexity involved.[...]
It is, therefore, necessary to seek more formal structured approaches to modeling. The

27
3. THE STUDY OF SHIP ENERGY SYSTEMS WITH A SYSTEMS
APPROACH

subject of system modeling is therefore hereafter introduced.

3.3.1 Systems modeling


Structured, mathematical modeling allows predicting the response of a system to differ-
ent inputs without the need of experimentation or prototyping (Blanchard & Fabrycky,
2006, p. 164). Depending on the problem to solve, different modeling approaches can be
used; a basic description of different criteria used when making decisions on approaches
and assumptions to be used in models is here provided, adapted from Grimmelius et al.
(2007) and Flood & Carson (1993, p.155):
Physical requirements: The model must be able to produce the outputs required
by its utilisation and must to reproduce the required physics. This for instance
might include the decision of whether the model should be dynamic or stationary,
or of which variables it is required to predict.
Accuracy requirements: Higher accuracy means better results. However, as re-
sources are limited, a target for model accuracy should be set in order to optimise
computational effort and final result. Accuracy requirements for a specific model
are normally related to the accuracy of other models in the same systems (in
order to avoid the ”bottleneck effect”), and to the accuracy of model inputs.
Data availability: A model should be constructed in accordance with the available
information on the system, both in terms of system parameters and system inputs.
Some parameters can be assumed based on previous work (physical constants,
empirical correlations), while in presence of experimental data it is also possible
to let some of the parameters vary in a calibration process, as long as enough
experimental points are available for performing both the calibration and the
validation of the model.
The main choice connected to the modeling approach refers to the use of mechanis-
tic (often referred to also as white-box, bottom-up, or deterministic) models, as opposed
to empirical (also referred to as black-box or top-down) models. Mechanistic models at-
tempt to describe the physical phenomena that characterise a system, thereby assuming
a deterministic approach; they typically make use of physical laws or empirical corre-
lations in order to model and predict the behavior of a system. In contrast, empirical
models treat the system to model as a black-box, and have no interest in the descrip-
tion of the underlying physical phenomena; starting from input-output databases, the
empirical modeller employes regression techniques in order to generate a model able
to predict system’s output. The difference between the two approaches is conceptually
visualised in Figure 3.5, while their characteristics are compared in Table 3.2.0
Finally, hybrid models (also known as gray-box or semi-empirical models) attempt
to mix the positive properties of both white- and black-box models.
1
The information summarised in this table is taken from (Duarte et al., 2004; Bieler et al., 2003,
2004; Braake et al., 1998; Groscurth et al., 1995; Bontempi et al., 2004; Grimmelius et al., 2007; Oliveira,
2004), which the reader is also referred to for further reading on various types of system modeling

28
3.3 The improvement of ship energy performance

(a) White box (b) Black box

Figure 3.5: Conceptual representation of white and black box modeling

Mechanistic models Empirical models


System knowledge required High Low
Validity of extrapolation High Low
Applicability in design phase High Low
Accuracy Low High
Improvement of system knowledge High Low
Amount of input data required Low High

Table 3.2: Comparison of mechanistic and empirical models1

29
3. THE STUDY OF SHIP ENERGY SYSTEMS WITH A SYSTEMS
APPROACH

3.3.2 Ship systems modeling


Application of energy system analysis in shipping dates as back as 1979 (Fake &
Pundyk, 1979). More recently, Shi et al. proposed models for predicting ship fuel
consumption in design and off-design conditions, both aiming at ships in general (Shi
et al., 2009, 2010) and in the specific case of dredgers1 (Shi & Grimmelius, 2010; Shi,
2013). The work of Shi (2013) focuses on the dynamic modeling of dredger energy
systems with the aim predicting energy consumption starting from the knowledge of a
limited number of external variables.
More focus on energy and fuel consumption while considering the ship as a system
was introduced by Dimopoulos et al., whose work first focused on LNG carriers (Di-
mopoulos & Frangopoulos, 2008a,b) and then more generally on marine energy systems
(Dimopoulos & Kakalis, 2010). In Dimopoulos et al. (2011) the use of optimisation al-
gorithms in order to optimise the design of marine energy systems, with a particular
focus on the implementation of WHR solutions and to the optimisation of related de-
sign parameters is proposed. The same authors also introduced exergy analysis applied
to ship energy systems (Dimopoulos et al., 2012).

Propulsion system modeling


The largest quantity of work in the field of ship systems modeling with an energy
perspective has been devoted to the propulsion system, as it constitutes the largest
share of ship energy consumption. Attempts in this sense started from the works of
DeTolla & Fleming (1981) in the US Navy during the oil crisis, and went to the tool
presented by Dupuis & Neilson (1997) and the modeling analysis described by Neilson
& Tarbet (1997).
Work by Benvenuto et al. (2005); Benvenuto & Figari (2011); Campora & Figari
(2003); Figari & Altosole (2007) focused on the propulsion system, proposing different
alternatives for the modeling of Diesel engines and gas turbines, ship dynamics, and
different control systems. Similar work was presented by Theotokatos (2007, 2008),
while Schulten (2005) focused on the interaction between the engines, the propeller and
the hull, with a particular focus of the dynamic events occurring during manoeuvring.
Grimmelius (2003); Grimmelius et al. (2007) also relates to modeling of ship propulsion
system, with a specific focus on the main engines. Tian et al. (2012) proposed and
validated a model for the prediction of the behavior of the propulsion system of a
RoRo vessel.
Some examples exist of the application of pure black-box models to ship propulsion
systems modeling. Among them, work from Leifsson et al. (2008) and Shi & Grimmelius
(2010) showed the comparison of the application of white-, grey-,and black-box models
to the prediction of ship performance. The results, in accordance to the theory, show
that black-box models provide the most effective prediction, but only when applied
inside the initial data range and when a large amount of input data is available. Other
1
A dredger is a vessel whose purpose is to excavate and remove material from the bottom of a body
of water.

30
3.3 The improvement of ship energy performance

attempts of predicting the influence of different control variables (speed, trim, etc.)
are mostly related to navigation, as proposed for instance by (Petersen et al., 2012).
In relation to the objectives of this thesis, however, black-box models fail to provide
additional insight on flows and phenomena in the energy system; furthermore, black-
box models are not suitable for the evaluation of alternative solutions, as they provide
only little possibility for extrapolation.

Auxiliary modeling
Studies related to ship auxiliaries are not common, as auxiliary power and heat demand
often constitute a negligible part of ship energy consumption. Some of the studies
mentioned in previous sections also include some part of the auxiliaries, such as in Shi
(2013) for the case of dredger pumps.
Among the few examples of auxiliary system modeling in literature, Balaji & Yaakob
(2012) analysed ship heat availability for use in ballast water thermal treatment tech-
nologies; others studied very specific types of auxiliary power consumption, such as
Fitzgerald et al. (2011) who focused on the consumption of refrigerated containers,
where Tilke et al. (2010) directed their interest on ship unloaders from bulk carriers.
Hulskotte & Denier van der Gon (2010) specifically studied ship consumption when
at berth, which is of particular relevance for the high impact of ship emissions when
released in densely populated areas (Winnes, 2010). Bidini et al. (2005) proposed in-
stead an analysis of the combined heat and power energy consumption of a small ferry
operating in lakes.
Work has been published on the design and optimisation of WHR applications
to ships already starting from the 70’s (Tarkir, 1979). Some studies, such as Tien
et al. (2007); Larsen et al. (2013), focused on a theoretical investigation of the WHR
cycle, while Ma et al. (2012); Grimmelius et al. (2010); Theotokatos & Livanos (2013);
Dimopoulos et al. (2011) proposed and evaluated different designs for the installation
of WHR systems on ships.

3.3.3 Ship operations


The first possibility for reducing ship energy consumption relates to the improvement
of ship operations. This alternative is of particular interest as it does not require
any installation of new equipment and, therefore, only limited investment. Ships are
normally optimised for one specific design point, while energy efficiency deteriorates in
off-design conditions. The design power for the case study ship refers to the condition
of sailing at a speed of approximately 15 kn in calm sea. Figures 3.6 and 3.7 show the
frequency distribution for ship speed, propulsion system load, and individual engines
load over one year of operations.
The figures show that the condition at design load (85% MCR) is far from being the
most frequent condition for ship operations, a consequence of the fact that the ship is
very seldom sailing at its design speed. In fact, it appears that the engines are mostly
operating between 40% and 70% load and that the ship mostly sails in the of range 8-12

31
3. THE STUDY OF SHIP ENERGY SYSTEMS WITH A SYSTEMS
APPROACH

0.20
Frequency of occurrence

0.15

0.10

0.05

0.00
0 5 10 15 20
Ship speed [kn]

Figure 3.6: Ship speed distribution over one year of operations for the case study ship

0.14 Individual engines


Propulsion system
Frequency of occurrence

0.12

0.10

0.08

0.06

0.04

0.02

0.00

0 20 40 60 80 100
Load [%]

Figure 3.7: Propulsion system and individual engines load distribution over one year of
operations for the case study ship

32
3.3 The improvement of ship energy performance

knots. Ship energy performance at low loads becomes therefore of primary importance.
This consideration, here deducted from the analysis of the case study ship, has also
been proved to apply to many other vessels operated today (Banks et al., 2013).
Propulsion is the main source of energy consumption on board of the ship and
should be addressed first. The systems approach suggests to look at interactions be-
tween components rather than at components themselves, which in the specific case is
corroborated by discussions with the partner company. The issue of engine-propeller
coupling, especially when the presence of a shaft generator requires to operate at con-
stant propeller speed, has also already been treated in literature (Woud & Stapersma,
2008; Van Beek & Van Der Steenhoven, 2005).
In this case, two alternative arrangements for ship propulsion are evaluated and
compared (see Figure 3.8):
Case 1 Fixed engine speed, auxiliary power provided by the shaft generator. This
arrangement corresponds to the standard operations in today’s settings.
Case 2 Variable engine speed, auxiliary power provided by the auxiliary engines. The
advantage in this case is the possibility of modifying propeller speed in order
to adapt to the conditions of best efficiency for different ship speeds. This case
represents the proposed alternative to be evaluated.

ME 1 ME 1

ME 2 ME 2

AE1 AE1

AE2 AE2

(a) Case 1 (b) Case 2

Figure 3.8: Schematic representation of the two operational modes compared in the study

As mentioned in the introduction to this section, the use of models rather than
prototypes or intuitive estimation is used to evaluate the different proposed alternatives.
In the case of the propeller, curves produced by the manufacturer displaying propeller
power as a function of ship speed, propeller pitch, and engine speed are available and
used in this study. In the case of the engine, however, more modeling effort is required.
In fact, all available information is related to engine operation at constant speed, which
does not reflect the operating conditions for Case 2. For this reason, an in-house engine
model is built in the Matlab© environment1 . The modeling of auxiliary engines and
1
A detailed description of the model is provided in Paper II.

33
3. THE STUDY OF SHIP ENERGY SYSTEMS WITH A SYSTEMS
APPROACH

of the shaft generator is based on simpler numerical regressions, widely used in this
field of engineering. All the different models are finally coupled and the behavior of the
system for the two different alternatives at different ship speed is evaluated.

3.3.4 Retrofitting
Retrofitting, i.e. the modification of an already existing vessel, is also considered as a
possible alternative for the improvement of ship energy performance.
Despite their high efficiency, Diesel engines reject to the environment a considerable
amount of energy, especially in the form of exhaust gas and cooling water. The imple-
mentation of WHR technologies is not new in shipping, even if still not common. The
utilisation of a systems perspective in the analysis of the feasibility of such systems is
hence employed. In particular, this thesis focuses on evaluating the recovery potential
compared to the actual need of auxiliary power, based on measurements relative to one
year of operations of the case study ship1 . A criteria was set for the evaluation of the
possible installation of such systems, as compared to standard operations: if installed,
it should provide the totality of ship auxiliary power need for a minimum of 80% of the
time spent sailing. The exergy efficiency of the WHR system required to achieve this
goal is then calculated for a number of different possible sources of waste heat.
The availability of waste heat is also influenced by engine load, and therefore by
ship operations. For this reason, in the second part of the study the focus is shifted
to the evaluation of the WHR potential as a function of ship speed. The same model-
ing approach as proposed for ship operations is used, where the potential for WHR is
evaluated using model output on exhaust gas temperature and mass flow. The arrange-
ment here proposed consists of positioning a WHR system on the exhaust gas flows of
the main engines (Case 3, see Figure 3.9). Case 3 is compared to both the previously
proposed Case 1 and Case 2.
In both parts of this study, the concept of exergy efficiency was used in order to
give an estimation of the technical complexity of the recovery system. This approach,
though only partially corresponding to real conditions, represents an improvement from
the use of energy efficiency, which does not allow to account for differences in energy
quality.

3.3.5 Design
As possible improvements both in operations and retrofitting are identified, the question
can be further moved to investigating how it is possible to improve the ship directly from
the design stage. Compared to common practice, the work presented in this thesis aims
at proposing a more holistic perspective, thus taking into account aspects such as fuel
selection, engine selection, WHR installation, energy performance and environmental
1
It should be noted that, as reviewed by Shu et al. (2013), several alternatives to power generation
exist for the exploitation of ship waste heat. However, given the ship needs of auxiliary heat are already
fulfilled by heat recovery and no additional requirement of, for instance, refrigeration was measured,
auxiliary power generation was the only possible utilisation of waste heat taken into account.

34
3.3 The improvement of ship energy performance

ME 1

ME 2

AE1

AE2

Figure 3.9: Schematic representation of the propulsion arrangement with WHR system
installed: Case 3

performance in the same study. In addition, the influence of the utilisation of a real
operational profile instead of a single design point, as suggested by Motley et al. (2012),
is evaluated.
The study is carried out by comparing a total of 11 different arrangements, as
showed in Table 3.3, which are generated assuming that the system should comply with
future emission limits in ECAs for SOX and NOX . The arrangements are generated
combining different alternatives of fuel type (HFO, MGO, or LNG), type of engine
(2-stroke and 4-stroke) and the utilisation of a WHR system1 . The performance of a
database of engines collected from technical documentation is evaluated using numerical
regressions. For each of the 11 proposed arrangements, the best performing engine was
selected. This allowed to compare ”best practice to best practice”. The analysis is then
completed by looking at how the picture changed when including a carbon footprint
analysis, thus including focusing on GHG emissions and employing an LCA approach.

1
The case of two-stroke engine with WHR using HFO as a fuel was not taken into account. This
case would require too complex arrangements in order to fulfil future ECA emission limits and therefore
was excluded from the study.

35
3. THE STUDY OF SHIP ENERGY SYSTEMS WITH A SYSTEMS
APPROACH

Case Fuel Engine WHR


1 HFO 2-st N
2 HFO 4-st N
3 HFO 4-st Y
4 MDO 2-st N
5 MDO 2-st Y
6 MDO 4-st N
7 MDO 4-st Y
8 LNG 2-st N
9 LNG 2-st Y
10 LNG 4-st N
11 LNG 4-st Y

Table 3.3: Alternative arrangements evaluated in ship design. 2-st and 4-st respectively
stand for two-stroke and four-stroke engines

36
4

Ship energy system analysis

The analysis of ship energy systems is applied to the case study ship. The objective
is to show the typical results of this application and the kind of information that it is
possible to extract. Energy and exergy analysis are applied to the case study ship.

4.1 Energy analysis


The results of the application of energy analysis to the case study ship are shown in
Figure 4.1 and in Table 4.1.
The first observable result is that consumption related to propulsion is the largest
figure, as expected, accounting for 70% of the overall energy consumption. This also
translates in the main engines consuming the largest share of the overall energy input
of the system, representing around 89% of the total.
However, the contribution from the auxiliaries cannot be neglected, particularly
because the ship spends a significant part of time waiting in port, when the only energy
demand comes from ship auxiliaries. Both auxiliary engines and boilers (respectively
representing 8.0% and 2.6% of ship energy input) on one side, and auxiliary power and
heat consumers (16% and 14% of ship energy output) on the other, should be given
significant attention. Boiler auxiliaries should be added to these considerations, as they
also represent a significant share of the total output (2.7%). Auxiliary boilers are also
run at low load most of the time, leading to low efficiency. Fuel heating also represents
a surprisingly high share of the overall ship energy consumption (7.8%).
Finally, a large amount of energy is wasted to the environment through the exhaust
gas (41% of main engines power output), the CAC (20%), JW cooler (22%) and the LO
cooler (24%). This suggests that there is a potential for the utilisation of these waste
flows for the generation of additional useful power.

4.2 Exergy analysis


Results of the application of exergy analysis to the case study ship are presented in
Figure 4.2 and in Table 4.2.

37
4. SHIP ENERGY SYSTEM ANALYSIS

The exergy analysis provides a different type of information on the system under
study. The irreversibility ratio (λ) quantifies the tendency of a component to deteriorate
energy quality in its internal processes. High values of λ correspond to high losses in
energy quality. It can be seen, for example, that according to this definition, boilers
(λ = 64%) are much less efficient than both main (37%) and auxiliary engines (38%).
In the case of heat exchangers, the utilisation of the irreversibility ratio as a figure
of merit can be misleading: a heat exchanger located on a large exergy flow (e.g.
on the exhaust gas) could have very high λ only because of its low heat exchange
area. This is the case, for example, of the EGE, which has a very low rate of exergy
destruction compared to the total exergy input (λ = 6.5%) and therefore appears to
be a very efficient component. If task efficiency (ηtask ) is used instead, it appears that
the EGE and all heat consumers are very badly designed exchangers, from an exergy
perspective, as they could achieve the same task with a much lower exergy destruction.
On the other hand the LT/SW exchanger, which is responsible of the highest rate of
exergy destruction among heat exchangers (δ = 5.3%), shows that this is connected to
its particular function; its task efficiency is not, in fact, particularly low (ηtask = 32%)
Figure 4.2 can be helpful in the evaluation of different waste flows with respect to
both energy quantity and quality. It is here shown that the exhaust gas is a much
larger source of potentially recoverable heat than the cooling systems, contrarily to
what could be deducted from Figure 4.1. When looking at the results of the exergy
analysis, exergy flow in the exhaust gas represent 54% of the total recoverable energy,
compared to 38% in energy analysis. The absolute estimation of the recoverable energy
is also redimensioned: the energy flow in the exhaust is estimated to be 41% of the
main engine output, while this value is decreased to 18% from an exergy perspective.
This reduction is even more pronounced when looking at the cooling systems. In this
case, it should also be noted that every step of heat exchange brings a decrease in the
recovery potential; as an example, the exergy flow entering the jacket cooling alone is
almost the same size as that flowing into the LT/SW heat exchanger. These results
show how exergy gives a much more realistic estimation of the actual power that could
be generated through a WHR system.
Even if smaller in size than that of the exhaust gas, the energy flow in the cooling
systems should not be discarded as it still constitute a relevant source of potentially
recoverable energy. Looking at task efficiencies leads to the identification of which
components could be improved in order to perform the same task while reducing exergy
destruction. This is particularly true for the charge air cooler (ηtask = 27.5%).
From the point of view of energy use, the very low efficiency of all auxiliary heat
consumers (tank cleaning, hotelling, and fuel heating respectively have task efficiencies
of 4.7%, 7.2% and 9.1%) indicates that it would be possible, by using a different heat
transfer fluid (or, in alternative, steam at a lower pressure), to generate the same heat
requirements while using much lower heat-grade sources. It is noted, for instance, that
fuel handling and hotelling only require temperatures as low as 70-80 ◦ C (a part from
fuel heaters before the engine, which warm HFO up to around 100◦ C ), which could be
provided at much lower temperature than by 9 bar steam.

38
4.2 Exergy analysis

Figure 4.1: Sankey diagram for the case study ship

39
4. SHIP ENERGY SYSTEM ANALYSIS

Figure 4.2: Grassmann diagram for the case study ship

40
4.2 Exergy analysis

Energy analysis
%inp %out ηen
Producers
Main engines 88.4 42
Auxiliary engines 8.0 36
Boilers 2.6 81
Consumers
Propeller 70.0
Nitrogen compressors 2.1
Cargo pumps 0.8
Boiler auxiliaries 2.7
Engine room 3.5
HVAC 1.8
Miscellaneous 2.6
Tank cleaning 3.1
Hotelling 5.5
Fuel heating 7.8
Internal flows
Gearbox 37.0 99
Shaft generator 4.0 91
Turbocharger 34.1 26
EGE 3.2
Exhaust (after EGE) 22.0
Charge air cooler 7.6
Lub oil cooler 9.0
Jacket water cooler 8.4
HT/LT heat exchanger 14.0
LT/SW heat exchanger 28.0

Table 4.1: Energy flows and efficiencies for system components

41
4. SHIP ENERGY SYSTEM ANALYSIS

Exergy analysis
ηex λ δ ηtask
Main engines
Engine 42.9% 37.2 % 64.2 % -
Turbocharger 35.9% 24.4% 6.4% -
Exhaust gas economiser 67.0% 6.5% 1.1% 4.8%
Charge air cooler 66.4% 6.8% 1.1% 27.5%
Lubricating oil cooler 59.3% 20.1% 1.1% 38.8%
Jacket water cooler 50.1% 15.9% 2.5% 49.5%
HT/LT heat exchanger 59.5% 11.6% 1.8% 38.5%
LT/SW heat exchanger 2.1% 85.3% 5.3% 31.5%
Auxiliary power
Auxiliary engines 61.7% 38.3% 5.3% -
Auxiliary heat
Boilers 29.7% 63.7% 5.4% -
Tank cleaning 25.3% 61.1% 0.8% 4.7%
Hotelling 37.8% 46.7% 0.6% 7.2%
Fuel heating 26.2% 62.1% 1.8% 9.1%

Table 4.2: Exergy flows and efficiencies for system components

Energy Exergy
%rec %M E,out %rec %M E,out
Exhaust gas 38 41 54 17.7
Charge air cooler 19 20 14 4.6
Jacket water cooler 21 22 20 6.7
Lubricating oil cooler 22 24 12 3.8

Table 4.3: Energy and exergy analysis of waste heat flows

42
5

Ship energy performance


improvement

Departing from the results of the first part of the thesis, the newly acquired knowledge
of the system can be used in order to improve its performance from an energy perspec-
tive. Measures of intervention are generally subdivided in the categories of operational,
retrofitting, and design, which are presented in Sections 5.1 to 5.3.

5.1 Ship operations


Figure 5.1 presents the result of the comparison of Case 1 and Case 2 operational modes.
The specific ship fuel consumption (SSFC), defined as the amount of fuel consumed by
the both the main and the auxiliary engines over one nautical mile distance (see Paper
II, equations 15 and 16), is plotted versus ship speed. The dashed and the solid lines
respectively represent the SSFC of the Case 1 and Case 2 arrangements.
The results indicate that a large reduction in SSFC can be obtained by modifying
the operating mode from fixed to variable propeller speed. The reason lies in the
very different efficiency of the propeller, at given ship speed, depending on propeller
speed, as shown in Figure 5.2. The propeller is designed for a ship speed of 15 knots,
where its best efficiency point corresponds to a rotational speed of 105 rpm. At lower
ship speeds, efficient propeller operations would require a slower rotation. Operations
in combinator mode do not allow reducing engine speed when only running on one,
high-loaded engine. As soon as operating on two engines is permitted, the reduction
of propeller speed brings an improvement of ship energy efficiency, as can be seen in
Figure 5.1 in the range between 11 and 13 kn. This advantage peaks at around 12 kn
and then diminishes when increasing ship speed. At around 14 knots, close to design
conditions, the performance of Case 1 becomes again more efficient than that of Case
2, as expected. With reference to Figure 5.1, it is possible to observe the moment
when the second main engine is clutched in, which in both cases corresponds to a sharp
increase in SSFC.

43
5. SHIP ENERGY PERFORMANCE IMPROVEMENT

80
Specific ship consumption [kg/nm]

60

40
Specific ship consumption [kg/nm]

80 20

Case 1: Fixed rpm


Case 2: Variable rpm
0
78 10 11 12 13 14
Ship speed [kn]

76

74

72

70

68
Case 1: Fixed rpm
66 Case 2: Variable rpm

10 11 12 13 14
Ship speed [kn]
Figure 5.1: Comparison of case study ship specific fuel consumption, fixed-speed versus
variable-speed setup

5.2 Retrofitting and waste heat recovery


The results of the first part of the analysis of WHR potential for the case study ship
are presented in Figure 5.3; for each value of WHR cycle exergy efficiency (X-axis),
the correspondent generated power is compared to the actual ship demand in auxiliary
power. The percentage of time during which the WHR system is able to generate the
required amount of auxiliary power is calculated (Y-axis). This process is repeated for
different choices on which waste heat flow is used as an energy source for the WHR
system. For each of these possibilities, the exergy efficiency of the WHR system required
for meeting auxiliary power demand for at least 80% of the time is shown.
When recovering on the exhaust gas alone the required efficiency amounts to ap-
proximately 58% if the WHR system is installed before the EGE (continuous line,
circular marks), while this value increases up to around 62% if the easier retrofitting
arrangement of installing the WHR after the EGE was employed (dashed line, trian-
gular marks). The required efficiency of the recovery cycle can be drastically reduced
by taking the cooling systems into account: either for the generation of auxiliary heat
(chained line, square marks) where 50% efficiency is required, and where all the energy
in the exhaust gas is available for WHR purposes; or for the WHR system itself (dashed
line, plus marks, and dotted line, rhombus marks) where 48% efficiency is required. It
should be noted however that these solutions would imply a shift in system complexity,

44
5.2 Retrofitting and waste heat recovery

Figure 5.2: Propeller power versus propeller speed, for different values of ship speed and
propeller pitch; kindly provided by the partner company

100 Only Exhaust, EGE before WHR boiler


Only exhaust
Only exhaust, no aux heat generation
Exhaust + LT cooling
Fulfillment of auxiliary power need [%]

Exhaust + CAC
80

60

40

20

0
30 40 50 60 70
WHR system exergy efficiency [%]

Figure 5.3: Fractional coverage of the auxiliary need for different recovery arrangements
versus cycle exergy efficiency

45
5. SHIP ENERGY PERFORMANCE IMPROVEMENT

80
Case 1: Fixed rpm
Specific ship consumption [kg/nm]

Case 2: Variable rpm


Case 3: WHR
75

70

65

60

10 11 12 13 14
Ship speed [kn]
Figure 5.4: Comparison of case study ship specific consumption, standard setup versus
WHR retrofitting

from the efficiency of the WHR cycle to the equipment required for recovering waste
heat from the cooling systems.
The possibility of adapting ship operations towards the optimisation of the effi-
ciency of the whole system (including WHR) was also explored. The results indicate
that the installation of a WHR system on the case study ship becomes easier if both
engines are operated simultaneously. In this case, in fact, the advantages of permitting
low-propeller speed operations and having additional exhaust flow neutralise the dis-
advantage of operating at low load. This is graphically shown in Figure 5.4, where the
SSFC of the propulsion system with WHR (Case 3) is compared to operations in the
current arrangement of the case study ship (Case 1 and Case 2, see Section 5.1). For
operating speeds high enough to allow the utilisation of two engines (around 11 kn),
the calculated minimum exergy efficiency required in order to provide auxiliary power
(364 kW) on board was estimated at 39% (auxiliary needs are lower or equal to 364
kW for 80% of the sailing time). The utilisation of a WHR system for auxiliary power
production also allows operating the engine in combinator mode, as presented in the
previous section, hence contributing to the large improvement displayed in figure 5.4.

5.3 Ship design


The results related to the application of a method for the comparison of different
propulsion arrangements, presented in Section 3.3.5, are shown in Figure 5.5. Different

46
5.3 Ship design

fuels perform in dissimilar manners depending on the selected propulsion system ar-
rangement, and vice verse. This distinction is particularly strong when WHR is taken
into account. It is therefore considered important to include the choice of both the
propulsion system and the fuel in the early phases of ship design.
Additionally, the study aims at evaluating the possible advantages of selecting the
propulsion system based on the whole operational cycle rather than on the design point.
The results show that there is no difference in the two approaches for 2-stroke engines,
which means that employing an engine selected based on its high efficiency at design
point also leads to the most efficient solution when the whole operational cycle is taken
into account. For 4-stroke engines instead a small improvement is computed when
using the ”operational-cycle approach”, evaluated in a decrease in fuel consumption
of 0.56% for the cases of HFO and MGO and 1.7% in the case of LNG. This latter
result is of particular interest and is connected to the fact that 4-stroke LNG-powered
engines operate according to a Otto cycle, which is known to have worse performance
at low-load than a comparable Diesel cycle-based engine. The use of the ”operational-
cycle approach” is therefore particularly advised when the choice of LNG as a fuel is
associated to a propulsion arrangement where 4-stroke engines are the prime mover,
which is becoming quite a common choice for ferries operated in ECAs (Aesoy et al.,
2011).

2
Note that in the LNG case pilot fuel injection is also taken into account

47
5. SHIP ENERGY PERFORMANCE IMPROVEMENT

0 2 HFO 4 MGO LNG


6 8Reference case10 12

150
Yearly energy consumption [TJ]

100

50

0
2-stroke 4-stroke 2-stroke 4-stroke 2-stroke 4-stroke
Propulsion arrangement

Figure 5.5: Cumulative fossil energy consumption over one year of operation for the al-
ternative design cases. Each group of bars represents the main fuel (HFO, MGO, LNG2 ).
In each group of bars, the first two represent the performance of a two-stroke engine ar-
rangement, while the last two that of a four-stroke engine arrangement. Finally, for each
pair, the first bar represent the yearly consumption of the vessel without WHR, while the
second also accounts for such system to be installed. The horizontal line gives the calculated
consumption of the existing ship arrangement. Finally, the error bars represent the extent
of the range of results for all the evaluated arrangements for each specific case. As the
main bars represent the most efficient engine choice for each arrangement, error bars only
stretch above the main level

48
6

Discussion

In Chapter 6 the implications connected to the results of this thesis are discussed,
together with the validity and generality of the conclusions that can be deducted (Sec-
tions6.1 and 6.2). A discussion related to the methodology employed in the study is
proposed in Section 6.3, referring to the validity of the thesis and to the quality of the
data.

6.1 Energy and exergy analysis


The application of energy and exergy analysis to the case study ship shows how these
methods can be used in order to improve the understanding of ship energy flows. These
methods can guide in the identification of main energy flows, therefore getting an
understanding of where intervention should be prioritised, and main inefficiencies, which
in turn can foster the ability of understanding where the potential for improvement is
located.
In this thesis, energy analysis led to results conceptually similar to what presented
in energy audits available in literature, such as Thomas et al. (2010) and Basurko
et al. (2013), which also include the identification of energy flows. The energy anal-
ysis proposed by the author, however, brings further detail in the analysis of internal
energy flows in systems such as the turbocharger, the cooling systems, the heat dis-
tribution systems and auxiliary power generators. This level of detail, to the best of
my knowledge, has never been presented in literature before and provides an improved
understanding of the relative and absolute sizes of different flows of different energy
types on board the ship.
The analysis of energy consumption led to the unexpected result of showing that
thermal (14% of total energy use) and electric (16%) users make up a large part of the
total ship consumption. Even if this result underscores the fact that the propulsion
line should be the first priority when approaching the issue of reducing ship fuel con-
sumption, it also shows that optimising ship energy systems performance based on the
propulsion power demand alone is not justified.
One additional contribution from this work is give by the introduction of exergy

49
6. DISCUSSION

analysis to the whole ship, which provided the basis for approaching the possibility of
implementing WHR systems on the ship as a way of reducing fuel consumption need
for auxiliary power generation. The estimation of the different energy flows including
considerations on energy quality allowed to select which flows could be of major interest
for energy recovery way that would not be allowed by energy analysis.
Additionally, exergy analysis allowed the identification of those components which
could be targeted in order to improve the overall potential for WHR. However, from a
technical point of view, improving the task efficiency of a heat exchanger is achieved by
reducing the temperature difference across the component. This reflects into an increase
in the exchange area required for maintaining the same amount of heat exchange,
which de facto translates into additional space and weight requirements. Furthermore,
improvements in the design of the network of heat exchangers could only be beneficial
if additional use of waste heat is planned. In the current arrangement all heat needs,
when the main engines are running, are met by the EGE and there is no need for an
optimisation.

6.2 Energy performance improvement


The results from the application of energy and exergy analysis to ship systems can be
used as the basis to ship energy performance improvement. The attention focused on
engine-propeller interaction, WHR systems, and whole propulsion system design.
It should be noted that the results presented in Chapter 5 only attempt to see
the matter from an energy perspective. Details regarding other aspects, such as the
economic (balance between savings and additional costs for installation, training, con-
tingent increased maintenance), human (increased workload in the engine room, lack of
expertise, safety), structural (additional space/weight requirements), and environmen-
tal (use of dangerous chemicals) were either not included, or considered qualitatively.
The results presented in this thesis should be seen as information to be put in a larger
context by the interested stakeholders when a decision is to be taken on how to operate,
retrofit, or build a ship.
The investigation of engine-propeller interaction led to the identification of possibil-
ities for improvement in this area for the case study ship. The results obtained through
the application of mathematical models to the interaction between engine and propeller
in the case study ship suggested that a reduction in ship fuel consumption of 5.8% can
be achieved while increasing ship speed from 10 to 12 kn if the ship is run in variable
propeller speed. Even though the outcome could vary for different ship types, the re-
sults suggest that this analysis can lead to the identification of possible improvements
of ship energy efficiency. Furthermore, although in this thesis the interaction between
engine and propeller was studied from an operational perspective, efforts can also be
directed to the retrofitting of existing ships and to the design of new vessels.
WHR systems are often suggested as one possible solution for decreasing ship energy
consumption, but the actual feasibility is seldom analysed in all its complexity. In this
sense, the main limitation in the approach proposed by the cited authors lies in the

50
6.2 Energy performance improvement

accounting of the operational mode of the selected ship. Ma et al. (2012) assume
that the system is operated at design conditions for 280 days a year, an assumption
that rarely reflects real operational conditions (Banks et al., 2013). Grimmelius et al.
(2010) and Theotokatos & Livanos (2013) both take into account a typical voyage and
quantitatively use this assumption for the calculation of expected system performance.
The ”typical voyage” approach represents a widely accepted approximation, but is
only partially able to account for the impact of real operations on fluctuations in ship
operational pattern (typical of ships operating on the spot market) and in boundary
conditions (increased resistance due to weather and waves). In the work presented by
Dimopoulos et al. (2011) ship operations are divided in 4 well-defined categories with
an assumption for the time spent during each operation, which additionally increases
the detail, but is more suitable for the operational pattern typical of container ships.
The main contribution of this thesis to the subject consists in the evaluation of the
WHR potential over one year of real ship operations. The results indicate a profitable
application of WHR to the case study ship would require a rather complex recovery
system, either from the point of view of the recovery cycle (efficiencies of around 60%
are required, which could only be met by advanced Organic Rankine Cycles (ORC) with
high-performance fluids (Larsen et al., 2013)), or from that of the type and amount of
waste heat to be recovered (by making use of the waste heat in the cooling system the
required efficiency can be reduced to 48%1 ).
The possibility of adapting ship operations in order to maximise the energy efficiency
of the propulsion system including WHR was also explored. The results suggest that
by adapting operations for running the ship on two engines, even at low load, the
advantages from improved propeller and WHR performance would outweigh the loss
in engine efficiency, while requiring a relatively simple recovery system (16% reduction
of fuel consumption with a required ηex of 39% recovering waste heat on the exhaust
gas alone). This result indicates that the installation of WHR systems should be
evaluated taking also operational aspects into account. This results is considered to
be of particular relevance as it underscores the importance of addressing ship energy
systems with a systems approach. For the case study ship, both the improved engine-
propeller interaction and the installation of WHR systems, if taken alone, would provide
much smaller benefits than the two applied together. Looking at ship energy systems
as a whole allowed the identification of this synergy, something that would not have
been possible otherwise.
A similar consideration can be done on possible energy savings on auxiliary power
consumers. Small reductions in auxiliary electric demand can make the difference
on the feasibility of a WHR system. The availability of measured data on auxiliary
consumption would allow further research in this direction.
Finally, the possibility of influencing the system from the design stage was explored.
Even if results in the initial phases of the design are only partly representative of
1
It should be noted that the use of cooling systems as a source of waste heat is not technically
easy. Cooling is vital to engine operations, and introducing an additional system could generate issues
in safety and control.

51
6. DISCUSSION

what the real systems will look like, the methods proposed in this thesis show very
promising results. Different engines, arrangements, and fuels can be compared and the
resulting estimated yearly consumption can be used as one of the basis for the decision of
which propulsive arrangement should be installed. The importance of including several
different parameters in the early design phase is demonstrated by the large variance in
the results depending on the chosen arrangement. Finally, the influence of including
the operational cycle in the evaluation proved rather limited, and only relevant in the
case of four-stroke dual fuel engines based on an Otto-cycle.

6.3 Methodology
As part of the analysis of the scientific value of this thesis, the quality of the work should
be critically discussed. Validity is often referred to as a measure of the quality of a
scientific work. It is often subdivided in internal validity, which refers to the extent
to which the results represent the real problem to be studied, and external validity,
which concerns instead the generalisability of the results to a larger sample than what
specifically addressed in the work. Finally, the quality of the data also has a very
important influence on the value of the results.

6.3.1 External validity and case studies


External validity deals with the extent to which the results and the methods presented
in a scientific study can be extended outside of the specific application presented in
the work (Mitchell & Jolley, 2001). In the case of this thesis, external validity is a
subject of major importance, as the use of case studies is often connected to a loss
generalisability.
The methods proposed in this thesis are potentially applicable to all ship types. The
main restriction to the external validity of the methods here presented is the assumption
of steady-state behavior, as all models and analysis are conceived for steady state
operation. This requirement implies that vessel types characterised by a very dynamic
behavior, such as tugs and inland ferries, would not be suitable for the application
of the proposed methods. Other possible differences, such as a different engine type
(two-stroke Diesel engine, gas turbine), different propeller type (FPP, water jet), and
propulsion arrangement would just require an additional effort in the modeling of a
different component.
On the other hand, the external validity of the numerical results is more limited. For
what concerns the analysis of ship energy systems, it has been observed that time spent
in different operational modes tends to remain approximately constant over different
years (Banks et al., 2013). The possibility of extending the results of this thesis to
future operations of the case study, is correlated to vessel speed distribution, as it has
a strong influence on fuel consumption for propulsion, and can be subject to large
variations over the years (Banks et al., 2013). When it comes to the extension to other
ship types, the possibility of extending the results presented in this thesis is limited by:

52
6.3 Methodology

Engine type: Two-stroke engines have lower exhaust temperatures (Theotokatos &
Livanos, 2013), therefore revolting all conclusion drawn in connection to the eval-
uation of WHR potential.

Propeller type: FPPs behave substantially differently from CPPs, and therefore all
results related to engine-propeller interaction would differ if a FPP is used instead.

Auxiliary power generation arrangement: Results related to engine-propeller in-


teraction are connected to the use of a shaft generator.

Ship type: All results, especially those connected to the evaluation of WHR potential,
are strictly related to the balance between ship requirement of propulsion power,
auxiliary heat, and auxiliary power. Different ship types might differ sensibly in
this matter.

These restrictions, should not lead the reader to think that the arrangement featured
by the case study ship is a special one. Four-stroke engines represent a significant
portion of the market (27.7% in terms of number of installed engines and 10.5% in terms
of installed power for engines with MCR > 2 MW (Haight, 2012)). Similarly, CPPs
have reached a stable market share of around 35% of the total number of propellers
installed; this is particularly true for some specific ship types, such as general cargo
vessels (80%), ferries (63%), tugs and offshore vessels (78%), and fishing vessels (89%)1 .
The utilisation of case studies is very widespread among applied researchers, es-
pecially when methods are proposed rather than specific designs. This is the case
for almost all work mentioned or referred to in this thesis: (Dimopoulos et al., 2011,
2012) refer to a specific containership with well defined loads and operational pattens;
Theotokatos & Livanos (2013) study one specific application to a bulk carrier, while
Thomas et al. (2010) present their method for energy audit applied to a fishing vessel.
Therefore, even if in applied science there is not such thing as a well defined scientific
method, as is instead the case in basic science (Niiniluoto, 1993), the thesis reflects
common practice in the field.
As an additional motivation to the choice of working on a case study, it should
be noticed that the close interaction with the company allowed to get a much better
insight of how the academic work here presented could be used in the ”real world”.
It is sometimes observed in academic researchers to become too much theoretical and
forgetting the challenges that arise when theories and models need to be applied in
practice. As highlighted by Cross et al. (1981), ”knowing how” is often as important as
”knowing that”, where the first refers to explicit, procedural knowledge and the second
to the knowledge that, despite its evident existence, is not structured.

6.3.2 Internal validity


Internal validity refers to the extent to which it is possible to identify a causal connection
between a study and its results (Brewer, 2000). Verification and validation are two of
1
Data referring to the 2000 to 2004 period, (Carlton, 2007, p. 21)

53
6. DISCUSSION

the main methods employed in the evaluation of internal validity (Oberkampf et al.,
2002).
Verification refers to the process of determining that a model implementation ac-
curately represents the developer’s conceptual description of the model and the solution
to the model ˝(DoD). Verification is very important in those cases, such as system
design, when no data for validation is available.
Validation refers to the process of determining the degree to which a model is
an accurate representation of the real world from the perspective of the intended use
of the model ˝(AIAA, 1998). The validation of a model is performed by comparing
model outputs versus experimental measurements. Different levels of validation can be
achieved depending on the thoroughness of this process (Oberkampf et al., 2002).
As described in the dedicated methodology section, energy and exergy analysis in-
volve both a top-down and a bottom-up approach. The two parts are connected to
different issues when it comes to internal validity. From a top-down perspective, the
analysis simply relates to the observation of existing measurements, which allows no
validation. From a bottom-up perspective, instead, the individual models used to sim-
ulate different components could, and should, be validated. In particular, the absence
of data for model validation when the propulsion system is operating at variable speed
can cast some doubts on the applicability of the model to such cases. Furthermore, no
information was available for the validation of individual components either.
The main uncertainty is related to the main engine model; even though a validation
of the model is proposed (see Paper II), it only refers to operations at fixed speed, as
no information is available on variable speed operations. However, the absence of the
related data was also a major justification for employing a larger mechanistic content in
the modelling: in fact, white-box models provide much larger safety when extrapolating
them out of the initial boundaries. This explains why not only pure black-box, but also
hybrid models that employ a lower grade of mechanistic content (such as mean-value
models) were excluded. If extensive data were available on off-load engine speed, the
task of modeling would have been approached in a different way, giving larger room
to the black-box part of the model. It should be noted, additionally, that despite the
modifications proposed in Paper II, the modeling approach used in this thesis is not
dissimilar to what can be found in literature (e.g. Scappin et al. (2012); Payri et al.
(2011); Benvenuto et al. (1998)) and therefore represents an application of common
practice. The validation of the overall results is discussed with the partner company,
and tests are being carried out for the purpose. Unfortunately, to the moment the
thesis is going to print, this kind of validation data has not been provided yet.
A verification process, comparing findings versus existing literature, was performed
instead. As studies available from scientific literature normally only include sailing
operations, the results from the energy analysis used in this thesis were recalculated
for only accounting for the time spent by the ship at sea. From the energy analysis
perspective, results presented by Thomas et al. (2010) and Basurko et al. (2013) can be
used as a confirmation of the orders of magnitude. Figures for total propulsion need,
for example, are quite in accordance in identifying propulsion as the main consumer

54
6.3 Methodology

on board (75.7% in the author’s work, 76% in Thomas et al. (2010) and 84.3%, 87.3%,
and 87.8% in Basurko et al. (2013)).
The results of the exergy analysis here proposed could be partially compared to
what was obtained by Dimopoulos et al. (2012). The paper confirms the fact that
most of system irreversibilities are connected to main engine operations (79% in this
thesis compared to 82% in literature1 ), while relevant losses are also connected to
turbocharger operations (respectively 7.5% and 3.9%). As the exergy analysis proposed
by Dimopoulos et al. (2012) refers to a system equipped with a WHR system it is
not possible to compare the results connected to other components, as the resulting
arrangement is much different for the two case-studies.
Results related to the evaluation of WHR systems are rather consistent with what
expressed in similar studies. The results presented by Dimopoulos et al. (2011) relate
to a similar ratio of propulsive over auxiliary power and show how auxiliary power
demand can be completely met by a recovery system only for high loads. In this case,
full power production using WHR system is guaranteed for only 17% of the time, but
this is mostly connected to the higher auxiliary power need compared to propulsion
(for normal speed transit the auxiliary power need represents 11.4% of total demand,
while on the case study ship this value is normally lower then 10%) and to the use
of two-stroke engines, known for their lower availability of waste heat (Theotokatos &
Livanos, 2013). Similar results are also obtained by Theotokatos & Livanos (2013), who
present the analysis of a simpler recovery cycle (around 34% exergy efficiency2 ) which
leads to a higher need for the operation of auxiliary engines in order to generate the
required electric power. It should be noted that even in this case, in sea-going mode,
the WHR system is able to provide around 72% of the power requirement.

6.3.3 Data quality


The availability and quality of the available information used in a scientific study are of
vital importance for the quality of the work itself. A perfectly built model can still give
misleading results if data are not handled correctly. A model can only be as accurate
as the data it is used to process.
Data quality is often defined as ”fitness for use”, i.e. can only be evaluated in light
of the purpose data are used for (Haug et al., 2001). According to Wang (1996) the
data qualities most found in reviewed scientific literature are:

Intrinsic: The extent to which data values conform to the actual variable to be mea-
sured.

Contextual: The extent to which data are useful to the purpose.


1
Note that the value referred to the result presented in this thesis refers to the aggregation of losses
in the main engine and in the cooling systems.
2
It was not possible, unfortunately, to evaluate the exergy efficiency of the WHR system proposed
by Dimopoulos et al. (2011) based on the information provided in the article. The estimation of
complexity is based on a purely qualitative evaluation of system design, ad described by the authors of
the two studies

55
6. DISCUSSION

Representational: The extent to which data are presented in a clear and unambigu-
ous manner.

Accessibility: The extent to which data are easy to obtain.

As already mentioned in section 3.1.2, it has not been possible to perform exten-
sive, own measurements. Data were made available from one shipping company, who
collaborated with the author, for one of their ships.
The strong collaborative framework with a ship operator in the related industrial
sector (in this case a shipping company) have both positive and negative sides. On the
one hand, in fact, it has not been possible to run own experiments, as the ship was sailing
in very different locations under hardly predictable schedules; moreover, the company
manifested the interest in sharing large amount of available information with the author,
but did not back up any experimental campaign. This translates in inconsistencies in
data quality: continuous monitoring, for instance, provides higher data quality than
noon reports, which is reflected in the accuracy of the results (Aldous et al., 2013).
Referring to the categories of data quality presented before, intrinsic quality was very
variable depending on the specific measurement; contextual quality was limited, as
extensive data were available on variables of little interest to the scope of the thesis
and lacked for other, more important variables. Additionally, data consistency was
rather poor, as very different sources had to be used to be able to perform the analysis.
No issue can instead be reported on data representational and accessibility quality.
On the other hand, the process showed that the proposed methodology is very
resilient to different data quality (this is often referred to as ecological validity (Shadish
et al., 2002)). The analyses was conducted using data from the real world, meaning
that the presented methods do not belong to a restricted community of researchers
but can be used in virtually any real life situation. The alternative of tailoring the
method on measurements performed in a proper scientific way would have improved
the accuracy of the results, but would have made the model of limited use outside the
specific proposed application.

56
7

Future work and


recommendations

Suggestions on how to proceed from the results of this thesis are here presented. These
are divided into proposals to the scientific community (Section 7.1) and recommenda-
tions to the industry (Section 7.2).

7.1 Proposals to the scientific community


• The application of energy and exergy analysis to ship energy systems should be
expanded to include different ship types. This would bring a larger literature on
which to base further studies like the one presented in this thesis.

• This thesis focuses on large vessels, operating on mostly international routes, for
which it is legitimate to assume a steady-state behavior. Some specific applica-
tions do not allow this kind of assumption, such as small inland ferries and fishing
vessels. The influence of transient phenomena can become relevant in particular
when phenomena characterised by different inertiae are coupled. This is for ex-
ample the case of WHR systems, where fast mechanical transients are coupled to
slower thermal ones.

• Engine-propeller interaction showed large margins of improvements in arrange-


ments were the propulsion system is operated at constant speed for using a shaft
generator, especially at low propeller load. The findings presented in this thesis
should be extended to other types of propulsion systems, including slow-speed
engines and fixed pitch propellers.

• Additional focus should be put on the environmental impact from the initial
stage of ship design. Future work should be focused on including not only GHGs
but also other relevant pollutant emissions, such as N OX , SOX , and PM to the
analysis. Issues related to different pollutants are, in fact, strongly interconnected
and should not be addressed separately.

57
7. FUTURE WORK AND RECOMMENDATIONS

• Even if extensive work on Diesel engine models has been performed to date, very
little has been done on their modeling for implementation into energy systems
modeling, especially when it comes to the cooling systems. As showed in paper
I there are large amounts of energy available through different flows from en-
gine cooling, and a more detailed modeling of thermal losses and cooling systems
should be available for the optimisation of WHR systems for using the highest
possible amount of available waste heat.

7.2 Recommendations to the industry


• ”Know what you do, know what to do”. The application of energy and exergy
analysis to the case study ship has been facilitated by the existence of a large,
reliable dataset based on a continuous monitoring system. The availability of
such large and complete databases is still not common in shipping. Resources
should be invested on reliable sensors and online data storage, since this would
expand the information available for a correct planning actions to improve energy
efficiency.

• WHR systems are a promising solution for improving ship energy efficiency. It is
recommended to consider this possibility for both retrofitting and new buildings.
However, as showed in this thesis, the profitability of this technology strongly
depends on ship operational mode, engine type and auxiliary demands. For this
reason, it is recommended to increase the expertise in such systems and in the
evaluation of waste heat availability and quality.

• As suggested by the results of this thesis and by existing literature, engine-


propeller interaction can have a strong influence on ship energy efficiency. This
aspect should be kept under strong focus, both in the operation, retrofitting, and
design of marine vessels. The impact of this interaction on propulsion systems
based on FPPs and two-stroke engines, as well as the influence of sea state and
hull fouling, should be investigated.

• The initial design phase of a vessel should be emphasised more. New knowledge
in the field of systems optimisation and environmental studies should not be
relegated to the last phases of the design, where most of the choices are already
taken, but rather be included in the initial phases of ship design.

• Sampling data is not enough. The analysis of these data is crucial. In this
sense, I would personally like to further emphasise the importance of academic
collaboration. We, as researchers, have an interest in improving our understanding
of systems and reality; you, running a business, have an interest in having access
to knowledge and expertise that might be missing in your human resources. It is
a win-win situation that should happen more often!

58
8

Conclusions

The main findings of this work can here be summarised:


• Energy and exergy analysis applied to ship energy systems provide a useful tool
for the identification of main energy flows and of inefficiencies. For the case study
ship, auxiliary heat and power consumption was found to account together for
30% of the total ship energy consumption. The evaluation of the waste flows
allowed to estimate that a large potential for WHR exists on the case study ship,
particularly in the exhaust gas.
• Engine-propeller interaction can have a strong impact on ship energy efficiency.
In the case study ship, the application of such models suggests that specific ship
fuel consumption can be reduced by 5.8% while increasing the speed from 10
to 12 knots by operating engine and propeller at variable speed instead of at
constant speed. The benefit of an increased energy efficiency of the propulsion
train outweighs the disadvantage of using the auxiliary engines instead of the
shaft generator for meeting on board auxiliary power demand.
• The profitable application of WHR systems is subjected to an analysis of the
availability of waste energy in comparison of auxiliary power demand. For the case
study ship, results suggest that a rather effective system (ηex of 58%) is required if
only heat in the exhaust gas is recovered. The situation can be improved either by
increasing the number of sources of waste heat (the required ηex when including
cooling water goes down to 48%) or by adapting ship operations in order to always
run on two engines (ηex = 39%.)
• Ship design can be improved, with both economic and environmental benefits,
if parameters such as engine type and fuel choice are taken into account from
the initial stages of the design. Results from the application of mathematical
models to the evaluation of alternative arrangements suggested that both fuel
consumption and carbon footprint are strongly influenced by these choices.
All the results and findings presented in this thesis were obtained by addressing
the challenge of improving ship energy efficiency with a systems perspective. This

59
8. CONCLUSIONS

translated in the application of energy and exergy analysis to ship energy systems, as
well in the use of mathematical models for the simulation of these systems without the
need to resort to time consuming sea trials. This approach proved to be particularly
beneficial, and systems thinking is expected to become more and more important in the
challenge of making shipping a cleaner and more energy efficient mean of transportation.

60
Bieler, P., Fischer, U. & Hungerbühler, K. (2004). Modeling
the energy consumption of chemical batch plants: Bottom-
up approach. Industrial and Engineering Chemistry Re-
search, 43, 7785–7795. 28

Blanchard, B. & Fabrycky, W. (2006). Systems engineering and


analysis. Prentice-Hall international series in industrial
and systems engineering, Pearson Prentice Hall, Upper
Saddle River, N.J, 4th edn. 28

References Böme, H. (1983). Ocean shipping: survival at the rock bottom.


Marine Policy, 7, 223–224. 3

Bontempi, G., Vaccaro, A. & Villacci, D. (2004). Semiphysical


modelling architecture for dynamic assessment of power
components loading capability. IEEE Proceedings: Gener-
ation, Transmission and Distribution, 151, 533–542. 28
Aesoy, V., Einang, M.P., Stenersen, D., Hennie, E. & Valberg, I.
Braake, H.B., van Can, H. & Verbruggen, H. (1998). Semi-
(2011). Lng-fuelled engines and fuel systems for medium-
mechanistic modeling of chemical processes with neu-
speed engines in maritime applications. SAE Technical Pa-
ral networks. Engineering Applications of Artificial Intel-
per 2011-01-1998 . 12, 47
ligence, 11, 507–515. 28

AIAA (1998). Guide for the verification and validation of com- Brady, E.F. (1981). Energy conservation for propulsion of
putational fluid dynamics simulations. Reston, VA. 54 naval vessels. Naval Engineers Journal, 93, 131–144. 3

Aldous, L., Smith, T. & Bucknall, R. (2013). Noon report data Brett, B. (2008). Potential market for LNG-fueled marine ves-
uncertainty. In Low Carbon Shipping Conference, London, sels in the United States. Master’s thesis, Massachusetts
UK. 22, 56 Institute of Technology. 3

Atkins, P.W. (1984). The second law . W. H. Freeman and Com- Brewer, M. (2000). Research design and issues of validity.
pany. 24 In Handbook of Research Methods in Social and Personal-
ity Psychology, Cambridge University Press, Cambridge,
Balaji, R. & Yaakob, O. (2012). An analysis of shipboard waste UK. 53
heat availability for ballast water treatment. Proceedings of
the Institute of Marine Engineering, Science and Technology Buhaug, O., Corbett, J.J., Endersen, O., Eyring, V., Faber, J.,
Part A: Journal of Marine Engineering and Technology, 11, Hanayama, S., Lee, D.S., Lee, D., Lindstad, H., Markowska,
15–29. 31 A.Z., Mjelde, A., Nilsen, J., Palsson, C., Winebrake, J.J.,
Wu, W. & Yoshida, K. (2009). Second imo ghg study 2009.
Banks, C., Turan, O., Incecik, A., Theotokatos, G., Izkan, S., Tech. rep., International Maritime Organization (IMO). 4
Shewell, C. & Tian, X. (2013). Understanding ship operat-
ing profiles with an aim to improve energy efficient ship Buxton, I.L. (1985). Fuel costs and their relationship with cap-
operations. In Low Carbon Shipping Conference, London, ital and operating costs. Maritime policy and management,
UK. 33, 51, 52 12, 47–54. 3

Basurko, O.C., Gabiña, G. & Uriondo, Z. (2013). Energy per- Campora, U. & Figari, M. (2003). Numerical simulation of ship
formance of fishing vessels and potential savings. Journal propulsion transients and full scale validation. Proceedings
of Cleaner Production, 54, 30–40. 22, 23, 49, 54, 55 of the Institution of Mechanical Engineers Part M: Journal
of Engineering for the Maritime Environment, 217. 30
Bazari, Z. & Longva, T. (2011). Assessment of imo man-
dated energy efficiency measures for international ship- Carlton, J.S. (2007). Marine Propellers and Propulsion. Else-
ping. Tech. rep., Lloyd’s Register, DNV. 5 vier Ltd., 2nd edn. 11, 53

Checkland, P. (1999). Systems thinking, systems practice. John


Bengtsson, S. (2011). Life cycle assessment of present and fu-
Wiley and Sons, Ltd, Chichester, UK. 10, 17
ture marine fuels. Ph.D. thesis, Chalmers University of
Technology. 5
Choi, B.C. & Kim, Y.M. (2012). Exhaust-gas heat-recovery sys-
tem of marine diesel engine (ii) - exergy analysis for work-
Benvenuto, G. & Figari, M. (2011). Dynamic simulation tech- ing fluids of r245fa and water. Transactions of the Korean
niques for the evaluation of the propulsion system behav- Society of Mechanical Engineers, B, 36, 593–600. 27
ior. Marine Technology and Engineering, 1, 321–348. 30
Clausius, R. (1850). Ueber die bewegende kraft der wärme und
Benvenuto, G., Campora, U., Carrera, G. & Casoli, P. (1998). A die gesetze, welche sich daraus für die wärmelehre selbst
two-zone diesel engine model for the simulation of marine ableiten lassen. Annalen der Physik und Chemie, 155, 368–
propulsion plant transients. In MARIND 98, Second Inter- 294. 23
national Conference on Marine Industry, Varna, Bulgaria.
54 Cross, N., Naughton, J. & Walker, D. (1981). Design method
and scientific method. Design Studies, 2, 195–201. 53
Benvenuto, G., Carrera, G., Figari, M. & Campora, U. (2005).
Interaction between ship propulsion plant automation and DeTolla, J.P. & Fleming, J.R. (1981). A computer model for
simulation, vol. 1. Francis, Taylor and, London, UK. 30 shipboard energy analysis. Naval Engineers Journal, 96,
33–45. 3, 30
Bidini, G., Di Maria, F. & Generosi, M. (2005). Micro-
cogeneration system for a small passenger vessel operating Devanney, J. (2011). The impact of the energy efficiency design
in a nature reserve. Applied Thermal Engineering, 25, 851– index on very large crude carrier design and co2 emissions.
865. 31 Ships and offshore structures, 6, 355–368. 5

Bieler, P., Fischer, U. & Hungerbühler, K. (2003). Modeling Dimopoulos, G., Georgopoulou, C. & Kakalis, N. (2011). Mod-
the energy consumption of chemical batch plants - top- elling and optimization of an integrated marine combined
down approach. Industrial and Engineering Chemistry Re- cycle system. In Proceedings of ECOS , Novi Sad, Serbia.
search, 42, 6135–6144. 28 30, 31, 51, 53, 55

61
REFERENCES

Dimopoulos, G., Georgopoulou, C. & Kakalis, N. (2012). The Fitzgerald, W.B., Howitt, O.J.A., Smith, I.J. & Hume, A. (2011).
introduction of exergy analysis to the thermo-economic Energy use of integral refrigerated containers in maritime
modelling and optimisation of a marine combined cycle transportation. Energy Policy, 39, 1885–1896. 31
system. In Proceedings of ECOS , Perugia, Italy. 27, 30,
53, 55 Flood, R.L. & Carson, E.R. (1993). Dealing with complexity -
an introduction to the theory and application of systems sci-
Dimopoulos, G.G. & Frangopoulos, C.A. (2008a). Synthesis, de- ence. The Language of science, Plenum Press, New York,
sign and operation optimization of the marine energy sys- 2nd edn. 9, 10, 17, 27, 28
tem for a liquefied natural gas carrier. International Jour-
nal of Thermodynamics, 11, 203–211. 30 Gauthey, J.R. & DeTolla, J.P. (1974). The energy crisis and
naval ship research. Naval Engineers Journal, 86, 97–105.
Dimopoulos, G.G. & Frangopoulos, C.A. (2008b). Thermoeco- 3
nomic simulation of marine energy systems for a liquefied
natural gas carrier. International Journal of Thermodynam- Gilbert, P. & Bows, A. (2012). Exploring the scope for comple-
ics, 11, 195–201. 30 mentary sub-global policy to mitigate co2 from shipping.
Energy Policy, 50, 613–622. 4
Dimopoulos, G.G. & Kakalis, N.M.P. (2010). An integrated
modelling framework for the design, operation and control Grimmelius, H., Mesbahi, E., Schulten, P. & Stapersma, D.
of marine energy systems. In CIMAC Conference, Bergen, (2007). The use of diesel engine simulation models in
Norway. 30 ship propulsion plant design and operation. In CIMAC
Congress, Wien, Austria. 28, 30
Dincer, I. & Rosen, M. (2013). Exergy: Energy, Environment
and Sustainable Development. Elsevier, 2nd edn. 23, 26, 27 Grimmelius, H.T. (2003). Simulation models in marine engi-
neering: from training to concept exploration. In COM-
DNV (2012a). Managing the complexity of ship machinery sys- PIT , Hamburg, Germany. 30
tems. Tech. rep., DNV Research and Innovation. 5
Grimmelius, H.T., Boonen, E.J., Nicolai, H. & Stapersma, D.
DNV (2012b). Maritime fuel price and uptake projections to (2010). The integration of mean value first principle diesel
2035. Tech. rep., DNV Research and Innovation. 3 engine models in dynamic waste heat and cooling load
analysis. In CIMAC Congress, Bergen, Norway. 31, 51
DNV (2012c). Shipping 2020. Tech. rep., Det Norske Veritas.
2, 3, 5, 14 Groscurth, H.M., Bruckner, T. & Kümmel, R. (1995). Modeling
of energy-services supply systems. Energy, 20, 941–958. 28

DoD (????). Verification, validation, and accreditation (vva)


Haight, B. (2012). Marine propulsion order survey. Tech. rep.,
recommended practices guide. 54
Diesel and Gas Turbines Worldwide. 53

Duarte, B., Saraiva, P. & Pantelides, C. (2004). Combined


Haug, A., Zachariassen, F. & van Liempd, D. (2001). The cost
mechanistic and empirical modelling. International Jour-
of poor data quality. Journal of Industrial Engineering and
nal of Chemical Reactor Engineering, 2. 28
Management, 4, 168–193. 55

Dupuis, R. & Neilson, J. (1997). Dynamic analysis: sail the ship


Heywood, J. (1988). Internal Combustion Engines Fundamen-
before it’s built. Naval Engineers Journal, 109, 83–91. 10,
tals. McGraw-Hill. 12
30

Hulskotte, J.H.J. & Denier van der Gon, H.A.C. (2010). Fuel
EEA (2013). The impact of international shipping on euro- consumption and associated emissions from seagoing ships
pean air quality and climate forcing. Tech. rep., European at berth derived from an on-board survey. Atmospheric En-
Environmental Agency. 5 vironment, 44, 1229–1236. 31

EIA (2013). Annual energy outlook. Tech. rep., US Energy IEA (2012). Co2 emissions from fuel combustion - highlights.
Information Administration. 3 Tech. rep., IEA International Energy Agency OECD,
Paris, France. 4
Eide, M., Longva, T., Hoffmann, P., Endresen, O. & Dalsren,
S.B. (2011). Future cost scenarios for reduction of ship co2 IFIAS (1974). Energy analysis workshop on methodology and
emissions. Maritime Policy and Management, 38, 11–37. 4 conventions. Tech. rep., International Federation of Insti-
tutes for Advanced Study. 23
EPA (2008). Global trade and fuel assessment - future trends
and effects of requiring clean fuels in the marine sector. Johnson, H., Johansson, M., Andersson, K. & Södahl, B. (2013).
Tech. rep., US Environmental Protection Agency. 3 Will the ship energy efficiency management plan reduce
co2 emissions? a comparison with iso 50001 and the ism
Eyring, V., Isaksen, I.S.A., Berntsen, T., Collins, W.J., Corbett, code. Maritime Policy and Management, 40, 177–190. 1,
J.J., Endresen, O., Grainger, R.G., Moldanova, J., Schlager, 3, 5
H. & Stevenson, D.S. (2010). Transport impacts on atmo-
sphere and climate: Shipping. Atmospheric Environment, Kalli, J., Karvonen, T. & Makkonen, T. (2009). Sulphur con-
44, 4735–4771. 12 tent in ships bunker fule in 2015 - a study on the impacts
of the new imo regulations and transportation costs. Tech.
Faber, J., Eyring, V., Selstad, E., Kågeson, P., Lee, D.S., Buhag, rep., Ministry of Transport and Communications, Finland.
O. & Lindstad, H. (2009). Technical support for european 2
action to reducing greenhouse gas emissions from interna-
tional maritime transport. Tech. rep., CE Delft. 4 Kotas, T. (1980). Exergy criteria of performance for thermal
plant: second of two papers on exergy techniques in ther-
Fake, T., J. Rozenman & Pundyk, J. (1979). Analysis of ship- mal plant analysis. International Journal of Heat and Fluid
board energy systems. In Proceedings of the Intersociety Flow , 2, 147–163. 26
Energy Conversion Engineering Conference. 30
Kuiken, K. (2008). Diesel engines for ship propulsion and power
Figari, M. & Altosole, M. (2007). Dynamic behaviour and sta- plants : from 0 to 100,000 kW . Target Global Energy
bility of marine propulsion systems. Proceedings of the In- Training, Onnen. 12
stitution of Mechanical Engineers, Part M: Journal of Engi-
neering for the Maritime Environment, 221, 187–205. 30 Laerke, S. (2012). Experiences with slow steaming. 12

62
REFERENCES

Larsen, U., Pierobon, L., Haglind, F. & Gabrielii, C. (2013). De- Payri, F., Olmeda, P., Martı́n, J. & Garcı́a, A. (2011). A com-
sign and optimisation of organic rankine cycles for waste plete 0D thermodynamic predictive model for direct injec-
heat recovery in marine applications using the principles tion diesel engines. Applied Energy, 88, 4632–4641. 54
of natural selection. Energy, 55. 31, 51
Petersen, J.P., Jacobsen, D.J. & Winther, O. (2012). Statistical
Lassesson, H. & Andersson, K. (2009). Energy efficiency in ship- modelling for ship propulsion efficiency. Journal of Marine
ping - review and evaluation of the state of knowledge. Science and Technology, 17, 30–39. 31
Tech. rep., Department of Shipping and Marine Technol-
ogy, Chalmers University of Technology and Lighthouse. Pons, M. (2009). On the reference state of temperature when
5 ambient temperature fluctuates. International Journal of
Thermodynamics, 12, 113–121. 26
Leifsson, L.T., Sævarsdóttir, H., Sigurdsson, S.T. &
Vésteinsson, A. (2008). Grey-box modeling of an ocean Rogelj, J., Hare, W., Lowe, J., Van Vuuren, D.P., Riahi, K.,
vessel for operational optimization. Simulation Modelling Matthews, B., Hanaoka, T., Jiang, K. & Meinshausen, M.
Practice and Theory, 16, 923–932. 30 (2011). Emission pathways consistent with a 2c global
temperature limit. Nature Climate Change, 1, 413–418. 4
Leo, T.J., Durango, J.A. & Navarro, E. (2010). Exergy analy-
sis of pem fuel cells for marine applications. Energy, 35, Sabinsky, S. (????). The danish maritime cluster. Presenta-
1164–1171. 27 tion. 3

Lijun, X., Zhou, S. & Yu, A. (1996). Exergy analysis and op- Sack, B.P. (1981). Navy shipboard energy conservation r&d
timum thermoeconomics of dual pressure waste heat re- program. Naval Engineers Journal, 88, 106–116. 3
covery systems for diesel ships. In Proceedings of the Inter-
national Conference on Energy and Environment, 250–255.
Scappin, F., Stefansson, S., Haglind, F., Andreasen, A. & Larsen,
27
U. (2012). Validation of a zero-dimensional model for pre-
diction of NO x and engine performance for electronically
Linné, P. (2012). Ships, air pollution and climate change - controlled marine two-stroke diesel engines. Applied Ther-
some reflections on recent legal developments in the ma- mal Engineering, 37, 344–352. 54
rine environment protection committee. MarIus, 414, 13–
32. 5
SCC (2005). Evolution of low sulfur marine fuel availability.
Tech. rep., Starcrest Consulting Group, LLC. 3
Ma, Z., Yang, D. & Guo, Q. (2012). Conceptual design and per-
formance analysis of an exhaust gas waste heat recovery Schulten, P. (2005). The interaction between Diesel engines,
system for a 10000teu container ship. Polish Maritime Re- ship and propellers during manoeuvring. Ph.D. thesis, Delft
search, 19, 31–38. 31, 51 University of Technology. 30

MaK (????). M32C Project Guide - Propulsion. MaK, last ac- Shadish, W., Cook, T. & Campbell, D. (2002). Experimental and
cessed: 08-08-2013. 21 Quasi-Experimental Designs for Generalized Causal Infer-
ence. Houghton Mifflin, Boston, US. 56
Matuszak, Z. (2008). Losses of exergy in a marine oil-fired
boiler. Archives of Thermodynamics, 29, 97–108. 27 Shi, W. (2013). Dynamics of Energy System Behaviour and
Emissions of Trailing Suction Hopper Dredgers. Ph.D. the-
Mazraati, M. (2011). Challenges and prospects of interna- sis, TU Delft. 30, 31
tional marine banker fuels demand. In OPEC Energy Re-
view 2011 , Blackwell Publishing Ltd, Oxford, UK. 3 Shi, W. & Grimmelius (2010). Comparison of modeling tech-
niques for simulation of fuel consumption of dredgers. In
Mitchell, M. & Jolley, J. (2001). Research Design Explained. COMPIT , 382–395, Gubbio, Italy. 30
4th edn. 52
Shi, W., Stapersma, D. & Grimmelius, H.T. (2009). Analysis of
Molland, A. (2008). The Maritime Engineering Reference Book: energy conversion in ship propulsion system in off-design
A Guide to Ship Design, Construction and Operation. operation conditions. WIT transactions on Ecology and the
Butterworth-Heinemann. 11 Environment, 121, 449–460. 30

Motley, M.R., Nelson, M. & Young, Y.L. (2012). Integrated Shi, W., Grimmelius, H.T. & Stapersma, D. (2010). Analysis of
probabilistic design of marine propulsors to minimize life- ship propulsion system behaviour and the impact on fuel
time fuel consumption. Ocean Engineering, 45, 1–8. 35 consumption. International Shipbuilding Progress, 57, 35–
64. 30
Neilson, J. & Tarbet, R. (1997). Propulsion system simula-
tions: Making the right choice for the application. Naval Shu, G., Liang, Y., Wei, H., Tian, H., Zhao, J. & Liu, L. (2013).
Engineers Journal, 109, 83–98. 30 A review of waste heat recovery on two-stroke ic engine
aboard ships. Renewable and Sustainable Energy Reviews,
19, 385–401. 14, 34
Nguyen, T., Pierobon, L., Elmegaard, B., Haglind, F., Breuhaus,
P. & Voldsund, M. (2013). Exergetic assessment of energy
systems on north sea oil and gas platforms. Energy, in SSPA (2011). North european lng infrastructure project. Tech.
press. 23 rep. 3

Niiniluoto, I. (1993). The aim and structure of applied re- Stone, R. (1999). Introduction to internal combustion engines.
search. Erkenntnis, 38, 1–21. 53 Palgrave MacMillan, London, UK, third edit edn. 12

Oberkampf, W., Trucano, T. & Hirsch, C. (2002). Verification, Stopford, M. (2009). Maritime economics. Routledge, London,
validation, and predictive capability in computational en- New York, 3rd edn. 2
gineering and physics. In Foundations for Verification and
Validation in the 21st Century Workshop, Laurel, Mary- Svensson, E. (2011). The regulation of global SOX emissions
land, US. 54 from ships. Ph.D. thesis, Chalmers University of Technol-
ogy. 5
Oliveira, R. (2004). Combining first principles modelling and
artificial neural networks: a general framework. Computers Tarkir, R. (1979). Application of rankine bottoming diesel en-
and Chemical Engineering, 28, 755–766. 28 gines to marine vessels. SAE Technical papers. 31

63
REFERENCES

Tchanche, B.F., Lambrinos, G., Frangoudakis, A. & Papadakis, UN (2013). World economic situation and prospects. Tech.
G. (2011). Low-grade heat conversion into power using or- rep., United Nations, Department of Economic and Social
ganic rankine cycles - a review of various applications. Re- Affairs. 2
newable and Sustainable Energy Reviews, 15, 3963–3979.
14
UNCTAD (2012). Review of maritime transport. Tech. rep.,
Theotokatos, G. (2008). Ship propulsion plant transient re- United Nations Conference on Trade and Development. 1
sponse investigation using a mean value engine model. In-
ternational Journal of Energy, 2. 30
Van Beek, T. & Van Der Steenhoven, S. (2005). Selection and
optimisation of the propulsion train of fishing vessels. In
Theotokatos, G. & Livanos, G.A. (2013). Techno-economic
Fishing vessels, fishing technology and fisheries, RINA, 99–
analysis of single pressure exhaust gas waste heat recov-
110, Newcastle upon Tyne, UK. 33
ery systems in marine propulsion plants. Proceedings of the
Institution of Mechanical Engineers Part M: Journal of En-
gineering for the Maritime Environment, 227, 83–97. 31, Vivid & Economics (2010). Assessment of the economic impact
51, 53, 55 of market-based measures. Tech. rep., Prepared for the Ex-
pert Group on Market-based measures, International Mar-
Theotokatos, G.P. (2007). A modelling approach for the over- itime Organization. 3
all ship propulsion plant simulation. In WSEAS Interna-
tional Conference on System Science and Simulation in En-
gineering, 80–87, Venice, Italy. 30 Wang, R.Y. (1996). Beyond accuracy: what data quality
means to data consumers. Journal of Management Infor-
Thomas, G., O’Doherty, D., Sterling, D. & Chin, C. (2010). En- mation Systems. 55
ergy audit of fishing vessels. Proceedings of the Institution
of Mechanical Engineers Part M: Journal of Engineering for
the Maritime Environment, 224, 87–101. 23, 49, 53, 54, 55 Winnes, H. (2010). Air pollution from ships. Ph.D. thesis,
Chalmers University of Technology. 31
Tian, G., Juando, A., Heslop, J., Solla, M., Roskilly, T., Saras-
quete, A. & Rodriguez, J. (2012). A holistic model for vari-
Woud, H.K. & Stapersma, D. (2008). Design of propulsion and
able speed generation systems on a roro vessel. In Low
electric power generation systems. IMarEST Publications,
Carbon Shipping Conference, Newcastle, UK. 30
London, UK. 11, 12, 14, 33
Tien, W.K., Yeh, R.H. & Hong, J.M. (2007). Theoretical anal-
ysis of cogeneration system for ships. Energy Conversion WTO (2012). International trade statistics. Tech. rep., World
and Management, 48, 1965–1974. 31 Trade Organization. 2

Tilke, C., Rakitsch, S. & Günthner, W.A. (2010). Specific en-


ergy consumption of ship unloaders. Bulk Solids Handling, Yates, F. (1978). Complexity and the limits to knowledge.
30, 250–254. 31 American Journal of Physiology, 4, R201–R204. 10

64
Paper I
Energy and exergy analysis of a ship, the case
study of a chemical tanker
Francesco Baldi, Cecilia Gabrielii, Karin Andersson
October 23, 2013

Manuscript ready for submission

Abstract
Shipping is already a relevant contributor to global carbon dioxide emissions,
and its share is expected to grow together with global trade in the coming years.
At the same time, bunker prices are growing and companies start to feel the
pressure of growing fuel bills in their balance sheet. In order to address both
challenges, it is important to improve the understanding of how ship energy
consumption is generated, through a detailed analysis of its energy systems. In
this paper, data from one year of operations of a chemical tanker were used,
together with some mechanistic knowledge of ship systems, in order to improve
the knowledge of ship energy use. Energy analysis applied to the case-study
vessel allowed comparing different energy flows and therefore identifying sys-
tem components and interactions critical for ship energy consumption. Exergy
analysis allowed instead identifying main inefficiencies and evaluate waste flows.
This last information was then processed in order to estimate the potential for
waste energy recovery under different conditions. Results showed that propul-
sion is the main contributor to ship energy consumption (70%), but that also
auxiliary heat (16.5%) and power (13.5%) needs are relevant sources of energy
consumption. The potential for waste heat recovery is relevant, especially in the
exhaust gas, which contain an exergy flow sized 14.4% of engine power output.
However, the analysis of feasibility for WHR systems showed how the exhaust
gas alone would not suffice for providing at least the auxiliary power need for at
least 80% of the sailing time, which would require also harvesting waste energy
from the cooling systems.

1
1 Introduction
1.1 Background
The shipping business is currently facing a number of challenges. The recent
global economic crisis and the increase in oil price have influenced negatively
most economic activities based on fossil fuels, a phenomenon that can be ob-
served very strongly in shipping. Shipping activities are strongly related to
trade, (more than 80% of global trade by mass is transported on sea [1]) and
this sector is therefore directly connected to the global economy. Freight rates
have already recovered in several sectors after the global economic crisis, and
the same has happened to bunker fuel prices, which have increased by a factor
of 3 since the 80’s [2] and have already returned to the pre-crisis levels. For the
period between 2006 and 2008, fuel costs were estimated for between 43% and
67% of total operating costs depending on vessel type [3].
This situation is not expected to improve in the future: stricter environmen-
tal regulations on sulphur oxides, nitrogen oxides, and greenhouse gases will
exert an additional leverage to fuel costs [4]. This phenomenon will be even
more pronounced in emission controlled areas (ECA), i.e. USA coastal waters,
the Baltic Sea, and the North Sea, where regulations will be even stricter [5],
and new ECAs (Japan coastal waters, Caribbean Sea) are expected in the com-
ing years. These conditions combined have motivated the shipping business to
start focusing on fuel efficiency. However, unlike what happened during the oil
crisis in the 80’s, efforts are expected to be more durable, as the increase in fuel
prices is not generated by a crisis, but by market evolution.
In order to save fuel, various measures are available and currently imple-
mented. They are often divided into operational and design measures, where
the first mostly relate to improving the performance of existing vessels, while
the second refer to a more efficient ship design for new-buildings. Some op-
erational measures are improvements in voyage execution, engine monitoring,
reduction of auxiliary power consumption, trim/draft optimization, weather
routing, hull/propeller polishing, slow-steaming. Examples of design related
measures can be the use of more efficient engines and propellers, improved hull
design, air cavity lubrication, wind propulsion, fuel cells for auxiliary power
generation, waste heat recovery, LNG as fuel, pump frequency converters, cold
ironing [6]. Several scientific studies have been conducted on these measures;
however, a more detailed treatise of these technologies is out of the scope of this
work. Even if efforts have been put in order to evaluate the benefits associated
with the use of each of these solutions and of their combined effect [6], these ad-
vantages should be evaluated on a ship-to-ship basis. For this reason, a deeper
understanding of energy use on board is considered to be vital.

1.2 Previous work


Studies can be found in scientific literature that report figures connected to ship
propulsion or for selected components. Thomas et al.[7] presented the energy
audit of a fishing vessel, also providing a certain detail about end energy use,
but did not provide any thermodynamic analysis of the different processes and
flows; Shi et al.[8, 9] proposed models for predicting ship fuel consumption in
design and off-design conditions, without considering auxiliary machinery and

2
with no attempt to refer to real operative cycles; Balaji and Yaakob[10] analysed
ship heat availability for use in ballast water treatment technologies, which is an
interesting but very limited domain of application. A more thorough, holistic
thermodynamic analysis of a ship, such as that proposed by Nguyen et al.[11] for
oil platforms energy systems, is lacking in scientific literature. A study on the
ship systems energy analysis, providing a detail modeling and understanding
of the different energy flows encountered on a commercial vessel, is therefore
needed.
Analysis purely based on the first law of thermodynamics lack insight of
the irreversibilities of the systems, as they do not account for the additional
knowledge provided by the second law of thermodynamics. Exergy analysis,
which is based on both the first and the second laws of thermodynamics, can
help addressing this shortcoming. Unlike energy, exergy can be destroyed in
real processes, and the rate of destruction is directly connected to the departure
of the analysed process from a reversible one. Exergy analysis also takes into
account energy quality, and not only quantity. Dincer et al. have thoroughly
explained the use of exergy analysis in several industrial applications [12], while
its utilisation in the shipping sector is still reduced to a limited amount of appli-
cations, mostly related to waste heat recovery systems [13, 14] and refrigeration
plants [15, 16]. Literature on exergy analysis applied to the overall ship energy
systems is however lacking.

1.3 Aim
The aim of this paper is to analyse the energy and exergy flows of an entire ship,
using a mixed top-down and bottom-up approach. The scope of the analysis is
to provide a better understanding of how energy is used on board and where
the largest potential for improvement is located.

2 Methodology
In this section, the methods of energy and exergy analysis are described for the
application to ship energy systems. Their application to a case-study vessel is
proposed.

2.1 Combined top-down and bottom-up approach


The analysis of energy systems is often categorised either top-down or bottom-
up. When a top-down approach is used to model the system’s behavior, large
amounts of experimental data points regarding system inputs and outputs are
processed leading to the construction of an empirical model which, based on past
experience, can predict the behavior of the system. These models tend to be
rather accurate but need large initial data sets for calibration and are not suit-
able for extrapolation [17]. On the other hand, a white-box approach makes use
of mechanistic knowledge of the system in order to model each component, link
them together, and produce the output of the system based on the assumptions
made on the single components. Bottom-up approaches are less accurate, but
provide a larger insight of the system, as well as allow a safer extrapolation of
the results [17]. The use of white- and black-box models is thoroughly discussed

3
in literature, and each of the two approaches might be more or less suitable
depending on the application, i.e. on the desired output, and the availability of
initial information.
In the present case, a combined approach was used, as some features of both
strategies were needed. Pure top-down analysis does not provide any insight of
the system under study, as no detailed analysis of the connections and the flows
is performed [17, 18]. On the other hand, bottom-up analysis does not allow
handling the large amount of data on different system inputs and outputs that is
available for the ship under study [17]. A combination of the two, however, may
be the most appropriate way to make the most use of the available information.

2.2 Exergy analysis


Exergy is defind as the maximum shaft work that can be done by the composite
of the system and a specified reference environment [12]. The exergy content of
a flow depends on the quality of the energy content. Additionally, differently
from energy, exergy is not conserved and can be destroyed, representing the
deterioration of energy quality.
For electrical, potential, kinetic, and mechanical energy exergy and energy
flows coincide, while this is not true for chemical and thermal energies. For
a given amount of matter, its thermal exergy content is defined as showed in
equation 2.2:
EX = m[(h − h0 ) + T0 (s − s0 )] (1)
where EX, h, and s respectively stand for exergy, specific enthalpy, and specific
entropy, while the subscript 0 represents reference conditions.
According to Dincer and Rosen [12] exergy flows calculated according to
equation 2.2 can be divided in three main categories:
Input (EX ˙ in ) : the flow of exergy entering the component.
˙ out ) : the flow of exergy leaving the component.
Output (EX
˙ : the amount of exergy lost in the component operation
Irreversibility (I)
(also known as exergy distruction). This part represents energy quality
deterioration and is defined as I˙ = T0 Ṡgen , where Ṡgen represents the rate
of entropy generation in the component.
The definition of the exergy efficiency of a component is quite debated, and
several authors propose different possible alternative definitions. In this study,
three different quantities will be used:
˙
EX
p
Exergy efficiency is defined for this study as ηex = EX ˙ in , where the sub-
scripts p and in respectively refer to products and inputs. In the case of
∆EX ˙ c
heat exchangers, the alternative definition of ηex = ∆ EX˙ h is used, where
subscripts c and h respectively refer to the cold and the hot fluid. The
definition of the ∆EX are adapted depending whether the component is
meant for cooling or heating. The exergy efficiency gives an estimation of
how efficient the component is in the generation of useful products.
Irreversibility ratio is used according to the definition proposed by Kotas et

al. [19], i.e. λ = EX
˙ in . The irreversibility ratio gives an estimation of
how much energy quality is lost in the component.

4
Task efficiency is here slightly modified from what proposed by Dincer adn
Rosent [12], and is defined as the ratio between the irreversibility in an
ideal exchange at constant temperature difference (here arbitrarily fixed
to 10°C) and the irreversibility in the actual process. The task efficiency
gives an estimation of how close the component behavior is to an ideal
process.
Measurements of seawater temperature are used as reference point for exergy
calculations throughout the entire study, as they represent ambient condition
with which to compare conditions of thermal equilibrium.

2.3 Ship description


The ship here selected as a case study is a Panamax chemical / product tanker.
Relevant ship features are provided in Table 1

Ship feature Value Unit


Deadweight 46760 dwt
Overall length 183 m
Draft 12.2 m
Installed power (main engines) 7680 kW
Installed power (auxiliary engines) 1364 kW
Shaft generator power 3200 kW
Exhaust boilers 1400 kgsteam /h
Auxiliary boilers 28000 kgsteam /h

Table 1: Selected ship main features

The ship is propelled by two 4-stroke Diesel engines rated 3840 kW each.
The two engine shafts are connected to a common gearbox. One of the gears
reduces the rotational speed from 600 rpm to 105.7 rpm, the design speed for
the controllable pitch propeller (CPP). Another shaft from the gearbox connects
it to the electric generator (S/G) which provides 60 Hz current to the ship. Ad-
ditionally, two auxiliary engines rated 682 kW can provide electric power when
the main engine is not in operation, or whenever there is a failure in the shaft
generator system. Figure 1 graphically represents the propulsion arrangement.

2.4 Energy conversion


Energy needs on board of the case-study vessel are provided by: the main
engines (MEs), the auxiliary engines (AEs), and the auxiliary boilers (ABs). In
addition, a shaft generator (S/G) is used to convert mechanical energy generated
by the main engines to electrical energy, while an exhaust gas boiler (EGB) is
used to recovery heat from the main engine exhaust.
Main engine fuel consumption is known from online measurements, while
its power output can be derived starting from measurements of propulsion and
auxiliary power using the following expression:
Pprop PS/G
( ηS + ηS/G )
PM E = (2)
ηGB

5
Figure 1: Ship propulsion system arrangement

Where the variables P and η refer to power and efficiency and subscripts ME,
prop, S/G, S, and GB respectively refer to main engines, propeller, shaft genera-
tor, shaft, and gearbox. PS/G and Pprop are available from online measurements;
ηS is assumed equal to 0.99, as suggested by Shi et al. [9]; ηGB is assumed equal
to 0.983, as reported by the shipyard. ηS/G,des is assumed equal to 0.95, as
reported in dedicated technical papers. As the shaft generator is often operat-
ing at very low load (see Figure 2), its efficiency dependence on load needs is
modelled using a polynomial regression calibrated on the experimental points
reported by Hau [20].
Inlet air conditions are determined using a polynomial regression for com-
pression ratio in the turbocharger based on engine technical data; air temper-
ature in the engine room, before the turbocharger, is assumed equal to 35°C.
Air is cooled after the turbocharger in order to increase its density through
a charge air cooler (CAC); the air temperature before the charge air cooler
is calculated assuming a polytropic compression, whose efficiency is based on a
polynomial regression; the air temperature after the charge air cooler is assumed
equal to 55°C according to common practice on the selected ship; the air mass
flow is calculated assuming unitary volumetric efficiency. The heat exchanged
to the cooling system in the charge air cooler can therefore be calculated as
PCAC = ṁair cp,air ∆TCAC .
The temperature of the exhaust gas is estimated through a polynomial re-
gression, while the mass flow is obtained combining air and fuel flows. The en-
ergy content in the exhaust flow is therefore simply calculated from the temper-
kJ
ature difference with reference conditions, assuming a constant cp,eg of 1.08 kgK .
Heat flows to engine jacket water (JW) cooling, lubricating oil and radiation are
estimated from design values available from technical documentation, under the
assumption that their relative share of the residual heat available in the engine
energy balance remains constant with load.
The engine is connected to a double-level cooling system, composed of a low-
temperature (LT) and a high-temperature (HT) part. JW heat is transferred to

6
the HT circuit, while heat from the CAC is subdivided among the two circuits.
Elaborating data from technical documentation, 22.4% of the charge air cooling
heat is transferred to the low-temperature circuit, while the rest is transferred
to the high-temperature circuit. Mass flows in the HT cooling, LT cooling and
lubricating oil (LO) circuit are assumed to be constant with engine load as they
are operated by engine-driven pumps. LT circuit inlet temperature, HT circuit
outlet temperature, and LO circuit inlet temperature are respectively assumed
equal to 34°C, 85°C, and 60°C according to operative experience coming from
communication with on board technical personnel.
Little technical documentation was available for auxiliary engines. The ef-
ficiency is calculated as a polynomial regression of available data points and
corrected according to ISO standards, while the residual energy is subdivided
among charge air cooling, jacket water, lubricating oil, and radiation according
to values at design point.

2.5 Consumers
Propulsion constitutes the largest energy need on board of the selected vessel. As
shown in figure 1 both main engines are coupled to one CP propeller through
a gearbox. The CP propeller allows operating at fixed main engine speed,
thus allowing the utilisation of a shaft generator. In this study, measured data
for propeller shaft power will be used in order to estimate, using a top-down
approach, the ship propulsion need.
In a similar way, auxiliary power consumption is available as measured value
from online measurements. The distribution of auxiliary loads varies between
different operational modes; this is exemplified in figure 2, which shows the
example of sailing loaded and ballast conditions.

Figure 2: Auxiliary power, frequency of occurrence for different operational


modes

In order to give an estimation of the power needed by different consumers,


data from the electric balance was used. This operation, however, required some
assumptions.

• For seagoing mode (loaded), it is assumed that the power consumption is


subdivided as suggested by the shipyard. Therefore, proportions between
different consumers are maintained. For all points were auxiliary load is

7
larger than 500 kW nitrogen compressors are assumed to account for the
difference between with actual consumption.
• For seagoing mode (ballast) the same repartition is assumed as for seagoing
mode (loaded) if auxiliary power is lower than 500 kW. If power consump-
tion is higher the difference is assumed to be connected to the operations of
nitrogen compressors and boilers auxiliaries (in connection to tank clean-
ing), which are subdivided according to design power.
• For manoeuvring the same assumptions as for seagoing mode (loaded) are
employed.
• For cargo loading and unloading all consumption going over 500 kW is
allocated to nitrogen compressors and cargo pumps, with repartition ac-
cording to maximum installed power. It should be noted that cargo load-
ing operations normally do not require the use of cargo pumps, as port
storage facilities can provide the needed overpressure for loading the cargo,
which is shown by the comparison of the respective distributions in figure
2.
• For waiting time the same proportions as reported in the ship electric
balance are used, with the exception of engine room consumption, which
is halved, since when waiting in port only auxiliary engines are used.
Fuel heating need is needed because of high fuel viscosity, and is computed
starting from the design heat balance and using sea water temperature and outer
air temperature measurements. Hotel facilities needs are calculated assuming a
linear correlation between the value given in the heat balance, assumed at an
outer temperature of 2°C, depending on outer air temperature. Fuel heating,
hotel facilities, and other needs are fulfilled by recovering heat from the exhaust
through an Exhaust Gas Economiser (EGE), when the main engine is active, or
by boilers, when the main engine is shut off. Heat consumption for fresh water
generation is assumed according to common practice equal to 7.5 kg h per crew
member. Since the generation of fresh water is connected to the low temperature
(LT) cooling systems, the value of heat of vaporisation for water was taken at
50°C and equal to 2382 kJ kg .
During ballast legs, steam is needed for tank cleaning. This is a very energy
intensive process, and requires the operation of the auxiliary boilers. Energy
use for tank cleaning is derived from the aggregated boiler fuel consumption,
under the assumption of 90% boiler efficiency accounting for combustion losses
and heat flow in the exhaust gas, limited at 200°C. Auxiliary boilers are also
used when the main engines are not in operation. In this condition, as boilers
are operated at very low load, a reduced efficiency of 80% was assumed instead.
A summary of the main auxiliary consumers is shown in Table 2:

2.6 Data sources and quality


As mentioned in section 2.1 a large amount of measured and technical data were
available and were used in the modeling and analysis process. This section will
shortly clarify the origin and the quality of the different data sources.
The main source of data for the top-down part of the analysis is the con-
tinuous monitoring system (CMS) installed onboard. The accuracy of these

8
Electricity consumers Heat consumers
Nitrogen compressors Fuel handling
Cargo pumps Hotel facilities
Boiler auxiliaries Tank cleaning
Engine room
HVAC
Miscellaneous

Table 2: Main auxiliary consumers on the case-study ship

data, with a sampling time of 15 minutes, is only related to the accuracy of the
sensors on board. Technical documentation (TD) was available for the main
engines and the shaft generator and used as input for numerical regressions.
Ship sea trials (ST) performed by the shipyard when the ship was first sailed
were also available. Finally, some values that were not available by any of the
previous sources were obtained by verbal input from the crew.

3 Results
In this section the results of the energy and exergy analysis of the case-study
ship are presented and analysed.

3.1 Energy analysis


Figure 3 represents the Sankey diagram of ship energy systems, while Table 3
summarises the energy inputs and outputs to different ship components. Propul-
sion represents the main source of energy consumption on the overall yearly op-
erations of the vessel, representing the 70% of the overall ship energy demand.
Hence, efforts directed towards the reduction of propulsive power are therefore
highly justified for the ship under study. Electric (13.5%) and heat (16.5%)
needs are however not negligible and should be taken also into account, espe-
cially because of the large time spent in port. Figure 3 also shows the large
amount of energy that is released to the environment through the funnel and
through the sea water cooling systems. The recovery of this waste energy is
limited to the auxiliary heat needs for accommodation and fuel handling, which
however leave a large part of the energy unused.

3.2 Exergy analysis


However, energy analysis becomes partly inappropriate and can be misleading
when thermal energy flows are compared to electric and mechanical flows; the
first law of thermodynamics, in fact, does not include any consideration about
energy quality [12]. For this reason, the concept of exergy is introduced and
employed in this work, in order to get a better understanding of the amount of
waste energy that can be recovered in order to produce additional electric or
mechanical power. This kind of information can be deduced from the observa-
tion of the Grassmann diagram (Figure 4). Here exergy flows, instead of energy

9
Figure 3: Sankey diagram of ship systems

10
Energy analysis
Ėin Ėout ηen
Main engines
Engine 192.0 80.2 41.7%
Turbocharger 73.3 19.3 26.4%
Exhaust gas economiser 9.0 -
Charge air cooler 16.3 -
Lub oil cooler 19.3 -
Jacket water cooler 18.0 -
HT/LT heat exchanger 30.7 -
LT/SW heat exchanger 52.6 -
Converters
Gearbox 80.2 79.2 98.7%
Propeller 70.5 -
Shaft generator 8.68 7.87 90.6%
Auxiliary power
Auxiliary engines 17.2 6.12 35.5%
Nitrogen compressors 2.15 -
Cargo pumps 0.79 -
Boiler auxiliaries 2.68 -
Engine room 3.55 -
HVAC 1.82 -
Miscellaneous 2.63 -
Auxiliary heat
Boilers 5.56 4.46 80.2%
Tank cleaning 3.16 -
Hotelling 5.56 -
Fuel heating 7.84 -

Table 3: Energy flows and efficiencies for system components. All flows are
given in TJ/year

11
flows, are represented, and a new entry (symbolised as red arrows in the figure)
representing energy destruction is included.

Figure 4: Grassmann diagram of ship systems

12
The results from the exergy analysis are presented in Figure 4.Table 4 shows
values of exergy efficiency (ηex ),

Exergy analysis
Component ηex λ δ ηtask
Main engines
Engine 42.9% 37.2 % 64.2 % -
Turbocharger 35.9% 24.4% 6.4% -
Exhaust gas economiser 67.0% 6.5% 1.1% 4.8%
Charge air cooler 66.4% 6.8% 1.1% 27.5%
Lubricating oil cooler 59.3% 20.1% 1.1% 38.8%
Jacket water cooler 50.1% 15.9% 2.5% 49.5%
HT/LT heat exchanger 59.5% 11.6% 1.8% 38.5%
LT/SW heat exchanger 2.1% 85.3% 5.3% 31.5%
Auxiliary power
Auxiliary engines 61.7% 38.3% 5.3% -
Auxiliary heat
Boilers 29.7% 63.7% 5.4% -
Tank cleaning 25.3% 61.1% 0.8% 4.7%
Hotelling 37.8% 46.7% 0.6% 7.2%
Fuel heating 26.2% 62.1% 1.8% 9.1%

Table 4: Exergy flows and efficiencies for system components. All flows are
given in TJ/year

The largest share of this exergy destruction is originated in a small number


of components: the main engines (64.2% of total exergy destruction) and their
turbochargers (6.4%), the auxiliary engines (6.1%) and the auxiliary boilers
(5.4%). Possible improvements in the performance of these components are a
concern of dedicated research and will not be dealt further in this work.
Figure 4 also helps in the evaluation of the different waste flows with respect
to both energy quantity and quality. It is here shown that the exhaust gas is
a much larger source of potentially recoverable heat then the cooling systems,
contrarily to what could be deducted from Figure 3. In addition, looking a the
task efficiency of the EGE (4.8%) allows understanding how its location and
use could be improved in order to increase the available waste heat. The same
amount of heat could be generated using much lower-grade energy.
Even though their exergy flow is lower than the exhaust gas, the cooling
systems should not be discarded as they still constitute a relevant source of
potentially recoverable energy. Here again the observation of the task efficiency
leads to the identification of which components could be improved in order to
perform the same task while reducing exergy destruction. This is particularly
true for the charge air cooler (ηtask = 27.5%), while improvements could also be
programmed in the lubricating oil cooler (38.8%) and the HT/LT heat exchanger
(38.5%). It should be noted, however, that improvements in the task efficiencies
of heat exchangers would only be beneficial in case installation of a WHR system
is expected, as heat is already supplied to auxiliary consumers in sufficient
quantity otherwise.

13
From the point of view of energy use, the very low efficiency of all auxiliary
heat consumers (tank cleaning, hotelling, and fuel heating respectively have task
efficiencies of 4.7%, 7.2% and 9.1%) shows that it would be possible, by using
a different heat transfer fluid (or, in alternative, steam at a lower pressure), to
generate the same heat requirements while using much lower heat-grade sources.
It is noted, for instance, that fuel handling and hotelling only require tempera-
tures as low as 70-80 °C (a part from fuel heaters before the engine, which warm
HFO up to around 100°C), which could be provided at much lower temperature
than by 9 bar steam.
Finally, figure 4 shows how the on board management of thermal energy can
be improved. Firstly, the use of boilers for generation of auxiliary heat should
be reduced to a minimum, as underlined by the very low exergy efficiency of
both these components.

3.3 Potential for waste heat recovery


Heat recovery is already performed on board, and when the ship is sailing no
additional fuel is required for providing energy for the auxiliary heat needs.
This does not mean, however, that the ship systems are thermodynamically op-
timised. In particular, the possibility for installing a WHR system in order to
utilise waste heat for providing auxiliary power can be an option to be explored
in order to reduce the consumption from auxiliary engines. This study focused
on the use of the results from the exergy analysis in order to estimate the feasi-
bility of the installation of a WHR system on board of the selected case-study
ship. The method here presented is expected to allow a much finer estimation
of whether and how WHR systems should be retrofitted on existing vessels.
Figure 5 shows the percentage of time during which the WHR system pro-

100 Only Exhaust, EGE before WHR boiler


Only exhaust
Only exhaust, no aux heat generation
Exhaust + LT cooling
Fulfillment of auxiliary power need [%]

Exhaust + CAC
80

60

40

20

0
30 40 50 60 70
WHR system exergy efficiency [%]

Figure 5: Fractional coverage of the auxiliary need for different recovery ar-
rangement versus cycle exergetic efficiency

14
vides sufficient power for the auxiliaries, as a function of the recovery arrange-
ment and of the exergy efficiency of the WHR system (the fraction only refers
to the time sailing, both loaded and ballast). The range for WHR exergy effi-
ciency was set between 0.4 (referring to the simple pressure cycle, as proposed
by [21]) and 0.6 (referring to complex, optimized ORCs as proposed by [22]).
The different lines represent the possibility of recovering energy from different
sources:
• Exhaust gas plus LT cooling (auxiliary heat harvested from the exhaust,
after the WHR boiler)
• Exhaust gas plus Charge air cooling (auxiliary heat harvested from the
exhaust, after the WHR boiler)
• Exhaust gas (auxiliary heat harvested from the exhaust, before the WHR
boiler; this solution would mean constitute the easiest modification of the
current arrangement)
• Exhaust gas (auxiliary heat harvested from the exhaust, after the WHR
boiler)
• Exhaust gas (auxiliary heat harvested from the cooling flows)

Vertical lines are drawn in order to show, for each arrangement, the efficiency
that grants that auxiliary power demand is fulfilled for at least 80% of the time.
Observing this figure it is possible to notice that fulfilling both heat and power
needs based on the exhaust alone is not a viable solution, unless very complex
recovery systems are employed (minimum efficiencies of 58% and 64%, the latter
not showed in the figure). When low-grade waste heat is harvested as well more
reasonable efficiencies become sufficient for justifying the installation and use
of WHR systems on board. If the efficiency required in order to provide the
totality of the auxiliary power demand for the 80% of the time spent sailing,
fuel consumption can be reduced by 350 tons, representing a saving of 6.78% of
the overall yearly bunker needs.

3.4 Discussion
The assumptions made in the analysis of the data collected were explained
in detail in section 2. The implications of these hypothesis will be here further
discussed, with a focus on their impact on the validity of the results as presented
in this section.
One strong limitation of the procedure employed lies in the inconsistency of
the input data used in order to elaborate the structure of on board energy flows.
Input data for calculations were obtained from online measurements, manufac-
turers technical documentation, shipyard technical documentation, and reported
measurements from the crew. This mixture of different data sources made it
possible to use all available information, with the drawback of giving up to data
consistency. More accurate data concerning boiler fuel consumption, tempera-
tures across the engines and the different cooling flows, as well as more detailed
information about all different auxiliary needs, would make the analysis much
more consistent and accurate. However, this information was not available, and

15
the authors resolved that employing all usable data, even if coming from dif-
ferent sources, was the best solution for getting the highest amount of insight
on the system under study, without compromising the scientific quality of the
work.
On the other hand, the availability of such a relatively wide and detailed
dataset of online measurements for ship propulsion and electric load, as well
as for main engine fuel consumption, is quite uncommon in the shipping sector
(only limited examples of similar datasets are reported in literature, such as what
presented in [23]). Therefore, even if the methodology described is supposed to
be applicable to any kind of ship, the specific choices and assumptions employed
in the modeling are almost exclusive of the specific case. The composition of
data would be hardly found in exactly the same form for other ships. However,
the overall method of decomposition and analysis, both from the energy and
exergy perspectives, does not depend strongly on the input data. In fact, the
composition of different data sources presented in this study shows indeed how
such analysis can be performed even if no complete database is available.
The results presented in the energy and exergy analysis are expected to be
valid for the selected vessel and its sister ships: as aggregated data over one year
of operation were used any voyage-specific feature (weather influence on propul-
sive power, sea water temperature, etc.) is supposed to be cancelled out when
accounting for longer periods of time. It should be noted, however, that some
phenomena can be observable only under longer time perspectives. In particu-
lar, today’s low markets and high fuel prices have pushed down the operative
speed of the vessel, and it is reasonable to expect that the share of propulsive
power would be larger (together with recoverable energy) if the vessel were to
operate at higher speed. The validity of the results can also be extended to
vessels of similar scope, even if of different sizes. It is legitimate to expect that
while quantities will vary sensibly from ship to ship, the conclusions and con-
siderations derived from the analysis of this ship can still be accounted as valid.
Much more care should be put when extending the validity of the analysis to
other ship types. As can be easily observed, the largest share of energy demands
are not type-specific: heat demand for fuel handling and accommodation, power
demand for auxiliary, and propulsion power need are present on most of mer-
chant vessels in operation. However, the applicability should, in this case, be
analysed on a case-by-case principle.

4 Conclusion
The paper presented the energy and exergy analysis of a chemical / product
tanker, based on a mixed top-down and bottom-up approach applied to one
year of ship operation. The exergy analysis was used as a basis for evaluating
the installation of waste heat recovery on the vessel.
The analysis showed, as expected, that propulsion power is the major energy
consumption, while also demonstrating that auxiliary demands of both electric
power and heat are not negligible. A large amount of energy is wasted to
the environment through engine cooling and exhaust gas. Through the exergy
analysis it has been possible however to identify what the largest sources of
exergy loss and destruction are. The largest exergy losses is connected to the
exhaust, while losses connected to engine cooling are negligible. Large amounts

16
of exergy are instead destructed in the boiler and in the cooling systems, as
exchanges are not optimised for conserving energy quality.
The analysis of the possibilities for energy recovery showed that there is
large potential for the application of such technology to the selected vessel.
Assuming the operation of a WHR system in sailing conditions the exergetic
efficiency required by the system in order to fulfil a minimum value of 80% of
the auxiliary power demand would be between 48% and 64% depending on the
exergy flows selected for energy recovery. Such a system would allow the ship
saving an estimated 6.78% of the yearly fuel consumption.

References
[1] UNCTAD, Review of maritime transport, Tech. rep. (2012).

[2] IEA, Co2 emissions from fuel combustion - highlights, Tech. rep., IEA In-
ternational Energy Agency OECD, Paris, France (2012).
[3] J. Kalli, T. Karvonen, T. Makkonen, Sulphur content in ships bunker fule
in 2015 - a study on the impacts of the new imo regulations and transporta-
tion costs, Tech. rep., Ministry of Transport and Communications, Finland
(2009).
[4] DNV, Shipping 2020, Tech. rep., Det Norske Veritas (2012).
[5] Vivid, Economics, Assessment of the economic impact of market-based
measures, Tech. rep., Prepared for the Expert Group on Market-based mea-
sures, International Maritime Organization (2010).

[6] DNV, Pathways to low carbon shipping - abatement potential towards 2030,
Tech. rep., Det Norske Veritas (2010).
[7] G. Thomas, D. O’Doherty, D. Sterling, C. Chin, Energy audit of fishing
vessels, Proceedings of the Institution of Mechanical Engineers Part M:
Journal of Engineering for the Maritime Environment 224 (2) (2010) 87–
101.
[8] W. Shi, D. Stapersma, H. T. Grimmelius, Analysis of energy conversion
in ship propulsion system in off-design operation conditions, Energy and
Sustainability 1 (2009) 461–472.

[9] W. Shi, H. T. Grimmelius, D. Stapersma, Analysis of ship propulsion sys-


tem behaviour and the impact on fuel consumption, International Ship-
building Progress 57 (1-2) (2010) 35–64.
[10] R. Balaji, O. Yaakob, An analysis of shipboard waste heat availability for
ballast water treatment, Proceedings of the Institute of Marine Engineer-
ing, Science and Technology Part A: Journal of Marine Engineering and
Technology 11 (2) (2012) 15–29.
[11] T. Nguyen, L. Pierobon, B. Elmegaard, F. Haglind, P. Breuhaus, M. Vold-
sund, Exergetic assessment of energy systems on north sea oil and gas
platforms, Energy (0).

17
[12] I. Dincer, M. Rosen, Exergy: Energy, Environment and Sustainable Devel-
opment, 2nd Edition, Elsevier, 2013.

[13] B. C. Choi, Y. M. Kim, Exhaust-gas heat-recovery system of marine diesel


engine (ii) - exergy analysis for working fluids of r245fa and water, Trans-
actions of the Korean Society of Mechanical Engineers, B 36 (6) (2012)
593–600.
[14] G. Dimopoulos, C. Georgopoulou, N. Kakalis, The introduction of exergy
analysis to the thermo-economic modelling and optimisation of a marine
combined cycle system, in: Proceedings of ECOS, Perugia, Italy, 2012.
[15] R. Y. Lin, X. Yu, J. Li, Y. Li, W. Wang, Exergy analysis for lng refrigeration
cycle, in: International Conference on Computer Distributed Control and
Intelligent Environmental Monitoring (CDCIEM), Changsha, Hunan, CH,
2011, pp. 211–214.
[16] L. Mattarolo, Energy economy and heat recovery with particular reference
to sea transport, International Journal of Refrigeration.
[17] B. Duarte, P. Saraiva, C. Pantelides, Combined mechanistic and empirical
modelling, International Journal of Chemical Reactor Engineering 2.

[18] G. Bontempi, A. Vaccaro, D. Villacci, Semiphysical modelling architecture


for dynamic assessment of power components loading capability, IEEE Pro-
ceedings: Generation, Transmission and Distribution 151 (2004) 533–542.
[19] T. Kotas, Exergy criteria of performance for thermal plant: second of two
papers on exergy techniques in thermal plant analysis, International Jour-
nal of Heat and Fluid Flow 2 (4) (1980) 147–163.
[20] E. Hau, Wind Turbines, Springer Berlin Heidelberg, 2013.
[21] G. Theotokatos, G. A. Livanos, Techno-economic analysis of single pres-
sure exhaust gas waste heat recovery systems in marine propulsion plants,
Proceedings of the Institution of Mechanical Engineers Part M: Journal of
Engineering for the Maritime Environment 227 (2) (2013) 83–97.
[22] U. Larsen, L. Pierobon, F. Haglind, C. Gabrielii, Design and optimisation
of organic rankine cycles for waste heat recovery in marine applications
using the principles of natural selection, Energy (0).

[23] J. P. Petersen, D. J. Jacobsen, O. Winther, Statistical modelling for ship


propulsion efficiency, Journal of Marine Science and Technology 17 (1)
(2012) 30–39.

18
Paper II
A validated zero-dimensional four-stroke medium speed
Diesel engine model for waste heat recovery marine
applications

Francesco Baldia,∗, Ulrik Larsenb , Cecilia Gabrieliia , Karin Anderssona


a Chalmers University of Technology; Department of Shipping and Marine Technology,
SE-41296 Gothenburg, Sweden
b Technical University of Denmark; Department of Mechanical Engineering, DK-2800 Kgs.

Lyngby, Denmark

Abstract
As fuel prices increase and global warming awareness raises, there is need
for more advanced tools for improving design and operation of energy systems
toward increasing energy efficiency. In this framework, the aim of this paper
is to derive a zero-dimensional single zone mathematical model for predicting
outputs suitable for a combined energy system analysis of four-stroke marine
medium speed Diesel engines and waste heat recovery systems. The use of
Redlich-Kwong real gas equation for high pressure phenomena is proposed; Chen
and Winterbone correlations for friction losses are tested and compared. The
model is calibrated and validated on acceptance test records for a Caterpillar
MaK 8M32C and a MAN 9L27/38 marine engine, with very accurate predictive
results. Tests on operational data show similarly high accuracy. The model
is applied to the optimization of the propulsive efficiency of a merchant vessel.
Analysis using the model shows that, at medium-low ship speeds, the higher effi-
ciency in auxiliary power generation related to the use of a generator connected
to the main engine instead of auxiliary engines is overcome by the decreased
propeller efficiency when operating at fixed speed. Moreover, the evaluation of
the waste heat recovery potential from the exhaust gas leads to the conclusion
that the standard auxiliary power needed onboard the vessel can be provided
by a rather simple recovery cycle (ηex = 0.32), thus generating large potential
reduction in fuel consumption.
Keywords:
Diesel engine, Modeling, gray-box, zero-dimensional, waste heat recovery

∗ Corresponding author
Email addresses: francesco.baldi@chalmers.se (Francesco Baldi), ular@mek.dtu.dk
(Ulrik Larsen), cecilia.gabrielii@chalmers.se (Cecilia Gabrielii),
karin.andersson@chalmers.se (Karin Andersson)

Preprint submitted to Applied Energy October 9, 2013


Nomenclature

DAE Differential Algebraic Equation


EVO Exhaust Valve Opening
GA Genetic Algorithm
LHV Lower Heating Value
WHR Waste heat recovery
Ea Activation energy [kJ]
αf r Friction weighting coefficient
s̄p Mean piston speed [m/s]
s̄p Mean piston speed
∆i Combustion duration [CAD]
η Efficiency
ω Engine speed [rad/s]
kg
Φ Ship specific fuel consumption [ nm ]
τid Ignition delay [ms]
θi Crank angle [CAD]
g
 
bsf c Brake specific fuel consumption kW h

f mep Friction mean effective pressure [bar]


J
H, h Enthalpy [J, kg ]

m Mass [kg]
n Engine speed [rpm]
P Power
p pressure, [bar]
Q Heat [J]
kJ
R Specific gas constant[ kgK ]
kJ
R0 Universal gas constant [ kmolK ]
T Temperature [K]
J
U, u Internal energy [J, kg ]

2
V Cylinder volume [m3 ]
kg
v Specific volume [ m3]

vship Ship speed [kn]


Vsw Swept volume, [m3 ]

W Work [W]
y Gas mass fraction
AE Auxiliary engine

b Brake
ex Exergy
f Fuel
f Fuel

fr Friction
fs Fixed speed
GB Gearbox

in Inlet manifold
in Inlet
max Maximum
ME Main engine

min Minimum
out Outlet manifold
out Outlet

s Propeller shaft
SG Shaft generator
vs Variable speed
w Wall

3
1. Introduction

1.1. Background
In recent times two main factors have exerted a strong influence on the
development and the use of fossil-fueled energy systems and, in particular, of
internal combustion engines. On one hand, the awareness is growing about
human contribution to climate changes in terms of greenhouse gases emissions
[1]. On the other hand, the fuel market is experiencing a new, large increase in
demand, mainly triggered by developing countries growing economies, that is
not matched by a proportionate increase in resources availability [2]. The joint
influence of these two elements has brought a new, rising interest in technologies
for reducing engine fuel consumption.
One of the directions researchers have started to look at is a better under-
standing of the connections that can be found in complex energy systems. In
this kind of structures a simple component-by-component optimization could
be inefficient and even lead to the undesired phenomenon of sub-optimization.
However, the many possible configurations that could be applied for each en-
ergy system do not allow a simple, straight-forward experimentation process,
that would turn out to be too expensive and time consuming. For this reason,
the analysis and optimization of energy systems is subject to the use of accurate
predictive calculation models.
If countless examples of the application of modeling and optimization tech-
niques can be found for land-based systems, such as cogeneration or industrial
plants, the activity in the sector of ship design and operation control has not
been put under the same focus. Rather extensive research has been published fo-
cusing on the main propulsion systems: Theotokatos presents a simplified mod-
eling approach for the overall ship propulsion system model, both in steady-state
and in transient operations [3]; Campora and Figari propose a similar analysis
making use of models with higher mechanistic content and providing valida-
tion of the system transient behavior [4]; Schulten and Stapersma present an
analysis of the uncertainty in relation with the validity of a ship’s model as a
complex system. [5]; Vrijdag et al. propose a modeling procedure, complete
of verification, calibration and validation, of ship propulsion [6]. However, only
little work has been done on the concept of energy system modeling in shipping,
even though very interesting results have been showed by Dimopoulos [7, 8] and
Grimmelius [9, 10].

1.2. Diesel engine modeling


Diesel engines are used as prime movers in most of the energy systems they
are part of, meaning that every error in their modeling will propagate all through
the system. Hence, a large effort is needed in the Diesel engine modeling. On the
other hand, as the model is used for energy system optimization, computational
time must be kept reasonably low.
Two main alternatives have been presented and used for energy system mod-
eling in shipping. Grimmelius et al. [9, 10] make use of mean value first principle
models, which are an evolution of ideal models, where a higher level of detail

4
in the description of the cycle is employed. Dimopoulos et al. [7, 8] instead
use a Diesel engine mean value model where the Diesel cycle is simulated as a
flow restriction with heat addiction. In spite of their apparent simplicity, both
modeling techniques show very impressive results in the prediction of engine
power output and efficiency.
Zero-dimensional thermodynamic models, also known as ”filling and emp-
tying” or ”crank angle” models, can be a possible solution for increasing the
simulation while keeping model complexity and calculation time at reasonable
levels [11]. In these models, the mass and energy conservation equations, to-
gether with the gas state equation, are solved in their differential form all along
the cycle, so that the properties of the gas within the combustion chamber, such
as pressure and temperature, can be calculated. Zero-dimensional models are
often further subdivided in single-zone and multi-zone models, the latter subdi-
viding the combustion zone in two ore more sub-zones for a better prediction of
local temperature peaks. As showed by both Pariotis et al.[12] and Kumar et al.
[13], single-zone models provide a reasonable accuracy for in-cylinder pressure,
while keeping computational time low.
Several studies have been performed on single-zone models for compression
ignition engines. However, medium-speed marine engines, even though still
based on a Diesel process, are significantly different from automotive engines.
Different limitations in terms of weight, space, cost, and maintenance allow
marine engines to reach much higher pressures (up to 200 bar and more) and,
therefore, much higher efficiencies than what is observed in most automotive
application, with values for brake specific fuel consumption (bsf c) down to 160
g
kW h . High pressures in particular, as will be explained in section 2, do not allow
using the ideal gas equation for modeling gas behavior during compression and
combustion.
Therefore, previous work on zero-dimensional single-zone modeling of Diesel
engines was deemed unfit for the application to marine energy systems. Papa-
giannakis et al. [14], Kannan et al. [15], Kökkülünk et al. [16], and Awad et al.
[17] validate their results on engines characterized by lower maximum pressures
g
(60 to 100 bar) and higher design bf sc (200 to 265 kW h ). Several additional
studies are available, sharing the same limitation to marine applications.
Applications of zero-dimensional models explicitly for marine engines are
less numerous. Scappin et al. [18] apply a two-zone model to a marine engine in
a similar pressure and efficiency range, but with a two-stroke cycle. Benvenuto
et al. [19] focus on a 4-stroke cycle instead, but employ a more computational-
intense two-zone model.
Finally, friction is often overlooked in most modeling studies. Benvenuto et
al. [19] refer to an ”ad-hoc” correlation provided by the manufacturer, while
none of the other cited references mention correlations or methods for calculating
friction losses. Livanos et al. [20] showed how, for a marine engine, friction losses
account for 2.6% to 5% of fuel energy input. This contribution is far from being
negligible.

5
1.3. Objective
The objective of this study is to provide a numerical tool for the modeling
of four-stroke marine Diesel engine performances to be used as part of a larger
energy system model. The model should therefore be able to combine good
accuracy and low computational time. Additionally, high cylinder pressure and
need for a friction correlation are addressed in the modeling. The model should
adapt to several different engines with relatively low effort, i.e. not requiring
extensive experimental campaigns and easily available calibration data. The
model focuses on the ability to predict engine main outputs for use in energy
system models, i.e. shaft power and exhaust gas energy content and quality.
The model is used in the operational optimization of the propulsion system of
a merchant vessel.
In section 2 model equations and assumptions are presented. In section 3 the
calibration and validation of the model are described. In section 4 results from
the application of the model are presented and discussed. Finally, in section 5,
conclusions on the presented work are drawn.

2. Mathematical model

Engine evolution was modeled throughout five main phases: compression, in-
jection, combustion, expansion, and post-exhaust valve opening (EVO)blowdown.
Each phase was modeled using a different set of differential equations, that will
be described in the following sections. Modeling the thermodynamic proper-
ties evolution during the gas-exchange phase is omitted, as the gas-exchange
phase involves complex thermo-fluid dynamics interactions, which are hardly
predictable with a 0-dimensional model. An approximation, described later in
this section, is therefore used for the exhaust and intake strokes.
A set of case-dependent parameters employed by the model needs to be de-
fined. This involves engine geometrical parameters (number of cylinders, cylin-
der bore and stroke) and calibration parameters, which will be described in
detail in section 3. Five input variables are needed to the model: inlet air tem-
perature and pressure from the turbocharger (after the cooling stage, if present),
inlet fuel flow, engine speed, and fuel lower heating value (LHV) . Inlet air flow
is calculated assuming unitary volumetric efficiency.

2.1. Model equations


Energy conservation is implemented in the form of the following differential
equation:
U̇ = Q̇f − Ẇ − Q̇w + ṁin hin − ṁout hout (1)
Where U represents the internal energy, Qf the heat released by the combustion,
W the work, Qw the heat lost to the cylinder walls, and ṁin hin and ṁout hout
the inlet and outlet enthalpy flows.

6
Equations for work, specific internal energy and specific enthalpy come from
basic thermodynamic principles. Volume evolution is represented by equation:
Vsw h p i
V = Vmin + f + 1 − cosθ − (f 2 − sin2 θ) (2)
2
where f represents the ratio between connecting rod length and crank shaft
diameter, Vmin the minimum cylinder volume, Vsw the swept volume, and θ the
piston position, in crank angle degrees. All these values were made available by
the manufacturer.
The ignition delay (τid ) is calculated as proposed by Heywood [11]:
( "   0.63 #)
1 1 1 21.2
τid = (0.36 + s̄p )exp Ea − (3)
6n R0 T 17190 p − 12.4

where n represents the engine speed [rpm], s̄p the mean piston speed [m/s], Ea
the activation energy [kJ], R0 the universal gas constant, T the gas temperature
[K], and p the pressure inside the cylinder.
Combustion is subdivided in a first, fast premixed combustion, resulting
from the mixing of fuel and air charge that occurs during the ignition delay,
and a second, slower mixing-controlled combustion [18, 21, 22, 23]. Both heat
releases are modeled using a Wiebe function, which was showed to be the most
reliable approximation, among others, by Ding et al. [23]:
 m "  m+1 #
180ω Qc,i θ − θ0 θ − θ0
Q̇f = a(m + 1) exp −a (4)
π ∆θi ∆θi ∆θi

where ω represents the angular speed rad


 
s and ∆θi the duration of the com-
bustion phase [CA], while a and m are parameters depending on the fuel. Since
no experimental data were available for heat release, values of respectively 5
and 2 where used, as suggested by Heywood [11]. Premixed combustion dura-
tion was set to 7◦ according to Miyamoto et al.[24], while diffusive combustion
duration was subjected to calibration. Gas composition is calculated under the
hypothesis of complete combustion.
Several empirical correlations are available for modeling heat losses to cylin-
der walls, as reviewed by [13]. Both Annand and Woschni correlation are used
in other examples of marine engine modeling [18] [19] . However, as the single
zone approximation averages temperatures over the whole combustion cham-
ber, the use of Annand correlation, which involves non-linear dependence on
gas temperature, was deemed unsuitable. Therefore the heat transfer coeffi-
cient is approximated using Woschini correlation:
 
0.8 0.8 −0.2 −0.55 W
h = 129.9p u d T (5)
m2 K

where p represents the cylinder pressure [bar], d the cylinder diameter [m], T the
cylinder gas temperature [K] and u is a coefficient depending on the mean piston

7
speed, cylinder swept volume and compression ratio. Cylinder wall temperature
is assumed to be constant, as experimentally verified by Rakopoulos et al. [25].
The blow-down that follows the opening of the exhaust valves is responsible
for a considerable loss of indicated work that should be accounted for. Equation
6 can be used in order to avoid discontinuities in the model and to predict engine
performance with reasonable accuracy without requiring detailed knowledge of
valve geometry [18]:

pEV O − pout eθEV O −θ−1


p= + pout (6)
e−1 −θEV O + θ + 1
where pEV O represents the cylinder pressure at the instant when the outlet
valve opens, pout the pressure in the exhaust manifold, and θEV O the angle of
exhaust gas opening.
As the model refers to heavily turbocharged engines, additional work is
produced during the intake/exhaust strokes. As these phases are not modeled
in detail, equation 7 is here proposed in order to approximate the additional
work produced in the pumping cycle:

Wpc = Vsw (pin − pout )ηpump (7)

where Wpc represents the additional work produced by the pumping cycle, pi n
the inlet pressure, po ut the outlet pressure, and ηpump the loss of efficiency due
to the non-ideal valve behavior and is subject to model calibration. Heat losses
occurring in the gas-exchange phase are assumed to be negligible, the largest
part of heat transfer taking place during the combustion phase [26, 22].
Exhaust gas temperature is calculated through an energy balance of the
engine over the entire cycle:

ṁf LHV − Ẇb − Ẇf r − Q̇w


∆heg = (8)
ṁair + ṁf

exhaust gas temperature is calculated starting from its enthalpy value and com-
position. Ideal gas behavior is assumed in this case, hence no dependence on
outlet pressure is considered.
Enthalpy difference in both compressor and turbine is calculated using the
concept of adiabatic (also referred as polytropic) efficiency, which was preferred
to the isentropic efficiency since it allows to take into account the variation
in the efficiency due to varying pressure drops through the turbomachine [27].
Finally, the balance ẆC = ẆT is used for iteratively calculating the pressure in
the exhaust manifold. Both the turbine and the compressor are assumed to be
adiabatic.

2.2. Real gas versus ideal gas equation


The choice of the state equation for gas evolution in the cylinder is a con-
trovert subject. Most researches make use of the ideal gas law, which has been
proved effective in a dedicated study by Lapuerta et all [28]. However it is well

8
known that for high pressures the ideal gas approximation becomes less justi-
fied. Danov et al. [29] showed how for pressure values higher than 80-90 bar
the difference between ideal and real gas behavior stops being negligible.
This model is built for heavy duty Diesel engines where pressure can reach
values up to more than 200 bar. As suggested by Wark [30], for pressure values
up to 300 bar, the Redlich-Kwong equation of state (equation 9) is recommended
for describing the thermodynamic behavior of gases:
 
a
p √ (v − b) = RT (9)
v(v + b) T
where p, T and v respectively represent pressure [Pa], temperature [K] and
specific volume of the gas inside the cylinder [m3 /kg], while a and b are two
coefficients that can be analytically derived for each gas component. According
to Danov et al. [29], the thermodynamic behavior of the mixture subordinates
to the Amago law up to pressures of 300 bar and temperatures of 3300 K.
Therefore, the following mixing rule applies for real gas mixtures:
X
w= yi wi (10)
i

where w represents a generic thermodynamic property and y the mass frac-


tion of the pure gas in the mixture.

2.3. Friction
The application of a numerical correlation was preferred to a more detailed
modeling of friction-related phenomena. Friction losses can increase up to 40
% because of high pressures [31], and marine Diesel engines generally feature
very high compression pressures. Therefore, correlations 11 and 12 respectively
proposed by Chen et al. [32] and Winterbone et al. [33] where considered for
use:
pmax
f mepC = 0.137 + + 0.162v̄sp (11)
200
pmax n
f mepW = 0.061 + + 0.294 (12)
60 1000
where pmax represents the maximum pressure in the cylinder [bar], n the engine
speed [rpm] and s̄p the mean piston speed [m/s].
The equations are conceptually identical and only differ by the coefficients,
and different authors implement different correlations. For instance, Yun et al.
[34] use the Winterbone correlation, while Schulten, after calibration, gets to
the same coefficients as Chen et al. [35].
The two correlations were separately tested versus the third possibility of
using a weighted average of equations 11 and 12:

f mep = αf r f mepC + (1 − αf r )f mepW (13)

where the friction weighting coefficient αf r was subject to model calibration.

9
3. Model calibration and validation

3.1. Parameters sensitivity analysis


Since a few parameters necessary for engine simulation were not available
from engine technical documentation, an estimation was needed. A serial gray-
box approach was used, in which experimental results were used to calibrate a
set of parameters. Data from two engines were used: a MaK 8M32C (A engine)
and a MAN 9L27/38 (B engine). Their main parameters are showed in Table 1.
For a more detailed description of the engines please refer to the engine technical
guides, available on MaK and MAN websites [36, 37].

A B
Rated power [kW] 3840 3060
Cylinders 8 9
Bore [mm] 320 270
Stroke [mm] 480 380
Design speed [rpm] 600 800

Table 1: MaK 8M32C and MAN 9L27/38 main technical details

Experimental measurements were available from acceptance test records


from both engines. For comparison with model results, data regarding SFOC
and exhaust gas temperature were used.
Eight parameters needed to be estimated: diffusive combustion duration
[CA], injection duration [CA], pumping cycle efficiency, turbine and compressor
polytropic efficiency, engine wall temperature [K], injection timing [CA, after
BDC], inlet valve timing [CA, after BDC], and compression ratio. Injection
timing is often a function of engine load, which in the model is taken into
account using a linear correlation, whose two parameters are evaluated through
model calibration.
The influence of different calibration parameters on model output was esti-
mated. A default value and a range of variation were assigned to each parameter,
according to the incertitude of the value. Injection timing was observed being
the most relevant parameter, followed by combustion duration and wall tem-
perature. These three parameters were therefore employed in the calibration
process. The remaining parameters, when unknown, were arbitrarily assumed,
under the assumption that their influence on model output is limited. Assump-
tions are 50 CA degrees for injection duration, 0.85 for turbocharger polytropic
efficiency, and 0.9 for pumping cycle efficiency.

3.2. Calibration
A Genetic Algorithm (GA) minimizing the total weighted residual error ()
for the three variables was used. Based on the principles of natural selection,
first formulated by Charles Darwin, the GA can be applied to resolve a large
variety of simulation issues, as for example an optimization or calibration [38].

10
In this work a GA coded as a Matlab tool box (GATbx) has been obtained
from the Department of Automatic Control and System Engineering of The
University of Sheffield, UK and it is freely available online [39].
The accordance between calibrated and experimental data is finally showed
in Tables 2 and 3, where M and P respectively refer to measured and predicted
ones:

Test bsfc[g/kWh] Exhaust Temp. [K]


[% MCR] M P  M P 
110% 192.9 192.5 -0.21% 799 802 0.38%
100% 190.9 190.8 -0.05% 777 780 0.39%
85% 189.0 190.1 0.58% 761 761 0%
50% 199.0 198.8 -0.10% 756 766 1.32%

Table 2: Calibration results for A engine

Test bsfc[g/kWh] Exhaust Temp. [K]


[% MCR] M P  M P 
110% 203.2 202.8 -0.19% 826 853 3.21%
100% 198.5 198.2 -0.16% 777 799 2.84%
85% 195.0 194.8 -0.13% 737 758 2.85%
75% 193.8 193.9 0.06% 716 734 2.45%
50% 197.5 198.3 0.41% 678 693 2.32%

Table 3: Calibration results for B engine

3.3. Validation
The proposed model was tested versus the calibration dataset of both men-
tioned engines, through the calculation of the ”prediction coefficient of determi-
2
nation” RP rediction statistics. When the lack of a large amount of experimental
points does not allow the repartition of the original dataset into a calibration and
a validation part, the predictive ability of the model can be estimated through
2
the RP rediction statistic [40].
The analysis shows that the model, regardless the engine used for testing,
2
features a good accuracy in the prediction of both bf sc (RP rediction for A and
B engine respectively of 0.820 and 0.858) and exhaust temperature (0.913 and
0.852). Contribution to the cumulated error is generated evenly over all different
points, thus showing that the model is able to predict the physical behavior of
the engine even for extrapolation cases.
In order to further verify model accuracy, measurements from six ships em-
ploying engine A were used. On the selected ship, the engine shaft is connected
to a gearbox, which transfers mechanical power to the propeller shaft and to
an electrical generator (hereafter called ”shaft generator”). The propeller shaft
power (Ps ) and the shaft generator power (PSG ) are calculated after measuring

11
shaft speed and torque (for the propeller shaft) and output voltage and cur-
rent (for the shaft generator). In order to calculate engines brake power (Pb ),
equation 14 was employed:
 
Ps PSG 1
Pb = + (14)
ηs ηSG 2ηGB

where ηs , ηSG , and ηGB represents the efficiencies of respectively propeller shaft,
shaft generator, and gearbox. According to Shi et al.[41] shaft efficiency can
range between 98% and 99%, therefore a mid-range value of 98.5% was used.
ηSG was assumed equal to 90% referring to manufacturers technical papers. A
value of 98.3% for ηGB was given by the shipyard. The total brake power is
divided by 2, since two main engines of the same model are simultaneously used.
Only fuel flow, engine speed and fuel LHV were available as inputs to the model.
The inlet temperature was set to a constant value of 55 ◦ C, which is used as
standard set point for ship operations. Other values were assumed as stated in
section 2.
Fig. 1 shows model results compared to measured values. Error bars for the
measured SFOC are calculated starting from a respective 2% and 3% measure-
ment accuracy for propeller power and fuel mass flow, as reported in technical
documents. Fig. 1a shows how the model is able to predict the engine output
largely within the accuracy of the measurements. The residuals are plotted ver-
sus load in Fig. 1b, showing no specific pattern. The final regression accuracy
is R2 = 0.742

(a) SFOC, predicted versus measured (b) SFOC, model residuals

Figure 1: Model validation

3.4. Application of a real gas equation in high-pressure Diesel engine models


The high maximum pressures in marine Diesel engines, that can reach up
to 200 bar, challenge the validity of the application of the ideal gas law. Fig.
2 illustrates the difference in pressure between the application of the Redlich-
Kwong and the ideal gas equation for engine A, at 110%, 100%, 85%, and
50% engine load. The use of the ideal gas equation underestimates maximum

12
pressure by between 8.89% and 3,6% depending on the load. Hence, at these
pressure levels the use of a the Redlich-Kwong equation instead of the ideal gas
equation for determining gas state is necessary.

Figure 2: Cylinder pressure for different engine loads, ideal gas versus Redlich-
Kwong state equation

The use of the two different equations was studied in the calibration process.
2
Observing the evolution in the RP rediction statistics for engine efficiency for both
engines (0.82 and 0.98 with real gas equation implemented, 0.71 and 0.97 with
ideal gas equation implemented) leads to the conclusion that the use of the real
gas equation in the model improves its predictive performance.

3.5. Comparison of friction correlations


The evaluation of two existing correlations for the calculation of friction
losses was proposed and compared with the use of a weighting coefficient as
additional calibration parameter to the mode.
Results show clearly that the correlation proposed by Chen et al. is leading to
2
the highest prediction ability for the model. RP rediction for power calculated in
the three different cases (respectively using Chen et al. correlation, Winterbone
et al. correlation, and the weighting approach) is equal to 0.820, -0.916, and
0.708 for the A engine and 0.966, 0.232, and 0.573 for the B engine. It is clear
that not only is the Chen et al. correlation more appropriate than Winterbone
et al.’s, but also that allowing an additional calibration parameter for friction
leads to over-fitting the model and, therefore, to a reduced predictive capability.

13
(a) Case 1 (b) Case 2 (c) Case 3

Figure 3: Schematic representation of the different propulsion system arrange-


ments

4. Simulation results

The model was applied in the analysis the propulsion system as described in
the validation chapter. The ship under analysis is operated at constant engine
speed in order to use the shaft generator for auxiliary power generation. As the
shaft generator is connected to he main engine, this arrangement allows gener-
ating auxiliary power more efficiently than by using auxiliary engines. However,
propeller efficiency varies with its speed, and especially during slow operations
the optimal working condition is not located at the design propeller speed. For
this reason the possibility of generating electricity by using auxiliary engines
while running the engine at variable speed is studied. The option of installing
a WHR system on the exhaust gas, to be used for electric power generation, is
also evaluated. The three different concepts are shown in Fig. 3. In this case
the use of engine modeling is deemed necessary for the prediction of both engine
efficiency and exhaust gas temperatures and flows, as a standard regression of
engine shop test would not allow to make any prediction for performance at a
different speed than the design.
The potential for energy recovery was evaluated using the concept of exergy,
often used for WHR applications [42]. Exergy efficiency is then defined as the
ratio between the amount of exergy contained in the useful outputs of the system
(i.e. mechanical power) and the exergy input to the system, and it represents
how close the cycle gets to a Carnot engine. As exergy efficiency depends on the
specific cycle employed, the choice was made to calculate, for each condition, the
minimum exergy efficiency needed by the WHR system for producing the needed
amount of auxiliary power onboard. The value issued from this calculation can
give an estimation of the level of complexity needed by the recovery cycle in
order to fulfill its purpose.

14
Some additional values are needed for the calculation. The power need
(Pprop ) and the corresponding engine speed (n) are summarized in Table 4
at different ship speeds (vship ), both for constant engine speed and for the
condition corresponding to the highest propeller efficiency permitted by engine
safety limits. Subscripts f s and vs respectively refer to fixed and variable speed
operations.

Ship speed Pprop,f s n Pprop,vs nM E,vs


kn kW rpm kW rpm
11.0 2959 600 2959 600
11.5 3286 600 3072 551
12.0 3578 600 3281 499
12.5 3949 600 3734 521
13.0 4332 600 4188 543
13.5 4797 600 4725 569
14.0 5274 600 5263 595

Table 4: Propulsion power need for different ship speed for fixed and variable
engine speed

Efficiency for auxiliary generators (ηAG ) is assumed at 92%. Auxiliary en-


gines efficiency (bsf cAE ) at full load was available from technical documentation
(195 g/kWh), while their off-load behavior was assumed by similitude with the
main engine. The auxiliary power (Pel ) was assumed equal to 364 kW, which
includes 80% of sailing time. The process heat needed onboard is assumed to
account for 300 kW, which is generated using heat from the exhaust gas and
that consequently cannot be used for auxiliary power generation.
The three propulsion system arrangements are compared through their re-
spective ship specific fuel consumption (Φ), calculated as fuel consumed per
nautical mile traveled , a common evaluation figure of merit used for shipping
applications:
  
Pprop,f s Pel
φ1 = vs−1 bsf cM E + (15)
ηGB ηs ηGB ηSG
 
Pprop,vs Pel
φ2 = vs−1 bsf cM E + bsf cAE (16)
ηGB ηs ηAG
 
Pprop,vs
φ3 = vs−1 bsf cM E (17)
ηGB ηs
Results for Φ are shown in Fig. 4a. Relative Φ compared to the reference case
(Case 1) is shown in figure Fig. 4b. It is clear from the figures that the benefit
deriving from running the propeller at its optimal point largely overcomes the
loss due to the less efficient auxiliary power generation.
The advantage deriving from the use of WHR is also clear, but is limited to
ship operations on two engines. In this case, however, the minimum efficiency
required for fulfilling the required auxiliary power need is of 39%; according

15
(a) (b) Case 1

Figure 4: Comparison of ship specific fuel consumption

to Larsen et al.[43], the maximum exergy efficiency for organic Rankine cycles
(ORCs) can range between 53% and 66% depending on the level of hazard
related to the working fluid, therefore demonstrating that the required power
needs could easily be met even with very low complexity in the recovery cycle.

5. Conclusion

A zero-dimensional gray-box model for four-stroke medium speed marine


Diesel engines was derived and validated with acceptance tests records from a
MaK 832C and a MAN 9L27/38 engines. The predictive ability of the model
2
was tested using the RP rediction statistic for both engines, while for the MaK
engine an additional dataset was also available. Results from the validation
procedure indicate a very good ability in predicting both engine power output
and exhaust gas temperature.
The application of the Redlich-Kwong real gas equation instead of the ideal
gas equation was shown to be required due to the very high pressures in marine
engines. The influence of such modeling choice was tested and shows significants
improvements in power prediction. Similarly, two different friction correlations
were tested; the comparison showed that the regression proposed by Chen et
al. features better performance when applied to the modeling of large 4-stroke
marine Diesel engines.
The model was applied to the propulsion system of a merchant vessel with the
aim of evaluating three different arrangements influencing both the propulsion
and the auxiliary power generation onboard. The application of the model
showed that relying on the shaft generator in order to generate auxiliary power
does not bring any advantage to the overall fuel consumption as this choice
does not allow operating the propeller at its maximum efficiency. Moreover, the
evaluation of the WHR potential from the exhaust gas leads to the conclusion
that the standard auxiliary power need can be met by using a rather simple
recovery cycle (ηex = 0.32).

16
Acknowledgments

This work is part of a larger project about ship energy systems modeling,
performed at the Shipping and Marine Technology Department of Chalmers
University of Technology and financed by the Swedish Energy Agency. The
authors would also like to thank Bengt-Olof Petersen and Laurin Maritime for
their support and discussions. A special acknowledgment is also given to pro-
fessor Sven Andersson from the Department of Applied Mechanics in Chalmers
University of Technology for the productive discussions that lasted all over the
development of this work.

References

[1] J. T. Houghton, G. J. Jenkins, J. J. Ephraums, Climate change: the IPCC


scientific assessment, Cambridge University Press, Cambridge, UK, 1990.
[2] R. W. Bentley, Global oil and gas depletion: An overview, Energy Policy
30 (2002) 189–205.
[3] G. P. Theotokatos, A modelling approach for the overall ship propulsion
plant simulation, in: 6th WSEAS International Conference on System Sci-
ence and Simulation in Engineering, Venice, Italy, 2007, pp. 80–87.
[4] U. Campora, M. Figari, Numerical simulation of ship propulsion transients
and full-scale validation, Proceedings of the Institution of Mechanical En-
gineers Part M: Journal of Engineering for the Maritime Environment 217
(2003) 40–52.
[5] P. Schulten, D. Stapersma, A study of the validity of a complex simulation
model, Proceedings of the Institute of Marine Engineering, Science and
Technology Part A: Journal of Marine Engineering and Technology (2007)
67–77.
[6] A. Vrijdag, D. Stapersma, T. Van Terwisga, Systematic modelling, veri-
fication, calibration and validation of a ship propulsion simulation model,
Proceedings of the Institute of Marine Engineering, Science and Technology
Part A: Journal of Marine Engineering and Technology (2009) 3–20.
[7] G. G. Dimopoulos, N. M. P. Kakalis, An integrated modelling framework
for the design, operation and control of marine energy systems, Paper 154,
in: CIMAC Congress, Bergen, Norway, 2010.
[8] G. G. Dimopoulos, C. A. Georgopoulou, N. M. P. Kakalis, Modelling and
optimization of an integrated marine combined cycle system, in: Proceed-
ings of ECOS, Novi Sad, Serbia, 2011, pp. 1283–1298.
[9] H. T. Grimmelius, E. Mesbahi, P. Schulten, D. Stapersma, The use of Diesel
engine simulation models in ship propulsion plant design and operation,
Paper 227, in: CIMAC Congress, Wien, Austria, 2007.

17
[10] H. T. Grimmelius, E. J. Boonen, H. Nicolai, D. Stapersma, The integration
of mean value first principle Diesel engine models in dynamic waste heat and
cooling load analysis, Paper 280, in: CIMAC Congress, Bergen, Norway,
2010.
[11] J. N. Heywood, Internal Combustion Engine Fundamentals, McGraw-Hill,
New Yori, USA, 1988.
[12] E. G. Priotis, G. M. Kosmadakis, C. D. Rakopoulos, Comparative analy-
sis of three simulation models applied on a motored internal combustion
engine, Energy Conversion and Management 60 (2012) 45–55.
[13] S. Kumar, M. Kumar Chauhan, Varun, Numerical modeling of compression
ignition engine: A review, Renewable and Sustainable Energy Reviews 19
(2013) 517–530.
[14] R. G. Papagiannakis, D. T. Hountalas, Combustion and exhaust emission
characteristics of a dual fuel compression ignition engine operated with pilot
diesel fuel and natural gas, Energy Conversion and Management 45 (18-19)
(2004) 2971–2987.
[15] K. Kannan, M. Udayakumar, Modeling of nitric oxide formation in single
cylinder direct injection diesel engine using diesel-water emulsion, American
Journal of Applied Sciences 6 (7) (2009) 1313–1320.
[16] G. Kökkülünk, G. Gonca, V. Ayhan, I. Cesur, A. Parlak, Theoretical and
experimental investigation od diesel engine with steam injection system on
performance and emission parameters, Applied Thermal Engineering 54
(2013) 161–170.
[17] S. Awad, E. G. Varuvel, K. Loubar, M. Tazerout, Single zone combustion
modeling of biodiesel from wastes in diesel engine, Fuel.
[18] F. Scappin, S. H. Stefansson, F. Haglind, A. Andreasen, U. Larsen, Val-
idation of a zero-dimensional model for prediction of NO x and engine
performance for electronically controlled marine two-stroke diesel engines,
Applied Thermal Engineering 37 (2012) 344–352.
[19] G. Benvenuto, U. Campora, G. Carrera, P. Casoli, A two-zone diesel en-
gine model for the simulation of marine propulsion plant transients, in:
MARIND 98, Second International Conference on Marine Industry, Varna,
Bulgaria, 1998.
[20] G. A. Livanos, N. P. Kyrtatos, Friction model of a marine diesel engine
piston assembly, Tribology International 40 (2007) 1441–1453.
[21] S. N. Danov, A. K. Gupta, Modeling the performance characteristics of
diesel engine based combined-cycle power plants - Part I: Mathematical
model, Journal of Engineering for Gas Turbines and Power 126 (2004) 28–
34.

18
[22] R. Stone, Introduction to internal combustion engines, third edit Edition,
Palgrave MacMillan, London, UK, 1999.
[23] Y. Ding, D. Stapersma, H. Knoll, H. T. Grimmelius, Characterising heat
release in a diesel engine: A comparison between seiliger process and vibe
model, in: CIMAC Congress, Bergen, Norway, 2010.
[24] N. Miyamoto, T. Chikahisa, T. Murayama, R. Sawyer, Description and
analysis of a Diesel engine rate of combustion and performance using
Wiebe’s functions, SAE Paper 850107 (1985).
[25] C. D. Rakopoulos, E. G. Giakoumis, D. C. Rakopoulos, Study of the short-
term cylinder wall temperature oscillations during transient operation of a
turbocharged diesel engine with various insulation schemes, International
Journal of Engine Research 9 (2008) 177–193.
[26] H. Hiroyasu, Diesel engine combustion and its modeling, in: Proceedings of
1st International Symposium on Diagnostics and Modeling of Combustion
in internal Combustion Engines, Tokyo, 1985, pp. 53–75.
[27] M. V. Casey, T. M. Fesich, The efficiency of turbocharger compressors
with diabatic flows, Journal of Engineering for Gas Turbines and Power
132 (2010) 1–9.
[28] M. Lapuerta, R. Ballesteros, J. R. Agudelo, Effects of the gas state equation
on the thermodynamic diagnostic of diesel combustion, Applied Thermal
Engineering 26 (2006) 1492–1499.
[29] S. G. A. K. Danov, Influence of imperfections in working media on diesel en-
gine indicator process, Journal of Engineering for Gas Turbines and Power
123 (2001) 231–239.
[30] K. Wark, Advanced thermodynamics for engineers, McGraw Hill, New
York, USA, 1995.
[31] D. E. Richardson, Review of power cylinder friction for diesel engines, Jour-
nal of Engineering for Gas Turbines and Power 122 (2000) 506–519.
[32] S. K. Chen, P. Flynn, Development of a compression ignition research en-
gine, SAE Paper 650733 (1965).
[33] D. E. Winterbone, Transient performance, in: The Thermodynamics and
gas dynamics of internal combustion engines, Vol. II, Oxford University
Press, Oxford, UK, 1986.
[34] K. T. Yun, H. Cho, R. Luck, P. Mago, Modeling of reciprocating internal
combustion engines for power generation and heat recovery, Applied Energy
102 (2013) 327–335.
[35] P. Schulten, The interaction between diesel engines, ship and propellers
during maneuvering, Tech. rep., TU Delft.

19
[36] MAN, Project guides, L27-2 Marine, http://www.mandieselturbo.
com/0001425/Products/Marine-Engines-and-Systems/Medium-Speed/
GenSets/Project-Guides.html.
[37] MaK, M32c project guide, propulsion, http://marine.cat.com/cda/
files/953438/7/.

[38] J. H. Holland, Genetic algorithms, Scientific American 267 (1992) 66–72.


[39] A. Chepperfield, P. Flemming, H. Pohlheim, C. Fonseca, Genetic algorithm
toolbox: for use with matlab, University of Sheffield - Department of Au-
tomatic Control and Systems Engineering.

[40] D. C. Montgomery, Design and Analysis of Experiments, 8th Edition, John


Wiley & Sons, Singapore, 2009.
[41] W. Shi, H. T. Grimmelius, D. Stapersma, Analysis of ship propulsion sys-
tem behaviour and the impact on fuel consumption, International Ship-
building Progress 57 (2010) 35–64.

[42] I. Dincer, M. Rosen, Exergy, 2nd Edition, Elsevier Ltd, 2013.


[43] U. Larsen, L. Pierobon, F. Haglind, C. Gabrielii, Design and optimisation
of organic rankine cycles for waste heat recovery in marine applications
using the principles of natural selection, Energydoi:http://dx.doi.org/
10.1016/j.energy.2013.03.021.

20
Paper III
Low Carbon Shipping Conference, London 2013

The influence of propulsion system design on the carbon


footprint of different marine fuels

Francesco Baldi*a, Selma Bengtssona, Karin Anderssona


ªShipping and Marine Technology, Chalmers University of Technology, 412 96 Gothenburg, Sweden

Abstract

Rising environmental awareness and stricter environmental regulations have increased the interest in new fuels and
energy efficiency measures in the shipping industry. Different fuels have different physical and chemical
properties that affect the performance of internal combustion engines, e.g. the efficiency, the exhaust gas
emissions, and the potential for energy recovery. This has an impact on the potential propulsion efficiency as well
as on the life cycle environmental performance. The aim of this study is therefore twofold. First, to assess the
potential for optimising the energy use of the propulsion system dependent on fuel choice and second, to assess the
overall life cycle global warming potential of the optimised systems. Three fuels are compared, heavy fuel oil
(HFO), marine gas oil (MGO), and liquefied natural gas (LNG), in combination with two exhaust gas cleaning
technologies, scrubbers and selective catalytic reduction (SCR) units. Data from one year of actual operation with
a product tanker are used as a base for the optimization. The results show that the solution with the lowest fuel
consumption and carbon footprint is a two-stroke engine with waste heat recovery (WHR) powered by LNG. The
synthesis of an optimization procedure for the propulsion system and an LCA approach leads to very interesting
results. The different carbon content of different fuels, together with methane slip, leads to a better estimation to
the carbon footprint of different propulsion systems. On the other hand, a better insight of the differences between
different propulsion arrangements allows performing a more accurate comparison between different fuels. The
potential for WHR has a particularly relevant influence on the final result.

Keywords: Propulsion system optimization, shipping, carbon footprint, marine fuels

1. Introduction
The shipping sector is going to face, in the coming years, a number of decisive challenges, which are
expected to modify ship design as it has not happened in the last decades. Rising bunker prices and low
freight rates demand for increasing fuel efficiency; rising global warming awareness is starting to also
affect the shipping sector, which despite its high transportation efficiency has become a significant
contributor of greenhouse gases (Buhaug et al., 2009); finally, new environmental regulations, mainly
related to the emissions of sulphur and nitrogen oxides (IMO, 2013a), will add complexity to the
design and operation of merchant vessels.

1.1 Background
In the preliminary phase of ship design, several different choices have to be made. In particular,
decisions connected to the propulsion system will strongly affect the carbon footprint of the ship, and it
is therefore of primary importance not to underestimate this phase. The selection of the fuel, the type of
engine, and the propulsion arrangement should be performed with maximum care.
The use of computer simulations for the identification of optimum designs has a long history of
application in several fields of knowledge, not last that of engineering. In naval architecture,
applications of computer models for energy systems can be dated as back as 1984 (DeTolla and
Fleming, 1984), while the work from Dimopoulos et al. (2011) is just a more recent example of
optimization procedures applied to ship propulsion systems. However, two aspects are very seldom
treated in such scientific publications: the accounting of the operational profile, and the selection of the
prime mover.

*
Corresponding author. Tel: +46-31-772-2615
Email address: francesco.baldi@chalmers.se

1
Low Carbon Shipping Conference, London 2013

As stressed by Motley et al., traditionally propulsion systems are designed to optimize performance at a
single or few design points (Motley et al., 2012), while the importance of taking the whole operational
profile into account was already highlighted by Dupuis and Neilson (1997). Only few examples of
different approach are available in scientific literature; Dimopoulos et al. propose the optimization of a
combined cycle for marine applications, subdividing ship operations in four well-defined conditions
(Dimopoulos et al., 2011); Doulgeris et al. propose the optimization of a propulsion system considering
a reference journey, with varying environmental conditions (Doulgeris et al., 2012); Sciberras and
Norman, in their work about multi-objective optimization of hybrid propulsion systems refer to the use
of extensive data on the operational profile; Motley et al. (2012) also propose a tentative for
operational lifetime optimization of a propulsion system, focusing on propeller design.
In the same line, the choice of the prime mover is often left outside of the scope of the work
(Dimopoulos et al., 2011, Motley et al., 2012); Doulgeris et al. propose the choice between two
alternative gas turbines ((Doulgeris et al., 2012); Sciberras and Norman, instead, use an extensive
component database (Sciberras and Norman, 2012); this latter method is applied in the present study.

From the technical point of view, when compared to traditional design practices, two new subjects are
today on the agenda for addressing fuel consumption and environmental concerns: fuel choice and
waste heat recovery (WHR) systems installation.

As a consequence of new environmental regulations and high fuel prices, the choice of heavy fuel oil
(HFO) as ship fuel is not anymore to be given for granted. Emission Control Area (ECA) emission
limits can be met either with additional machinery on board (scrubbers, catalysts), or by employing a
different fuel, such as marine gas oil (MGO), for compliance with SOX emissions, or liquefied natural
gas (LNG), for compliance with both SOX and NOX emissions.

On the other hand, WHR systems are getting largely on the agenda of shipyards and naval architects
for the design of new ships; several installations can already be observed on-board existing vessels, and
there are expectations for the number to grow rapidly in the future. WHR is a well-known technology
which is extensively employed in land-based applications and it allows significant improvements in the
efficiency of Diesel engines with using a proven technology and having a rather short payback
time(Dimopoulos et al., 2011). It should be noted, however, that WHR systems do not come without
any drawback. Lowering exhaust gas outlet temperature can lead to condensation of particles and
sulphuric acids on the heat exchangers, thus leading to corrosion. Additional space and weight for the
exhaust boiler, turbine, and condenser need to be allocated. Furthermore, the interaction of WHR
systems with abatement technologies can lead to complex configurations. The implications of these
aspects in the modelling will be discussed in sections 2.1 and 2.2.

The evaluation of different design alternatives is often performed on the basis of energetic or economic
principles. However, it is in the beliefs of the authors that environmental impact should be a major
parameter to take into account in order to compare design alternatives, especially when the use of
different fuels involves diversity both in direct and indirect carbon emissions. In order to evaluate the
environmental performance of possible marine fuels life cycle assessment (LCA) can be used. LCA is a
method where the environmental performance of products or services throughout the life cycle of the
product or service is assessed, ideally including all processes in the life cycle from raw material
acquisition to end-of-life disposal.

Several LCAs of marine fuels have been published previously (Corbett and Winebrake, 2008,
Winebrake et al., 2007, Bengtsson et al., 2011, Bengtsson et al., 2012, Bengtsson et al., 2013). Corbett
and Winebrake (2008) and Winebrake et al. (2007) considered a specific trip with different engine load
for different parts of the trip, while Bengtsson et al. (2011, 2013) only considered operation at sea with
85% engine load. Furthermore, Bengtsson et al. (2012) considered the energy use during one year of
operation. The engine efficiency was assumed to be the same independently on fuel used in all studies.
Nor did any of these studies try to optimize the propulsion system with regard to the fuel. If one fuel
has a larger potential for gaining higher propulsion system efficiency than the others, this could have
an impact on its environmental performance and should ideally be considered in an LCA.

2
Low Carbon Shipping Conference, London 2013

1.2 Aim
The aim of this paper is twofold. From one perspective, a procedure is proposed for the optimization of
the choice of the propulsion system with regard to energy efficiency, taking into account the
environmental constraints, and with respect to the choice of fuel, engine type, and possible application
of WHR systems. To add one more dimension to the work the overall life cycle global warming
potential (GWP) of the optimised systems is compared using an LCA approach, i.e. the carbon
footprint.

2. Method
The method here proposed for propulsion system optimization and GWP assessment is supposed to be
applicable to ships of any kind and size. However, it will be presented applied to a specific case-study
ship, as to make it easier to show its principles and potential. The case study of a product tanker with a
deadweight of 45 000 tons will be used. In order to apply the propulsion system design optimization
based on real operational data instead of a fixed design point, measured data from one year of
operations of a similar ship were used in order to represent the expected ship behaviour.

Data on propulsion and auxiliary power needs in kW are available. The inclusion of propeller design in
the optimization process would involve additional complexity, and is left for future studies. The
implications of the use of different propeller types and diameters, as well as the complex issues of
engine-propeller coupling will therefore not be considered here. The use of a controllable pitch
propeller, associated with a gearbox and a shaft generator for auxiliary power generation is assumed in
every case, even if this corresponds to an important limitation in the process. The main engine load
used in the calculations is presented in Figure 1. Three main operational patterns can be identified:
high-load operation (peak at 90% load), when sailing at maximum speed, low-load operation (peak at
55% load), when sailing at economical speed, and port operations (peak at 20% load), when unloading
the cargo

Figure 1. Frequency of occurrence at different engine loads

2.1 Environmental implications


The ship is supposed to to sail into ECA areas. Hence, the chosen design must be able to fulfil ECA
areas requirements, i.e. maximum of 0.1% sulphur in the fuel starting from 1st of January 2015 and Tier
III NOX requirements for new buildings starting from 1st of January 2016.†

Sulphur emission limits require additional technical arrangements depending on the fuel choice. MGO
and LNG have sulphur content below the level required in ECAs. HFO, however, has much higher
sulphur content in the fuel (average 2.51 wt. % in 2012 (IMO, 2013b)) that has to be taken care of. The
use of a scrubber is assumed in this case, which allows for even high sulphur heavy fuels to respect
ECA limits. The electricity needed for pumps and the backpressure imposed on the engine are


This regulation is under review by the International Maritime Organization (IMO).

3
Low Carbon Shipping Conference, London 2013

estimated to generate a fuel penalty between 1-3% of the engine capacity (Kjølholt et al., 2012). A 2%
fuel penalty is added for scrubbers in this study.

Limitations on NOX emissions require the installation of a SCR unit, apart from the case of 4-stroke
LNG powered engines. While little literature is published on the influence of SCRs on engine
performance, manufacturers often claim the absence of any fuel penalty (Hagström, 2013); no
increased fuel consumption connected to the use of SCRs for NOX emissions reduction is therefore
considered in this study. Furthermore, the effect on exhaust gas temperature is considered to be
negligible. The carbon footprint implications of SCRs (direct CO2 emissions and the GHG impact of
urea production and transportation) will be further explored in section 2.4.

2.2 Propulsion system optimization


This paper’s first aim is that of describing a methodology to optimize the design of a propulsion system
in the early stages of the design. A number of different variable design choices are proposed and
analysed separately.

The design choices concern:


1. The number and type of engines: starting from the experience of today existing installation,
two possible arrangements were selected:
 One 2-stroke engine
 One or two 4-stroke engines
2. The availability of a waste heat recovery system
3. The primary fuel used for propulsion:
 Heavy fuel oil (HFO)
 Marine gas oil (MGO)
 Liquefied natural gas (LNG)

The combination of all the three possible design variables generates 12 possible cases. However, the
case of 2-stroke engine, HFO fuelled, equipped with WHR system is not considered because of the
technical difficulties in providing the good combination of high temperatures and sulphur concentration
in the exhaust required by SCR operations (Magnusson et al., 2012). This leaves a total of 11 different
cases to be separately optimized and compared.

The authors have gathered a large dataset from the documentation of the main manufacturers active on
the market of Diesel engines for marine propulsion. Data from each engine referred to engine power,
efficiency, and exhaust gas properties (temperature and mass flow) for a number of operational points.
For 2-stroke engines, data was available for a large number of operative points; hence, spline curves
were deemed the most suitable regression technique. For 4-stroke engines data were available only for
3-5 operative points, and second degree polynomial regressions were used instead. For all types of
engines the extrapolation of engine performances outside of the original load range was performed
using a linear approximation, in order to avoid numerical “explosion” of efficiencies, temperatures, and
flows, particularly at very low loads. Efficiency for all engines was corrected in order to account for
the typical 5% tolerance allowed under ISO conditions; furthermore, ISO corrections for the addition
of engine-driven pumps and for more realistic ambient conditions were applied.

The range of total installed power was set between 7500 and 10000 kW, where the original installed
power was equal to 7700 kW. This limited the amount of possible propulsion trains to:
 18 possibilities for 2-stroke engines
 14 possibilities for 2-stroke dual fuel engines
 143 possibilities for 4-stroke engines
 8 possibilities for 4-stroke dual fuel engines

The relatively low number of total runs (183 runs) allowed testing all possible configurations, therefore
avoiding any numerical uncertainty in the results. The computational time for the overall evaluation
process is 30 seconds on an Intel(R) Core(TM) i7-2620M CPU at 2.7 GHz processor.

The WHR system is accounted for as a net, free contribution to on-board power production. Exhaust
gas temperature and flow are interpolated as a function of load for each engine starting from available

4
Low Carbon Shipping Conference, London 2013

data, thus allowing calculating the energy content in the exhaust, which is considered in this work to be
the input to the WHR system. Despite the existence of several proposals for WHR systems also
harvesting other sources of waste heat, such as charge air cooling (Teng et al., 2007), only the exhaust
will be considered in this study.

In order to account for variations in energy quality connected to exhaust temperature, the concepts of
exergy is used. Exhaust exergy flows are evaluated, and the power available from WHR system is
finally calculated assuming a fixed value for the recovery cycle’s exergy efficiency. This approach
allows making an assumption of how much the system approaches the ideal, which can be considered a
constant for all propulsion system scenarios (Dincer and Rosen, 2013). Values up to 60% can be
reached by optimized ORCs (Larsen et al., 2013). However, for the present study, a more traditional
single pressure steam based cycle is assumed, as it represents more closely the state-of-the-art in the
shipping industry. A reference value of 40% exergy efficiency, consistent with the design proposed by
Theotokatos and Livanos (2013), will be used. The implications of using a WHR system in the design
are strongly connected to both the choice of the prime mover (2-stroke or 4-stroke engines) and of the
fuel.

The implications of fuel choice are mostly related to the environmental impact of the propulsion
system, and in particular with its compliance to regulations on the emissions of NOX and SOX. Fuel
choice also has a strong influence on the efficiency of the propulsion system when the installation of a
WHR system is taken into account. A non-negligible amount of sulphur is contained in HFO and
MGO. For this reason, sulphur oxides and, as a consequence, sulphuric acid will still be generated after
the combustion process. In order to prevent sulphuric acid condensation on the heat exchangers and on
the walls of the funnel exhaust gas temperature is not allowed to drop below 150 °C. This limit can be
discarded when LNG is used as fuel, as it usually contains less than 10ppm sulphur by mass. The
handling of high-viscosity HFO also adds to an additional heat requirement, which was assumed equal
to 300 kW regardless the load of the engine, and needs to be produced using exhaust gas heat.

2.3 Analysis of the GHG potential of the optimized designs: an LCA approach
It is important to consider the whole life cycle when comparing the carbon footprint of the optimised
designs. Carbon footprint is a limited life cycle assessment considering only the emissions of
greenhouse gases in the life cycle (Weidema et al., 2008). The carbon footprint of the optimised
designs will be compared by calculating the global warming potential at a 100 year time perspective
(GWP), considering emissions of CO2, CH4 (1 g CH4=25 g CO2-equvivalent) and N2O (1 g N2O=298 g
CO2-equivalent) (IPCC, 2007). The whole fuel life cycle is considered starting from raw material
acquisition, i.e. crude oil or natural gas, followed by fuel production, distribution and finally
combustion in the marine engines. However, the manufacturing, maintenance and scrapping of the
vessel is not included. Nor is the manufacturing of the exhaust abatement technologies. The
greenhouse emissions and energy use during the manufacturing of the ship has been shown to be very
low, less than 3%, when considering the life cycle of the vessel according to Johnsen and Magerholm-
Fet (1998) and Walsh and Bows (2011).

Data for the raw material acquisition, fuel production and fuel distribution for the fuels are from
Bengtsson et al. (2011). These data are representative for shipping in the northern part of Europe. Two
distribution alternatives for LNG are included in Bengtsson et al. (2011), i.e. from the North Sea or
from Qatar. Data for natural gas transported from the North Sea is considered in this comparison.

Ammonia is used as a reducing agent in the SCR, normally supplied by a water solution of urea. The
data for urea production and distribution are taken from Andersson and Winnes (2011). The emission
of CO2 from the use of urea is included in the data for production of urea (Davis and Haglund, 1999).
The estimates for the needed amount of urea is calculated based on data for the NOX emissions from
existing engines (Bäckström, 2010, NTM, 2008) and the amount of urea needed to reduce the NOX
emissions to the Tier III level Is calculated based on the assumption that in order to reduce the NOX
emissions with 2 moles, 1 mole of urea needs to be added to the SCR unit.

The CO2 emissions from the fuel combustion are calculated from the carbon content in the fuel
assuming that 99.5% of the fuel forms carbon dioxide. Fuel properties for MGO and HFO are average
data reported in scientific journals (Fridell et al., 2008, Cooper, 2003, Cooper, 2005, Winnes and
Fridell, 2009, Winnes and Fridell, 2010), while the fuel properties for Norwegian LNG are from

5
Low Carbon Shipping Conference, London 2013

Edwards et al. (2011). The emissions of methane are taken from NTM (2008) for MGO and HFO (both
for 4-stroke and 2-stroke engines), from Bäckström (2010) for 2-stroke LNG engines, and from
Wärtsilä published by Bengtsson et al. (2011) and from Nielsen and Stenersen (2010) for 4-stroke LNG
engines . The methane emissions from the dual fuel engines are uncertain and very high emissions, i.e.
averaging 8 wt. % of the fuel, is reported for the first generation of 4-stroke dual fuel engines (Nielsen
and Stenersen, 2010). A 4 wt. % methane slip for the 4-stroke dual fuel engine is considered as a base
in the calculations. The emissions of nitrous oxide for the existing engines are taken from Cooper and
Gustafsson (2004). Information regarding emissions of nitrous oxide from the gas engines has not been
found. The data used is presented in Table 1.

Table 1. Data used for the carbon footprint calculations.

HFO (1 MJ) MGO (1 MJ) LNG (1 MJ) Urea (1 g)


Well-to-tank
Primary energy use (MJ) 1.09 1.16 1.1 0.03
Emissions of CO2 (g) 6.68a 7.02b 6.97 1.99
Emissions of CH4 (g) 0.073 0.078 0.046 0.0021
Emissions of N2O (g) 0.0002 0.0002 0.0036 -
Tank-to-propeller
Urea consumption (2-stroke) (g) 1.19 1.10 1.06 -
Urea consumption (4-stroke) (g) 0.835 0.835 0 -
Emissions of CO2 (g) 77.2 73.1 54.4 -
Emissions of CH4 (2-stroke) (g) 0.000756 0.000744 0.04609 -
Emissions of CH4 (4-stroke) (g) 0.000463 0.000465 0.816c -
Emissions of N2O (2-stroke) (g) 0.00388 0.0039 - -
Emissions of N2O (4-stroke) (g) 0.00352 0.00352 - -
a
The open-loop sea water scrubbers discharge the scrubber water in the open sea, thereby indirectly releasing CO2 to the
atmosphere. Approximately 2 moles of CO 2 is formed for every mole of SO2 released. This would increase the CO2 emissions
from use of scrubbers with approximately 1.5 g/MJ HFO combusted and is not included in the LCA. b Increased energy use in
the refineries for producing MGO after 2015 is not included in Bengtsson et al. (2011). However, the CO 2 emissions are
estimated to increase with approximately 7 g/MJ MGO based on Avis and Birch (2009), when allocating the total increase in
CO2 to the production of MGO. b This parameter have been varied between 0.28 g/MJ and 1.63 g/MJ representing between 1.4
and 8 wt. % methane slip.

3. Results
Here follow the results for the propulsion system optimization as well as the carbon footprint
calculations.

3.1 Propulsion systems


A total of 183 design alternatives were evaluated and compared. In order to validate the results, the
modelled consumption was tested versus the measured one using the data corresponding to the
arrangement of the original ship. This showed reasonable accuracy, as calculated consumption differs
from measured consumption by 5%, which could be explained by engine wear and non-optimal
operating conditions. Moreover, measurement accuracy derived from fuel meter and shaft power meter
is equal to around 5%, meaning that the two results are not significantly different.

Results for the total fuel consumption over one year of operation (valid only for the main engines) are
shown in Figure 2, displayed as TJ/year. Values are given for each fuel choice, and for each type of
engine selected. Two values are showed for each combination, referring to the case with and without
the installation of WHR systems, apart from the case of 2-stroke propulsion arrangement with HFO as
a fuel, where technical limitations do not allow the application of WHR. The horizontal red line refers
to the original arrangement for the selected ship (hereafter referred to as “reference case”). For each
scenario “range bars” are also displayed, representing the average, maximum and minimum results for
each possible arrangement. It should be noted however that only the best point was used for the further
analysis.

The best selected arrangement for all 2-stroke cases coincided with the most efficient engine at design
point, as is also showed in Table 2. In the case of 4-stroke arrangements, instead, a father-and-son
arrangement proved to be the best efficient in all cases, with a large, efficient engine (rated power of
7200 kW in HFO and MGO cases, 5850 kW in the LNG case) supported by a smaller engine (1020 kW

6
Low Carbon Shipping Conference, London 2013

and 2700 kW) at low and peak loads. The benefit of operating in off-design conditions for shorter time
in the case of 4-stroke, dual fuel engines is justified by the lower efficiency of these engines at low-
load operations, as they operate according to a Otto-cycle (Aesoy et al., 2011).

Figure 2. Energy consumption from one year of operation for the presented cases.

One of the aims of this paper was that of estimating the relevance of different details related to the
design of the propulsion system when evaluating the carbon footprint of the selected ship, both in
absolute and relative terms for comparison between fuels and arrangements. In order to compare the
influence of the different aspects, Table 2 shows, for each case, the modifications observable in the
yearly energy consumption when adding more detail to the modelling, where:
 Level 1: Constant engine load, constant engine efficiency assumed based on existing data
(engine load fixed at 75% according to EEDI calculation; engine efficiency assumed from
Buhaug et al. (2009) equal to 170 g/kWh for 2-stroke engines and 190 g/kWh for 4-stroke
engines).
 Level 2: Real operational load, constant engine efficiency based on the most efficient
arrangement at design conditions.
 Level 3: Real operational load, load-dependent engine efficiency based on the most efficient
arrangement at design conditions.
 Level 4: Real operational load, load-dependent engine efficiency based on the most efficient
arrangement taking the operational cycle into account.
 Level 5: Real operational load, load-dependent engine efficiency based on the most efficient
arrangement taking the operational cycle into account, with WHR.

Table 2. Energy consumption in TJ for one year of operation and for different levels of modeling detail.

2-stroke engine 4-stroke engine


HFO MGO LNG HFO MGO LNG
Level 1 223 249
Level 2 151(-33%) 148(-34%) 156(-30%) 168(-33%) 165(-34%) 164(-34%)
Level 3 166(9.9%) 163(10.1%) 164(5.1%) 178(6.0%) 175(6.1%) 179(9.1%)
Level 4 166(0.00%) 163(0.00%) 164(0.00%) 177(-0.56%) 174(-0.57%) 176(-1.7%)
Level 5 162(-2.4%) 157(-3.7%) 150(-8.0%) 164(-7.3%) 159(-8.6%) 156(-11%)

7
Low Carbon Shipping Conference, London 2013

3.2 Carbon footprint


The carbon footprints of the optimised systems are presented in Figure 2. The case with the overall
lowest carbon footprint is LNG with 2-stroke engines combined with WHR. The 4-stroke dual fuel
engines have a significant higher carbon footprint caused by a large methane slip from the engine
during combustion. The 4-stroke dual fuel engine cases vary from having the third and fourth lowest
carbon footprint to the highest dependent on the estimates of the methane slip.

The largest contribution to the carbon footprint originates from the combustion of fuels in the marine
engines (between 84-88%). Acquisition of raw material, fuel production and distribution stands for
between 10-13% while urea production and transportation stands for between 0-3% of the carbon
footprint.

Figure 3. The carbon footprint of the optimised systems divided on the contributing greenhouse gases
and the different phases in the life cycle. The range for the 4-stroke dual fuel engines represents the
difference between a methane slip of 1.4 and 8 wt. % during combustion of LNG. Results with and
without WHR are presented for all alternative except for the alternative with HFO and 2-stroke
engines. The second bar for each engine arrangement represents the system with WHR.

Table 1Table 3 gives an overview of the result comparing the propulsion system energy use and the
carbon footprint for the selected cases compared to the reference case, representing the modelled
energy use with today’s propulsion system on the product tanker. The propulsion system energy use is
reduced by 28% while the carbon footprint is reduced by 36% for the case with a 2-stroke LNG engine
with WHR.

Table 3. Comparison of the propulsion energy use and the carbon footprint for the selected cases. The difference
compared to the reference case (182 TJ) is presented in parenthesis.

Propulsion energy use (TJ/year) Carbon footprint (tonne CO2-eq./y)


HFO 2-stroke 166 (91%) 14800 (91%)
HFO 4-stroke 177 (97%) 15700 (97%)
HFO 4-stroke with WHR 164 (90%) 14500 (89%)
MGO 2-stroke 163 (90%) 13900 (86%)
MGO 2-stroke with WHR 157 (86%) 13400 (82%)
MGO 4-stroke 174 (96%) 14700 (90%)
MGO 4-stroke with WHR 159 (87%) 13500 (83%)
LNG 2-stroke 164 (90%) 11300 (70%)
LNG 2-stroke with WHR 150 (82%) 10400 (64%)
LNG 4-stroke 176 (97%) 14800 (91%)
LNG 4-stroke with WHR 156 (86%) 13100 (81%)

4. Discussion
The discussion is divided in three parts discussion on the results, on the method and on potential future
work.

8
Low Carbon Shipping Conference, London 2013

4.1 Result discussion


The analysis of Figure 1 shows that 2-stroke dual fuel engines are the most efficient solution from a
purely energetic perspective. The 4-stroke engine solution (which is also the current arrangement for
the existing ship) is a competitive alternative only when a WHR system is in place, as 4-stroke engines
generally feature higher exhaust temperature, when only direct fuel consumption is accounted for. The
observation of the results for the carbon footprint, however, has a strong influence on the evaluation.
On the one hand, the lower carbon content in the fuel and the very low sulphur content makes 2-stroke
LNG-powered solutions even more convenient (carbon footprint is reduced by 36% compared to the
reference case when WHR is used). On the other hand, taking the methane slip into account makes the
LNG 4-stroke solution much less viable, even if this result is strongly affected by the uncertainty in the
data.

The observation of the range bars shows the size of the variation in the results, thus confirming the
additional value of implementing a propulsion system optimization algorithm in order to evaluate the
overall yearly consumption. The variation between the two extremes of the range varies between 6%
and 10% depending on the case, while the standard deviation lies between 1.5% and 3%. The
utilization of a complete operational profile for the evaluation of the propulsion arrangement instead of
a single operational point only has an influence when a multi-engine arrangement is employed. In fact,
for all 4-stroke engine cases, multi-engine solutions prove being the most efficient, regardless the fuel.
The advantage compared to the use of the most efficient engine alone is rather small in the HFO and
MGO cases (0.7% improvement), but larger in the LNG case (1.7% improvement). The presence of
WHR has no influence on the choice of the engine, regardless the case. The small amount of power
produced by the WHR system does not justify the choice of an engine with lower efficiency but, for
instance, higher exhaust temperature. However, as observable in Figure 1, taking WHR systems into
account can significantly decrease the overall consumption (between 2.4% and 6.7%, with the assumed
recovery efficiency).

The propulsion system optimization has shown to have a significant impact on the life cycle GWP. The
propulsion energy needed spans between 150-177 TJ per year depending on the selected designs. The
increased potential for WHR for LNG compared to HFO and MGO is important and should be
considered when evaluating the environmental performance of LNG for future marine propulsion. The
carbon footprint when using 2-stroke or 4-stoke dual fuel engines was significantly different, showing
that the propulsion system should not be neglected when calculating the carbon footprint.

From an engine optimization perspective, this approach provides additional insight about the climate
impact of the propulsion system. It is here shown that the energy use and the carbon footprint are not
directly proportional, as the carbon content is different between the different fuels and the non-CO2
emissions contributing to global warming are different dependent on the different propulsion systems,
especially in the case of dual fuel engines. From an LCA perspective it is possible to gain insight of the
potential for energy efficiency improvement for the different fuels, e.g. the LNG propulsion systems
can extract more energy from a WHR system due to the very low sulphur content. We suggest that this
approach could contribute to a more complete understanding of the propulsion systems available and
their pros and cons. This approach could therefore be used to provide decision support when selecting
propulsion systems for new vessels.

4.2 Method discussion


The implications of the results presented in this paper would not be complete without a discussion of
the methodology employed in this paper, and specifically of the different assumptions that had to be
made.

Data for different engines were obtained from publicly available technical documents from different
engine manufacturers. Data for 2-stroke engines from MAN were very extensive, with detailed fuel
consumption, exhaust mass flow and temperature for 14 different load points. The same cannot be said,
unfortunately, for data concerning 4-stroke engines, obtained from Wärtsilä and MaK: in these cases,
in fact, data points were available for only 3 to 5 different loads, which involves less accurate
approximations in the mathematical regression. For all engines, in addition, the variations generated by
the possibility of influencing cylinder loading, valve timing, and turbocharger matching were

9
Low Carbon Shipping Conference, London 2013

neglected. The availability of better and more extensive data would definitely improve the quality of
the approximation.

The installation of the SCR was also taken into account, with no fuel penalty. There is some
controversy on this point, as manufacturers claim no fuel penalty (Hagström, 2013) while some tests
have shown the opposite (Subramaniam et al., 2011). Moreover, the presence of an SCR would allow a
much more efficiency-driven engine tuning, as Tier II engines normally comply with NO X emissions
regulations by reducing combustion pressures and temperatures and, thus, engine thermodynamic
efficiency. No study is known to the authors that could show the effect of the combination of the two
different phenomena. Similarly, the approximation employed in this study in order to account for the
scrubber power needs is quite strong, and, in presence of more extensive data, could be further
improved.

The result of an LCA study depends on the goal and scope of the study and the data used in this paper
are specific for operation in the northern part of Europe considering marine transportation in 2015-
2020. Ideally, the data used in the LCA should be specific for the region and the time in which the new
ship will operate. In this study the geographical operation and time scale are not defined. However, the
part of the life cycle that contributes the most to the carbon footprint is the combustion of fuel in the
tank-to-propeller phase and this data is specific for the different types of engines used. The second
biggest contributor to the carbon footprint is the fuel production phase. This data may vary between
different production facilities, but still when considering future operation it can be expected that state-
of-the-art production today independent on geographical scope can be representative.

It is highlighted that the methane slip from the dual fuel engines has a significant impact on the carbon
footprint. The uncertainty on this value is very high and needs to be verified with actual on-board
measurements during operation. The only data available today are from engine manufacturers and a
few measurements on vessels operating in Norway (Nielsen and Stenersen, 2010). 2-stoke dual fuel
engines are operating according to the diesel process and have a much lower methane slip than the 4-
stroke lean burn dual fuel engines included in this analysis. It is also possible to use 4-stroke dual fuel
engines operating according to diesel process; this would result in lower methane slip and higher NOX
emissions, which is the case for the 2-stroke LNG engines. Another possibility may also be to use
oxidation catalysts in order to reduce the CH4 emissions. Furthermore, there are no emission factors for
nitrous oxides for the dual fuel engines.

4.3 Future work


This study opens for a number of alternatives for future work.

As mentioned in the description of the method employed in this study, leaving the propeller out of the
optimization procedure constitutes a very strong approximation, which is hardly justifiable in practical
terms. In particular, 2-stroke engines performance would then be even higher, as no gearbox losses are
involved in a directly coupled arrangement between engine and propeller. Moreover, engine speed
should then be taken into account, as it strongly influences propeller efficiency, and vice versa. As an
additional consequence of this simplification, auxiliary needs (though rather limited in the case study
taken into account), are assumed to be a part of the main engine load, as the current arrangement is
provided with a shaft generator. This is definitely a viable alternative, but accounting of different
possible designs for auxiliary power generation would be an improvement to the quality of the results.

The methane emission from the combustion of natural gas in marine engines is shown to have a great
impact on the overall result. The methane slip is markedly load-dependent and could potentially be as
high as 15 wt. % at low engine loads (Nielsen and Stenersen, 2010). Hence, the availability of load-
dependent methane slip emissions would be an additional improvement to the calculation of the
greenhouse gas emissions from combustion.

The approach of combining a propulsion system optimization with carbon footprint calculations gives
information on two important aspects when selecting the propulsion system in the design of a new ship.
It would, however, be possible to combine the engine optimization with a full LCA and not only the
carbon footprint and thereby providing information about a number of different environmental impacts.

10
Low Carbon Shipping Conference, London 2013

As have been mentioned before a number of LCA studies on marine fuels have been published and
data from these could rather easily be combined with data from engine optimizations. When
performing a complete LCA there are more emissions from combustion that could be calculated based
one the operational profile of the ship, e.g. nitrogen oxides, carbon monoxide and particles; thus,
making the result even more accurate.

When making a decision of which propulsion system to choose there are more parameters that are
important to consider then only the energy use and the environmental impacts, e.g. the cost of the
system, the reliability and ease to operate, the knowledge about the system and so on. This could, for
example, be considered in a multi criteria decision analysis.

5. Conclusions
In this paper we have presented a combination of a design optimization methodology with a carbon
footprint analysis which could be used as decision support in the design process for new ships. The
most efficient propulsion system was selected for a number of different engine types and fuels, taking
into account a real operational pattern, referred to one year of operation of a similar existing ship. This
procedure showed to improve the efficiency of the design when 4-stroke engines are employed,
especially with LNG as fuel.

The 2-stroke dual fuel engine, equipped with a WHR system, proved to be the arrangement with the
lowest carbon footprint. The WHR system proved to significantly improve the overall performance.
The 4-stroke dual fuel engine solution, even though showing promising performance from an energy
perspective, loses appeal when methane slip is also taken into account. When compared to traditional
LCA studies, this work showed the importance of the selection of the propulsion system, as compared
to the assumption of a common efficiency for all engines, in particular when off-design performance is
taken into account.

Acknowledgments
This work is the synthesis between two distinct projects, “life cycle assessment of marine fuels” and
“energy systems modelling in shipping”. The authors want therefore to acknowledge the contribution
of the financers of the studies, respectively Vinnova and the Swedish Energy Agency.

References

AESOY, V., MAGNE EINANG, P., STENERSEN, D., HENNIE, E. & VALBERG, I. 2011. LNG-Fuelled
Engines and Fuel Systems for Medium-Speed Engines in Maritime Applications. SAE International.
ANDERSSON, K. & WINNES, H. 2011. Environmental trade-offs in nitrogen oxide removal from ship engine
exhausts Proceedings of the Institution of Mechanical Engineers, Part M: Journal of Engineering for the
Maritime Environment 225, 33-42.
AVIS, M. J. & BIRCH, C. H. 2009. Impacts on the EU refining industry & markets of IMO specification changes
& other measures to reduce the sulphur content of certain fuels. Purvin & Gerts Inc.
BENGTSSON, S., ANDERSSON, K. & FRIDELL, E. 2011. A comparative life cycle assessment of marine fuels;
liquefied natural gas and three other fossil fuels. Proceedings of the Institution of Mechanical Engineers,
Part M: Journal of Engineering for the Maritime Environment 225, 97-110.
BENGTSSON, S., FRIDELL, E. & ANDERSSON, K. 2012. Environmental assessment of two pathways towards
the use of biofuels in shipping. Energy Policy, 44, 451-463.
BENGTSSON, S. K., FRIDELL, E. & ANDERSSON, K. E. 2013. Fuels for short sea shipping: A comparative
assessment with focus on environmental impact. Proceedings of the Institution of Mechanical Engineers,
Part M: Journal of Engineering for the Maritime Environment.
BUHAUG, Ø., CORBETT, J. J., ENDRESEN, Ø., EYRING, V., FABER, J., HANAYAMA, S., LEE, D. S., LEE,
D., LINDSTAD, H., MARKOWSKA, A. Z., MJELDE, A., NELISSEN, D., NILSEN, J., PÅLSSON, C.,
WINEBRAKE, J. J. & WU, W., YOSHIDA, K. 2009. Second IMO GHG Study 2009. London, UK:
International Maritime Organization.
BÄCKSTRÖM, U. 2010. RE: E-mail conversation the 28th of April 2010 with Ulf Bäckström, Managing Director,
MAN Diesel Sverige AB, regarding estimated emissions from 6S60ME-C and S60ME-C-GI.
COOPER, D. & GUSTAFSSON, T. 2004. Methodology for calculating emissions from ships: 1 Update of
emission factors. Swedish Methodolgy for Environmental data (SMED). Norrköping.
COOPER, D. A. 2003. Exhaust emissions from ships at berth. Atmospheric Environment, 37, 3817-3830.

11
Low Carbon Shipping Conference, London 2013

COOPER, D. A. 2005. HCB, PCB, PCDD and PCDF emissions from ships. Atmospheric Environment, 39, 4901-
4912.
CORBETT, J. J. & WINEBRAKE, J. J. 2008. Emissions tradeoffs among alternative marine fuels: Total fuel cycle
analysis of residual oil, marine gas oil, and marine diesel oil. Journal of the Air and Waste Management
Association, 58, 538-542.
DETOLLA, J. P. & FLEMING, J. R. 1984. COMPUTER MODEL FOR SHIPBOARD ENERGY ANALYSIS.
Naval Engineers Journal, 96, 33-45.
DIMOPOULOS, G. G., GEORGOPOULOU, C. A. & KAKALIS, N. M. P. Modelling and optimization of an
integrated marine combined cycle system. ECOS, 2011 Novi Sad.
DINCER, I. & ROSEN, M. A. 2013. Exergy - Energy, environment, and sustainable development, Elsevier.
DOULGERIS, G., KORAKIANITIS, T., PILIDIS, P. & TSOUDIS, E. 2012. Techno-economic and environmental
risk analysis for advanced marine propulsion systems. Applied Energy, 99, 1-12.
DUPUIS, R. & NEILSON, J. 1997. Dynamic analysis: Sail the ship before it's built. Naval Engineers Journal,
109, 83-91.
EDWARDS, R., LARIVÉ, J.-F. & BEZIAT, J.-C. 2011. Well-to-wheels Analysis of Future Automotive Fuels and
Powertrains in the European Context WTT APPENDIX 1 - Description of individual processes and
detailed input data. Luxembourg: JEC - Joint Research Centre-EUCAR-CONCAWE collaboration.
FRIDELL, E., STEEN, E. & PETERSON, K. 2008. Primary particles in ship emissions. Atmospheric
Environment, 42, 1160-1168.
HAGSTRÖM, L. WP3 - Exhaust Gas Cleaning. Eff-ship final seminar, 2013 Gothernburg, Sweden.
IMO 2013a. MARPOL ANNEX VI AND NTC 2008 WITH GUIDELINES FOR IMPLEMENTATION, London, CPI
Group (UK) Ltd.
IMO 2013b. Sulphur monitoring programme for fuel oils for 2012. Note by the Secretariat. Marine Environmental
Protection Commitee (MEPC), 65th session, MEPC 65/4/9. London: International Maritime
Organization.
IPCC 2007. Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth
Assessment. Report of the Intergovernmental Panel on Climate Change. In: SOLOMON, S., D. QIN, M.
M., Z. CHEN, M. MARQUIS, K.B. AVERYT, TIGNOR, M. & MILLER, H. L. (eds.).
JOHNSEN, T. & MAGERHOLM-FET, A. 1998. Screening Life Cycle Assessment of M/V Color Festival.
KJØLHOLT, J., AAKRE, S., JÜRGENSEN, C. & LAURIDSEN, J. 2012. Assessment of possible impacts of
scrubber water discharges on the marine environment. København: Environmental Protection Agency,
Danish Ministry of Environment.
LARSEN, U., PIEROBON, L., HAGLIND, F. & GABRIELII, C. 2013. Design and optimisation of organic
Rankine cycles for waste heat recovery in marine applications using the principles of natural selection.
Energy.
MAGNUSSON, M., FRIDELL, E. & INGELSTEN, H. H. 2012. The influence of sulfur dioxide and water on the
performance of a marine SCR catalyst. Applied Catalysis B: Environmental, 111-112, 20-26.
MOTLEY, M. R., NELSON, M. & YOUNG, Y. L. 2012. Integrated probabilistic design of marine propulsors to
minimize lifetime fuel consumption. Ocean Engineering, 45, 1-8.
NIELSEN, J. B. & STENERSEN, D. 2010. Emission factors for CH4, NOX, particulates and black carbon for
domestic shipping in Norway, revision 1. Klima og forurensningsdirektoratet: Marintek.
NTM 2008. Environmental data for international cargo and passenger sea transport. Version 2008-10-18 ed.:
Nätverket för transporter och miljön.
SCIBERRAS, E. A. & NORMAN, R. A. 2012. Multi-objective design of a hybrid propulsion system for marine
vessels. IET Electrical Systems in Transportation, 2, 148-157.
SUBRAMANIAM, M. N., HAYES, C., TOMAZIC, D., DOWNEY, M. & BRUESTLE, C. 2011. Pre-turbo
aftertreatment position for large bore diesel engines - Compact & cost-effective aftertreatment with a
fuel consumption advantage. SAE Technical Papers.
TENG, H., REGNER, G. & COWLAND, C. 2007. Waste Heat Recovery of Heavy-Duty Diesel Engines by
Organic Rankine Cycle Part I: Hybrid System of Diesel and Rankine Engines. SAE Technical Paper.
THEOTOKATOS, G. & LIVANOS, G. A. 2013. Techno-economic analysis of single pressure exhaust gas waste
heat recovery systems in marine propulsion plants. Proceedings of the Institution of Mechanical
Engineers Part M: Journal of Engineering for the Maritime Environment, 227, 83-97.
WALSH, C. & BOWS, A. Incorporating lifecycle elements into the energy efficiency design index. Low Carbon
Shipping 2011 International Conference on Technologies, Operations, Logistics and Modelling, 22-24
June 2011, 2011 Glascow.
WEIDEMA, B. P., THRANE, M., CHRISTENSEN, P., SCHMIDT, J. & LØKKE, S. 2008. Carbon Footprint.
Journal of Industrial Ecology, 12, 3-6.
WINEBRAKE, J. J., CORBETT, J. J. & MEYER, P. E. 2007. Energy use and emissions from marine vessels: A
total fuel life cycle approach. Journal of the Air and Waste Management Association, 57, 102-110.
WINNES, H. & FRIDELL, E. 2009. Particle Emissions from Ships: Dependence on Fuel Type. Journal of the Air
& Waste Management Association, 59, 1391–1398.
WINNES, H. & FRIDELL, E. 2010. Emissions of NOX and particles from manoeuvring ships. Transportation
Research Part D: Transport and Environment, 15, 204-211.

12

View publication stats

You might also like